Abstract

The description of the formation, structure, and reactivity of coordination networks and metal–organic frameworks (MOFs) remains a real challenge in a number of cases. This is notably true for compounds composed of Zn2+ ions and terephthalate ligands (benzene-1,4-dicarboxylate, BDC) because of the difficulties in isolating them as pure phases and/or because of the presence of structural defects. Here, using mechanochemistry in combination with operando Raman spectroscopy, the observation of the formation of various zinc terephthalate compounds was rendered possible, allowing the distinction and isolation of three intermediates during the ball-milling synthesis of Zn3(OH)4(BDC). An “NMR crystallography” approach was then used, combining solid-state NMR (1H, 13C, and 17O) and density functional theory (DFT) calculations to refine the poorly described crystallographic structures of these phases. Particularly noteworthy are the high-resolution 17O NMR analyses, which were made possible in a highly efficient and cost-effective way, thanks to the selective 17O-enrichment of either hydroxyl or terephthalate groups by ball-milling. This allowed the presence of defect sites to be identified for the first time in one of the phases, and the nature of the H-bonding network of the hydroxyls to be established in another. Lastly, the possibility of using deuterated precursors (e.g., D2O and d4-BDC) during ball-milling is also introduced as a means for observing specific transformations during operando Raman spectroscopy studies, which would not have been possible with hydrogenated equivalents. Overall, the synthetic and spectroscopic approaches developed herein are expected to push forward the understanding of the structure and reactivity of other complex coordination networks and MOFs.
Introduction
In recent years, the use of mechanochemistry in material synthesis has undergone an exponential development. Ball-milling (BM) syntheses have notably proved to be of great interest in pharmaceutical materials screening, battery development, or even catalysis.1,2 More than offering a drastic reduction of bulk-solvent use in the syntheses, it offers a precise control over the stoichiometry of the reaction, as well as simple and straightforward synthetic procedures.3,4 However, because of the closed environment of mechanochemical reactions (especially when conducted in stainless steel jars), these syntheses are still considered as “black boxes”. In this context, the last decade has seen the emergence of in situ time-resolved analysis techniques applied to BM syntheses, which allow the evolution of the reaction medium to be monitored without needing to interrupt the milling process and open the jar. Among them, two are of particular interest: powder X-ray diffraction (pXRD) and Raman spectroscopy.5−11 While the first requires the use of a synchrotron beamline, the second can be set up in standard laboratories thanks to bench-top Raman spectrometers and probes. Such systems permit reaction rates to be followed, complex reaction schemes to be understood, and can also shed light on the appearance of yet unknown/unobserved reaction intermediates.12,13 Furthermore, these experimental installations can also be coupled with thermal imaging cameras to have an even more precise description of the course of the chemical reactions.14,15 It is worth mentioning that meticulous kinetic studies from operando Raman spectroscopy experiments have been recently proposed.16,17
Metal–organic frameworks (MOFs) represent a major family of porous materials, with porosities that can exceed 50% of their volume. The impressive range of their applications extends from fuel/gas storage to catalysis or even cancer therapy.18−20 Various synthetic routes have been developed for obtaining and/or shaping these materials, like solvothermal or microwave syntheses, and also mechanochemistry.21−24 The numerous properties of MOFs being strongly and directly related to their crystallographic structures, it is then of primary importance to precisely control their synthesis to obtain pure phases through meticulous and reproducible procedures.25 In this context, in situ monitoring experiments can help observe or even trap interesting intermediates. For example, recently, Jones et al. presented the use of in situ1H liquid-state NMR for observing the formation of a Ni-MOF.26 Insights into the nucleation and crystal growth processes of this MOF were obtained through the kinetic studies carried out at different temperatures. Concerning MOFs formed by mechanosynthesis, a recent example by Karadeniz et al. revealed the interplay between MOF-525 and its polymorph PCN-223, two porphyrinic zirconium MOFs. In their case, in situ pXRD monitoring allowed the MOF-525 to be detected as an intermediate during the formation of PCN-223, depending on the synthetic conditions used.27 A similar experimental procedure also allowed the presence of metastable MOFs in the formation of dia-Zn(MeIm)2 (MeIm: 2-methylimidazole) to be demonstrated.6
Among MOFs, those based on zinc (Zn-MOFs) have been the focus of much interest because of their high porosity and/or interesting applications. In particular, because zinc is a biocompatible metal, zinc-containing materials such as the porous scaffold Zn-xWE43 (WE43 being a Mg alloy) or the Zn2(BDC)2(diazabicyclooctane) MOF (with BDC standing for 1,4-benzenedicarboxylate) have been studied as candidates for implants and drug delivery.28,29 Furthermore, Zn-MOFs have shown great potential for applications as catalysts, anodes for lithium batteries, or materials for uranium adsorption/detection.30−32 Among the numerous Zn-MOFs described to date, coordination complexes based on Zn2+-dicarboxylate motifs have been found to exist as a very broad diversity of structures.33 For instance, the well-known structure of MOF-5 (Zn4O(BDC)3) exhibits 4-fold-coordinated zinc sites with bridging carboxylate functions, while MOF2 (Zn2(BDC)2) presents one 5-fold-coordinated zinc ion with four bridging oxygens from carboxylate ligands and one additional coordinated water molecule.34,35 Interconversion between these two structures has also been demonstrated in specific synthetic conditions (high pressure, temperatures up to 120 °C).36 Furthermore, the different investigations performed so far on coordination networks involving Zn2+ ions and terephthalate ligands (noted Zn-BDC from hereon) have enabled many other structures to be described, e.g., Zn(BDC)(H2O),37 Zn3(OH)4(BDC)·6H2O,38 and (Zn3(BDC)4·4H2O)n,39 just to name a few. Here, we decided to focus our attention on four other Zn-BDC structures: a dihydrated zinc terephthalate coordination polymer, Zn(BDC)(H2O)2,40 and three lamellar phases involving terephthalate and hydroxyl ligands (see Table 1).
Table 1. Zn-BDC Compounds Studied in This Contribution, Together with Structural Details, as Initially Reported in the Corresponding Refs (41−44).

The Zn(BDC)(H2O)2 phase (noted here ZTA1) consists of a “zig-zag” arrangement of chains formed by distorted tetrahedral zinc environments, in which each Zn2+ ion is linked to two monodentate BDC ligands and two water molecules (Table 1). Regarding the three lamellar structures, two are polymorphs of general formula Zn2(OH)2(BDC) (noted here as ZTA2a and ZTA2b). The structure of the latter polymorph (ZTA2b) has been more thoroughly described so far:43 it exhibits only one penta-coordinated zinc environment, with well-ordered π-stacked BDC units. The structure of the other polymorph (ZTA2a), a contrario, reveals more uncertainties. In fact, although several articles have reported the presence of this phase (by looking at pXRD data), no consensus has been found in the literature about its structure. Rodríguez et al. attributed this compound to a distorted MOF-5 phase caused by a default in one of the binding modes of carboxylate groups to Zn2+.45 Thirumurugan and Rao observed this phase as an intermediate in their hydrothermal reactions when using a Zn/BDC 1/1 molar ratio but were not able to propose an exact formula for it.46 Hirai et al. were unable to isolate a pure phase of this material but were able to detect it in a microwave synthesis when using a Zn/BDC ratio of 2/1 and working at specific pH (between 5.7 and 4.9).38 In 2016, a crystal structure was reported in the Cambridge crystallography database (CCDC ECATIO), which matches the pXRD pattern of this same phase and corresponds to the Zn2(OH)2(BDC) chemical formula (with a Zn/BDC ratio of 2/1). Unfortunately, no information on its synthesis could be found.42 When looking at the crystallographic file, it appears that it corresponds to a lamellar structure with two distinct zinc sites in octahedral geometry, which are separated by one bridging oxygen belonging to a hydroxyl (Table 1). Between two layers, the BDC ligands exhibit a “zig-zag” configuration of the aromatic rings. Hence, both polymorphs differ by subtle modifications of the BDC arrangements, resulting in different coordination modes of the metal centers, and the ZTA2a phase displays a slightly smaller interlamellar spacing than ZTA2b (about 10.1 vs 10.8 Å, respectively) (see Table 1). Lastly, the fourth phase of interest in this work is Zn3(OH)4(BDC) (noted ZTA3).44 This other lamellar Zn-hydroxide terephthalate is formed in a 3/1 Zn/BDC stoichiometry. Three distinct zinc environments were found to be present in the Zn–O polyhedra layers (with tetra/penta/hexacoordination, according to the initially reported crystallographic file), which are connected to each other through terephthalate and hydroxyl anions.
Diffraction techniques remain, at the moment, the most common analytical tool to obtain structural information on Zn-BDC MOFs. However, preparing good-quality crystals can be very challenging. Moreover, providing exact positions for hydrogen atoms from pXRD data remains an issue, and only two out of the four reported structures in Table 1 fully report hydrogen positions (aromatics and hydroxyls/water). In complement to XRD, chemical and structural information on these phases can also be obtained through FTIR spectroscopy. For example, the O–H stretching modes of hydroxyls and water molecules are easily observed above 3000 cm–1 and can inform on the H-bond networks in presence. Moreover, the C–O stretching modes of the carboxylate groups are identified by the presence of antisymmetric and symmetric vibrations (noted νas(COO–) and νs(COO–)) between 1650–1540 and 1450–1360 cm–1, respectively, and information on the binding mode of the carboxylate ligands to zinc metal centers can a priori be proposed on the basis of the difference in wavenumbers between these two frequencies (Δ = νas(COO–) – νs(COO–)).47,48 However, very few articles have used solid-state NMR (ssNMR) techniques to probe the local environment of the atoms in Zn-BDC structures, even for some of the more readily observable isotopes such as 1H and 13C.37,49−51 This can be seen as surprising considering that the use of ssNMR alongside XRD analyses is more and more frequent to elucidate the structure of MOFs and coordination polymers, including phases prepared using mechanochemistry,52−54 such as MOF crystal-glass composites (MIL-53/ZIF-62),55 cadmium-imidazole ZIFs,54 or even paramagnetic MOFs (e.g. Ni(II)-MOF-74, bimetallic ZnCu-MOF-74).56,57 More generally speaking, such techniques could help numerous debates, some of which will be discussed more extensively later in this contribution, which still remain about such zinc structures and their intermediates.
To further elucidate the details of the structure of these phases, 17O ssNMR is a very attractive tool. Oxygen-17 is a spin-5/2 quadrupolar nucleus of low natural abundance (0.04%), which presents a very wide chemical shift range (exceeding 1000 ppm) and also a very wide variation of quadrupolar interaction (QI) parameters (CQ, ηQ). Moreover, recent investigations have shown that 17O ssNMR spectroscopy can be of high interest for studying the structure of coordination polymers and MOFs.58−61 For example, Martins et al. have shown the powerful use of ultra-high-field 17O ssNMR experiments (35.2 T) applied on 17O-enriched Mg and Al MOFs (α-Mg3(HCOO)6 and MIL-53(Al)). Impressively, they were able to distinguish up to 12 inequivalent oxygen sites for as-made and activated phases of α-Mg3(HCOO)6.62 While 17O ssNMR is still not routinely used to investigate the structure of MOFs, our recent developments of cost-efficient 17O-labeling schemes using mechanochemistry,60,63,64 including for carboxylic ligands like terephthalic acid,65 imply that such nonroutine NMR now become conceivable for studying Zn-BDC structures like those in Table 1.
In this contribution, we report a detailed investigation of the structure and interconversion between the four aforementioned Zn-BDC systems (ZTA1, ZTA2a, ZTA2b, and ZTA3) synthesized using mechanochemistry. We first show how by modifying the initial stoichiometry of the reaction, and carefully monitoring the course of the BM reactions using operando Raman spectroscopy and thermal analysis of the milling jar, each of these different phases can be easily isolated in a pure form and subsequently studied by ex situ methods (FTIR, powder X-ray diffraction, multinuclear ssNMR), allowing further information about their structures to be gained. Moreover, we then demonstrate for the first time how, by using this synthetic methodology, it becomes possible to perform selective 17O isotopic labeling of the different types of oxygen environments in such MOF structures (by enriching either the water/hydroxyl ligands or the carboxylate ones), thereby opening the way to a level of structural insight, which had not been achieved before. Lastly, we show how the combination of high-resolution 17O NMR analysis of these compounds and ab initio calculations of NMR parameters using the gauge-including projector-augmented wave method (GIPAW) can improve and/or correct the structural models of these phases, thereby paving the way to other investigations on the structure/property relationships in complex coordination networks and MOFs.
Materials and Methods
Reagents
The following commercial precursors were used as received without further purification: terephthalic acid (C8H6O4, Janssen Chemicals, 98% purity, noted here H2BDC), deuterated terephthalic acid (C8D4H2O4, with deuteration on the aromatic H positions only, Cambridge Isotope Laboratories, 98% purity, noted here d4-BDC), ZnO (particle size < 5 μm, Sigma-Aldrich, 99.9% purity), and D2O (Sigma-Aldrich, 99.9% purity).17O-labeled water (with ∼40, 70, or ∼90% 17O-enrichment) was purchased from CortecNet.
17O-enriched H2BDC was prepared using two different mixer mills: either a vibratory mixer mill (Restch MM400 apparatus, as published elsewhere)65 or by scaling-up the procedure using a planetary ball mill (Fritsch P7 apparatus), as described below.
H2BDC-Enrichment in a Planetary Mixer Mill (Fritsch P7)
H2BDC (330 mg, 1.98 mmol, 1.0 equiv) and CDI (708.6 mg, 4.37 mmol, 2.2 equiv) were introduced into the stainless steel grinding jar (20 mL inner volume), to which 80 stainless steel beads (5 mm diameter) were then added. The jar was closed and subjected to grinding for two cycles of 10 min at 500 rpm, with a break of 30 min between each cycle in the P7 planetary ball mill. 17O-labeled (91.5%, 106 μL, 5.96 mmol, 3.0 equiv) water was then added into the jar, and the mixture was subjected to further grinding for two cycles of 10 min at 500 rpm, with a break of 30 min between each cycle. To help recover the product, nonlabeled water (2 mL) was added into the jar, and the content was subjected to grinding for 1 min at 500 rpm. Then, the suspension was transferred to a beaker (together with 9 mL of nonlabeled water used here to rinse the jar). The medium was acidified under stirring to pH ∼ 1 with an aqueous solution of HCl (6 M, 44 drops). The white precipitate was immediately filtered on a glass frit, washed (3 × 2 mL of 1 M HCl and then 3 × 2 mL of ultra-pure water), and then dried under vacuum. Yield: 312.7 mg (∼95%). Enrichment yield: 84%, enrichment level: 38.7%, as determined by the MS.
Mechanochemical Syntheses of Zn-BDC Networks
The Fritsch Pulverisette-23 (P23) vertical mixer mill was used in all reactions. BM syntheses were conducted in 10 mL inner-volume Perspex jars,7 with two 10 mm diameter zirconia beads.
A typical mechanochemical synthesis was performed as follows. First, ZnO was introduced into a Perspex jar followed by H2BDC (see Table 2 for exact masses). Both powders were gently mixed with a spatula before adding the two zirconia beads. Then, the appropriate amount of water was added. The jar was quickly closed, and parafilm was added to cover the junction of top/bottom parts of the jar to avoid any leaks as well as to increase air/water tightness. The system was then subjected to grinding for milling times ranging from 5 to 90 min in the P23 mixer mill operating at 30 or 50 Hz. The reactor was then opened, and the resulting compound was left to dry in air for several hours. The solid powder was subsequently recovered by scraping the edges of the reactor and beads with a spatula. All samples were obtained as white powders. After several uses, reactors can exhibit signs of erosion; however, no trace of Perspex was detected in the final compounds by 13C ssNMR and ex situ Raman spectroscopy.
Table 2. Synthetic Details for BM Experiments of Zn-BDC Compoundsa.
| milling parameters |
||||||
|---|---|---|---|---|---|---|
| code name | formula | mZnO (mg) | mH2BDC (mg) | VH2O (μL) | frequency (Hz) | duration (min) |
| ZTA1 | Zn(BDC)(H2O)2 | 129.9 | 265.1 | 115 | 30 | 30 |
| 1.59 mmol | 1.59 mmol | 6.38 mmol | ||||
| 1 equiv | 1 equiv | 4 equiv | ||||
| ZTA2a | Zn2(OH)2(BDC) | 178.4 | 182.1 | 40 | 50 | 20 |
| 2.19 mmol | 1.09 mmol | 2.22 mmol | ||||
| 2 equiv | 1 equiv | 2 equiv | ||||
| ZTA2b | Zn2(OH)2(BDC) | 197.8 | 201.9 | 350 | 50 | 90 |
| 2.43 mmol | 1.21 mmol | 19.43 mmol | ||||
| 2 equiv | 1 equiv | 16 equiv | ||||
| ZTA3 | Zn3(OH)4(BDC) | 262.1 | 178.4 | 310 | 50 | 90 |
| 3.22 mmol | 1.07 mmol | 17.21 mmol | ||||
| 3 equiv | 1 equiv | 16 equiv | ||||
| 17O-Labeled Compounds | ||||||
| ZTA1-*BDC | Zn(BDC)(H2O)2 | 34.0 | 71.0* | 17 | 50 | 5 |
| 0.42 mmol | 0.42 mmol | 0.94 mmol | ||||
| 1 equiv | 1 equiv | 2 equiv | ||||
| ZTA2a-*BDC | Zn2(OH)2(BDC) | 60.9 | 63.6* | 15 | 50 | 12 |
| 0.75 mmol | 0.37 mmol | 0.83 mmol | ||||
| 2 equiv | 1 equiv | 2 equiv | ||||
| ZTA2a-*OH | 178.4 | 182.1 | 40* | 50 | 20 | |
| 2.19 mmol | 1.10 mmol | 2.20 mmol | ||||
| 2 equiv | 1 equiv | 2 equiv | ||||
| ZTA2b-*OH | Zn2(OH)2(BDC) | 171.2 | 174.7 | 40* | 50 | 90 |
| 2.10 mmol | 1.05 mmol | 2.20 mmol | ||||
| 2 equiv | 1 equiv | 2 equiv | ||||
| ZTA3-*BDC | Zn3(OH)4(BDC) | 121.1 | 84.3* | 50 | 50 | 90 |
| 1.49 mmol | 0.49 mmol | 2.77 mmol | ||||
| 3 equiv | 1 equiv | 5 equiv | ||||
| ZTA3-*OH | 119.6 | 81.4 | 50* | |||
| 1.47 mmol | 0.48 mmol | 2.77 mmol | ||||
| 3 equiv | 1 equiv | 5 equiv | ||||
| Deuterated Compounds | ||||||
| ZTA1-d4BDC | Zn(d4BDC)(H2O)2 | 125.0 | 262.0d | 315 | 30 | 15 |
| 1.54 mmol | 1.54 mmol | 17.48 mmol | ||||
| 1 equiv | 1 equiv | 11.2 equiv | ||||
| ZTA2a-d4BDC | Zn2(OH)2(d4BDC) | 197.4 | 207.5d | 350 | 50 | 90 |
| 2.42 mmol | 1.22 mmol | 19.43 mmol | ||||
| 2 equiv | 1 equiv | 16 equiv | ||||
| ZTA2b-OD | Zn2(OD)2(BDC) | 188.1 | 192.8 | 335d | 50 | 90 |
| 2.31 mmol | 1.16 mmol | 18.47 mmol | ||||
| 2 equiv | 1 equiv | 16 equiv | ||||
| ZTA3-d4BDC | Zn3(OH)4(d4BDC) | 242.0 | 168.7d | 90 | 50 | 90 |
| 2.97 mmol | 0.99 mmol | 5.00 mmol | ||||
| 3 equiv | 1 equiv | 5 equiv | ||||
| ZTA3-OD | Zn3(OD)4(BDC) | 250.7 | 171.1 | 350d | ||
| 3.08 mmol | 1.03 mmol | 19.06 mmol | ||||
| 3 equiv | 1 equiv | 19 equiv | ||||
“*” is used for 17O-labeled materials and “d” is used for deuterated materials (prepared from D2O or d4-BDC precursors).
Further details on the amounts of reagents and milling conditions used to isolate each of the four phases, in their nonlabeled, 17O-labeled, or deuterated forms, are provided in Table 2. It is worth noting that for the synthesis of 17O-labeled compounds, amounts of reagents were reduced to further decrease the cost of obtaining an enriched material for ssNMR analyses.
“Routine” Characterizations of the Different Phases
Powder XRD analyses were performed on an X’Pert MPD diffractometer using Cu Kα1 radiation (λ = 1.5406 Å), with the operation voltage and current maintained at 40 kV and 25 mA, respectively. Diffractograms were recorded between 2θ = 5 and 50° in the Bragg–Brentano configuration, with a step size of 0.050° and a time per step of 60 s.
Infrared (IR) spectra were recorded on a Perkin Elmer Spectrum 2 FTIR spectrometer, using the ATR measurement mode (4000–400 cm–1 range, four repetitions).
Operando Raman Spectroscopy and Thermal Analysis
In this manuscript, we use the term operando to insist on the fact that spectroscopic measurements were acquired during the milling (and without stopping the milling), as opposed to in situ analyses, which can be performed on a medium when the milling is stopped for a few seconds/minutes.66
The Raman measurements were performed with a WP 785 ER Raman spectrometer (Wasatch Photonics) using an IPS (Innovative Photonic Solutions) laser source (working at ∼300 mW, wavelength of λ = 785 nm) and a contactless probe head (the probe’s focal length is 11 mm for a 1 mm diameter spot). The positioning of the Raman probe was optimized such that the focus of the laser was localized at the inner jar wall. A laser exposure time of 5 s with five accumulations was chosen, and spectrum collection was repeated every 30 s during the milling time. Raman spectra were collected across a range from 220 to 3120 cm–1, and a dark spectrum (laser OFF) was also recorded before starting the experiment and directly subtracted from the recorded spectra. No degradation of the jar and/or the compound by the laser was observed during the operando Raman experiments. The thermal imaging camera showed that during the milling, the laser beam did not prompt any additional heat around its focal point on the Perspex jar walls (compared to the jar’s overall temperature). Moreover, no noticeable degradation of the reagents and products was observed under the operating conditions used here for operando Raman analyses. The obtained Raman spectra were processed with the OriginPro 2021 program. Baseline correction (asymmetric least-squares smoothing) was first applied to each spectrum, and then normalization of the signal of the Perspex jar (at 1728 cm–1 where no overlapping from reagents nor product is expected) was performed, prior to the subtraction of the signal of the vibrating empty jar (see Supporting Information Figure S1). The reproducibility of the kinetic measurements performed is illustrated in Figure S2.
Temperature measurements were carried out using an OPTRIS PI450i thermal imaging camera. The optical resolution of the device is 382 × 380 pixels. The temperature accuracy measured using this system is estimated to be about ±2 °C. Four different temperatures were recorded every 2 s during the milling, focusing on three to four different areas, corresponding to the top, center, and bottom of the jar, and also to the “hot spot” at each time point.
Cautionary remarks: the laser source used here is designed as a class IV during all procedures of operation, maintenance, and service, meaning that safety measures must be taken accordingly. The use of laser-specific safety goggles is required when operating this laser, as failure to wear appropriate eye protection may result in permanent eye damage. Caution must be taken to never look directly into the laser beam and make sure the laser direction points toward the wall. The power range for the laser is 0–580 mW. Measurements presented in this contribution were performed at ∼300 mW in a dedicated isolated dark room with proper safety signs on the door, and operators received specific security training before using this equipment.
Solid-State NMR
1H ssNMR
1H ssNMR experiments were performed on a Varian VNMRS 600 MHz (14.1 T) NMR spectrometer, using a Varian 3.2 mm probe tuned to 1H (599.82 MHz). Spectra were recorded under magic angle spinning (MAS) conditions, with a spinning speed of 16 kHz and under a regulated airflow (0 °C). A 1H direct excitation pulse of 3.3 μs was used. Recycle delays used for the Zn-BDC compounds were 4 and 15 s for pure H2BDC. Additional 1H MAS NMR experiments were performed on a Bruker Avance NEO 850 MHz (20.0 T) NMR spectrometer, using a Bruker 1.3 mm probe, tuned to 1H (850.23 MHz), with a spinning speed of 60 kHz. Adamantane was used as a secondary reference for setting 1H chemical shifts (δiso = 1.8 ppm with respect to tetramethylsilane).
13C ssNMR
13C ssNMR experiments were performed on a VNMRS 600 MHz (14.1 T) NMR spectrometer, using a Varian 3.2 mm HX probe tuned to 1H (599.82 MHz) and 13C (150.81 MHz). All one-dimensional (1D) 13C NMR spectra consist of cross-polarization experiments (CP) under MAS conditions, with spinning frequencies from 4 to 15 kHz. A 1H excitation pulse of 3.3 μs was used prior to a ramped spin-lock pulse of 5 ms contact time. Acquisition was performed under spinal-64 1H-decoupling (∼75 kHz RF). Recycle delays used for the Zn-BDC compounds were 4 and 15 s for pure H2BDC. 13C chemical shifts were referenced to adamantane used as a secondary reference (high-frequency peak at 38.5 ppm with respect to tetramethylsilane).
17O ssNMR
17O ssNMR experiments were first performed on all samples on a Varian VNMRS 600 MHz (14.1 T) NMR spectrometer, using either Varian 3.2 mm HX or HXY probes, or a Phoenix 3.2 mm probe, tuned to 1H (599.82 MHz) and 17O (81.31 MHz). Spectra were recorded under MAS conditions under a regulated airflow (0 °C), with spinning speeds ranging from 16 to 20 kHz, depending on the sample. Additional 17O MAS NMR experiments were performed at lower field on a VNMRS 400 MHz (9.4 T) NMR spectrometer, using a Varian 3.2 mm HXY probe tuned to 1H (399.92 MHz) and 17O (54.21 MHz). Moreover, higher-field 17O NMR analyses were performed first on a Bruker Avance NEO 850 MHz (20.0 T) NMR spectrometer, using Bruker 1.3 or 4 mm probes tuned to 1H (850.23 MHz) and 17O (115.26 MHz) or on a Bruker Avance III HD 800 MHz (18.8 T) NMR spectrometer using a 3.2 mm Low-E probe spinning at 16 kHz (νL(1H) = 800.12 MHz and νL(17O) = 108.47 MHz) to perform multimagnetic field data fitting (see Supporting Information Figures S14–S17). A double frequency sweep (DFS) excitation scheme was used consisting of a 1 ms pulse with a sweep between 1 MHz and 50 kHz67−69 for 17O-signal enhancement (when possible) prior to the excitation pulse, and spinal-64 1H-decoupling was applied during acquisition. The acquisition parameters used for each sample (including number of scans and recycle delay D1) can be found in Table 3. The D-HMQC (dipolar-heteronuclear multiple-quantum coherence) sequence used an SR412 recoupling scheme on the 1H channel with a radio-frequency field of twice the spinning speed; the detection is performed on the 17O channel.7017O chemical shifts were referenced externally to D2O at −2.7 ppm (or tap water at 0.0 ppm).
Table 3. Main Acquisition Parameters for 17O NMR Experiments.
| sample | field (T) | Ørotor (mm) | expt | νrot (kHz) | D1 (s) | NS (/#t1) | νRF (kHz) | pulses (μs) | dec {1H} (kHz) |
|---|---|---|---|---|---|---|---|---|---|
| ZTA1-*BDC | 14.1 | 3.2 | DFS-One pulse | 18 | 2 | 1024 | ∼40 | 2 | 75 |
| 20.0 | 3.2 | DFS-One pulse | 14.286 | 4 | 256 | 10 | 8.33 | 60 | |
| ZTA2a-*BDC | 9.4 | 3.2 | DFS-One pulse | 18 | 1 | 20480 | ∼40 | 2 | 50 |
| 14.1 | 3.2 | DFS-One pulse | 18 | 1 | 6144 | ∼40 | 2 | 75 | |
| 18.8 | 3.2 | MQMAS | 16 | 5 | 768/22 | ∼16 | 3/1/5 | 62.5 | |
| 20.0 | 4 | DFS-One pulse | 14.286 | 1 | 4096 | 10 | 8.33 | 60 | |
| ZTA2a-*OH | 9.4 | 3.2 | DFS-One pulse | 18 | 2 | 512 | ∼40 | 2 | 50 |
| 14.1 | 3.2 | DFS-One pulse | 18 | 2 | 3072 | ∼40 | 2 | 75 | |
| 20.0 | 4 | DFS-One pulse | 14.286 | 2 | 256 | 10 | 8.33 | 60 | |
| ZTA2b-*OH | 9.4 | 3.2 | DFS-One pulse | 18 | 4 | 3072 | ∼40 | 2 | 50 |
| 14.1 | 3.2 | DFS-One pulse | 18 | 4 | 1024 | ∼40 | 2 | 75 | |
| 20.0 | 4 | DFS-One pulse | 14.286 | 4 | 512 | 10 | 8.33 | 60 | |
| ZTA3-*BDC | 9.4 | 3.2 | DFS-One pulse | 18 | 2 | 6144 | ∼40 | 2 | 50 |
| 14.1 | 3.2 | DFS-One pulse | 18 | 2 | 12288 | ∼40 | 2 | 75 | |
| 18.8 | 3.2 | MQMAS | 16 | 5 | 768/32 | ∼16 | 3/1/5 | 62.5 | |
| 20.0 | 4 | DFS-One pulse | 14.286 | 2 | 1024 | 10 | 8.33 | 60 | |
| 1.3 | (DFS) D-HMQC | 60 | 4 | 9216/24 | 20 | 4.167/8.33 | 75 | ||
| ZTA3-*OH | 9.4 | 3.2 | DFS-One pulse | 18 | 4 | 10240 | ∼40 | 2 | 50 |
| 14.1 | 3.2 | DFS-One pulse | 18 | 4 | 1024 | ∼40 | 2 | 75 | |
| 18.8 | 3.2 | MQMAS | 16 | 5 | 480/32 | ∼16 | 3/1/5 | 62.5 | |
| 20.0 | 4 | DFS-One pulse | 14.286 | 4 | 256 | 10 | 8.33 | 60 | |
| 1.3 | (DFS) D-HMQC | 60 | 2 | 1536/23 | 20 | 4.167/8.33 | 75 |
The NMR parameters were obtained by fitting the spectra using DMfit.71
NMR Conventions
In this article, the “Herzfeld–Berger” convention is used to describe the magnitude of the chemical shift anisotropy (CSA) arising from the magnetic shielding interaction.72 The isotropic chemical shift (δiso), span (Ω), and skew (κ) are given as follows
| 1 |
| 2 |
| 3 |
The principal components of the chemical shift tensor are ordered as follows: δ11 ≥ δ22 ≥ δ33.
The quadrupolar interaction is described by two parameters, the quadrupolar coupling constant (CQ) and the asymmetry parameter (ηQ)
| 4 |
| 5 |
Here, V11, V22, and V33 are the principal components of the traceless electric field gradient (EFG) tensor, with |V33| ≥ |V22| ≥ |V11|, e is the fundamental electronic charge, Q is the nuclear quadrupole moment, and h is Planck’s constant.
GIPAW-DFT Calculations
The unit cell parameters were set to the X-ray diffraction parameters and kept fixed during geometry optimizations to ensure consistency between experimental and optimized structures. Missing protons (in CCDC PUCYAO and PEKGAO structures) were initially positioned to be consistent with the expected structure of the system, and successively, only protons and then all atomic positions were relaxed with the VASP (Vienna ab initio simulation package) code73 based on the Kohn–Sham density functional theory (DFT) and using a plane-wave pseudopotential approach. For the two structures mentioned above, several initial OH proton positions were tested. The NMR parameters were then calculated within the Kohn–Sham DFT using the QUANTUM-ESPRESSO code,74,75 keeping the atomic positions equal to the values previously calculated with VASP. The PBE generalized gradient approximation76 was used, and the valence electrons were described by norm-conserving pseudopotentials77 in the Kleinman–Bylander form.78 The shielding tensor was computed using the gauge-including projector-augmented wave (GIPAW) approach,79 which enables the reproduction of the results of a fully converged all-electron calculation.80 The isotropic chemical shift δiso is defined as δiso = −[σ – σref], where σ is the isotropic shielding and σ ref is the isotropic shielding of the same nucleus in a reference system. For 1H, 13C, and 17O, previously used referencing was adopted using glycine,81 calcite,82 and a series of silicates,63 respectively. Diagonalization of the symmetrical part of the calculated tensor then provides its principal components σ11, σ22, and σ33 from which the chemical shift components δ11, δ22, and δ33 can be calculated. The principal components Vxx, Vyy, and Vzz of the electric field gradient (EFG) tensor are obtained by diagonalization of the tensor. The quadrupolar interaction can then be characterized by the quadrupolar coupling constant CQ and the asymmetry parameter ηQ, which are defined previously. The experimental value of the quadrupole moment of 17O (Q = −25 × 10–30 m2) was used to calculate CQ.83
Results and Discussion
Given the rich variety of structures involving Zn2+ and BDC ligands, in terms of Zn/BDC ratios, H2O or OH species linked to the Zn2+, and also carboxylate coordination modes, we decided to investigate the formation of these compounds using mechanochemistry, a synthetic method that had not yet been tested in a systematic way for these phases. The general idea was to see if conditions allowing each of these phases to be isolated in a pure form and high yield could be achieved using mechanochemistry. Coupled with the use of 17O-enriched precursors, this approach would allow all of the different forms of Zn-BDC phases to be analyzed with an unprecedented level of detail, thereby providing novel information in the debate related to the structures of these materials.
Formation and Isolation of Zinc (Hydroxyl-) Terephthalate Coordination Complexes
Observation of Several Intermediate Phases Using Operando Raman Analyses
The three reagents, namely, H2BDC, ZnO, and H2O, were introduced in a Perspex milling jar to enable the monitoring of different vibration bands by Raman spectroscopy. An illustrative example of the formation of various Zn-BDC phases in a one-pot mechanochemistry synthesis is shown in Figure 1 for a ZnO/H2BDC/H2O stoichiometry of 2/1/16. In Figure 1a, the evolution of the Raman signals during the milling for two selected regions of the spectra is presented after data processing (see the Materials and Methods Section for more information). It is worth noting that water plays multiple roles in this system: (i) it is directly involved in the reaction, as a reagent, by forming complexes with the Zn2+ cations and in some cases leads to the formation of coordinated hydroxyls, (ii) it plays the role of a liquid-assisted grinding (LAG) agent, allowing a better mixing of the solid reagents and thus a better efficiency of the milling process,84,85 (iii) it guarantees the formation of paste-like materials during the milling, which will be homogeneously spread over the inner surfaces of the milling jar. The latter point is of great importance for the operando Raman measurements, justifying the use of an excess amount of water in this particular example (≥16 equiv).
Figure 1.
(a) Top: starting compounds used for the operando experiment with their respective equivalents and milling conditions. Bottom: time-resolved operando Raman spectra of two selected areas (gray and orange, see panel (b)): 820–900 cm–1 range on the left and 1380–1480 cm–1 range on the right (see Figure S2 for repetitions of the analysis in this zone); two-dimensional (2D) plots have been overlapped with spectra from panel (b). (b) Experimental ex situ Raman spectra of H2BDC (in black), ZTA1 (in green), ZTA2a (in red), and ZTA2b (in blue) when isolated under the milling conditions reported in Table 2. (c) Experimental ex situ pXRD diffractograms of the corresponding “pure” compounds when isolated under the milling conditions reported in Table 2 (in black), and comparisons with the simulated pXRD data for these Zn-BDC phases (colored lines; pXRD simulated from available CIF files in the CCDC database, as reported in Table 1).
The system undergoes several structural transformations during the milling process, as illustrated by the shifts of specific vibration modes in the time-resolved experiment (Figure 1a). Between 800 and 900 cm–1 (Figure 1b, gray box), we focus on the aromatic breathing and out-of-plane (γCH) deformation region.86−88 The signal from the starting H2BDC molecule quickly shifts from 832 to 878 cm–1, and after ∼8 min of milling, it evolves to 862 cm–1, and finally moves to 865 cm–1 in the final compound. Accordingly, corresponding modifications of the νs(COO–) band can also be easily followed between 1380 and 1480 cm–1.
The products successively formed were identified as ZTA1, ZTA2a, and ZTA2b. Their Raman spectra and corresponding pXRD patterns are presented in Figure 1b,c in green, red, and blue, respectively. No formation of ZTA3 occurred when using an initial Zn/BDC molar ratio of 2/1 in the milling.
Kinetic Study of the Transformations
In view of describing further the reactions occurring during the milling, the kinetic evolution of the reaction medium established by Raman spectroscopy (as determined by the integration of the previously mentioned vibration modes between 800 and 900 cm–1, for example) was complemented with the thermal imaging of the milling jar (Figure 2). Simultaneous recording of temperatures at three different areas of the jar (referred to as top, center, and bottom) led to the plots displayed in Figure 2b. The inset (dashed gray area in Figure 2a) presents the first 500 s of the experiment, corresponding to the transformation of H2BDC into ZTA1 (from 832 to 878 cm–1), which has a 1/1 Zn/BDC molar ratio.40 This reaction is then followed by the introduction of an additional Zn2+ center in the structure and the deprotonation of two water molecules to form, after less than 8 min (<480 s) of milling, the ZTA2a phase (with characteristic bands at 862 and 1424 cm–1 in the two aforementioned regions), which has a Zn/BDC ratio of 2/1 (red curve). The reaction is probably exothermic, as shown by the increase of the temperature up to ∼40 °C (brown curve in Figure 2b). The final transformation observed in Figure 1a corresponds to the structural modification of ZTA2a to form the polymorph ZTA2b (865 and 1449 cm–1), with a second increase of the temperature up to ∼47 °C (for the center temperature). Such temperature behaviors have been observed in various other systems during milling syntheses.14,89Figure 2b exhibits, for comparison, the milling temperature profile of a mixture of H2BDC (400 mg) and H2O (350 μL) (without ZnO but using similar milling conditions as in Figure 1a), allowing a difference of about 10 °C to be highlighted after ∼15 min of milling (and >15 °C after 1 h) between the Zn-BDC synthesis and the control experiment (brown vs gray curves).
Figure 2.
(a) Reaction profiles corresponding to signals highlighted in Figure 1b. Empty symbols (ZTA2a and ZTA2b at 1424 and 1449 cm–1, respectively) are shown only for the last half of the milling time for clarity. The inset on the right-hand side corresponds to a zoom of the first 500 s of the time-resolved Raman experiment. (b) Temperature profiles of three different areas of the jar (top, center, and bottom); the light gray curve corresponds to a control experiment in which the temperature evolution caused by the milling of pure H2BDC with H2O in the same experimental conditions as in Figure 1a was followed. (c) Picture obtained with the thermal imaging camera of the milling jar, highlighting the three measurement zones. The shaking direction is emphasized by the orange arrows.
The fact that the apparent proportion of ZTA2a is not null at the end of the reaction in Figure 2a (despite that its resonances have disappeared from the Raman data) can be explained by the way in which these kinetic plots were fitted here in the 860 cm–1 region using simple Gaussian signals centered at the different maximum frequencies of the intermediates, with intensities that were then left to freely evolve at each time point. This created ambiguities in the fitted intensities due to the closely overlapping signals of ZTA2a and ZTA2b (only ∼3 cm–1 difference in this spectral region). More accurate kinetic plots related to the transformation of ZTA2a into ZTA2b could be obtained by fitting the 1380–1480 cm–1 region, where the Raman resonances of the two phases are more distinctly resolved. This is shown in Figure 2a, where the Raman signal from ZTA2a (red empty circles) decreases close to 0% at the end of the reaction (in line with pXRD analyses of the final compound). Because the purpose of this manuscript was not to establish a full kinetic model of the formation of these Zn-BDC phases, no attempt was made to further refine or analyze this operando Raman data. Yet, it is worth noting that this example illustrates the importance of having access to different regions of the Raman spectra to evaluate the quality of such kinetic fits and potentially resolve overlapping resonances. Other possibilities along this line (e.g. with deuterated precursors) will be discussed at the end of this article.
In the frame of this work, the clear identification of each transformation thanks to the operando setup involving the combined use of the Raman instrument and the thermal camera allowed each of these structures to be isolated as a pure phase for further ex situ structural analyses (using the optimized reaction conditions reported in Table 2). Similarly to ZTA1, ZTA2a, and ZTA2b, the formation of ZTA3 could also be followed in the same fashion by operando Raman spectroscopy when starting from a 3/1 ratio between the ZnO and H2BDC precursors. Hence, by choosing the adequate stoichiometry and by stopping the milling at carefully selected times, the different Zn-BDC phases could be isolated pure, without them changing with time (see Supporting Information Figure S3). It then became conceivable to refine their structures, especially the proton positions, and to gain more insights into the terephthalate binding modes.
Refinement of the Zn-BDC Structures Using an “NMR Crystallography” Approach
When the limits of XRD methods are reached, for example, for phases that cannot be isolated as single crystals, in which some atoms cannot be positioned (e.g. protons) or which contain defects, ssNMR methods can be particularly useful when used in conjunction with computational modeling to gain precise information on the crystallographic structures. Such an approach, referred to as “NMR crystallography”,90,91 implies that the structures of interest possess NMR-active isotopes, which can be readily analyzed by high-resolution NMR.
In the case of Zn-BDC structures, the vicinity of the Zn2+ cations and more specifically the precise coordination modes of the BDC and hydroxyl ligands to Zn2+ are of great importance. From an NMR perspective, all atoms present in these phases have at least one NMR-active isotope, which can be studied by NMR, making them potentially attractive targets of analysis: 1H (I = 1/2, 100% natural abundance), 13C (I = 1/2, 1.07% natural abundance), 17O (I = 5/2, 0.04% natural abundance), and 67Zn (I = 5/2, 4.11% natural abundance). Even though 67Zn NMR has recently shown promising results for providing a detailed description of zinc local environments, it is still considered as very challenging92 because of the quadrupolar nature of zinc-67 (which can lead to very broad lines)93 and its poor receptivity (low natural abundance and very low resonance frequency (νL(67Zn) < 10% νL(1H))).92,94 In the present work, we thus set our focus on the other three nuclei (13C, 17O, and 1H) to reach a more detailed description of the binding mode of the dicarboxylate ligands and for positioning the H-atoms of the hydroxyls within the structures.
13C NMR Analyses of Carboxylate Binding Modes in Zn-BDC Structures
As briefly explained earlier (vide supra), IR spectroscopy has been used for decades as a first approach for proposing binding modes of carboxylates to various metal ions. The difference in wavenumbers between the antisymmetric and symmetric stretching modes, defined by the parameter Δ (=νas(COO–) – νs(COO–)), has been shown to change depending on the binding mode of carboxylates, allowing, as a first approximation, the following tendency to be derived upon coordination to divalent metal ions: Δ(Chelating) < Δ(Bridging) < Δ(Ionic) < Δ(Monodentate) (see Figure 3a).47 However, this method can be ambiguous as Δ can be affected by other parameters, depending on the types of ligands involved and how they interact with each other within the structures. Moreover, it relies on the clear attribution of the vibration modes (νas(COO–) and νs(COO–)), which actually remains uncertain in many experimental cases, due to numerous overlapping vibration bands (e.g., νas(COO–), νs(COO–), ν(CC), β(CCH), etc.) on the IR spectra (see Supporting Information Figure S4). In the present case, when looking at the FTIR spectrum of ZTA2a, the Δ = 210 cm–1 splitting is significantly higher than expected for bridging ligands. Indeed, the other phases with bridging ligands, namely, ZTA2b and ZTA3, display values at Δ = 171 and 180 cm–1, respectively. Moreover, the monodentate bound ligand in ZTA1 shows a splitting of Δ = 192 cm–1. Overall, this tends to show the limitations of IR spectroscopy in providing direct information on the terephthalate binding modes, but may also hint that ZTA2a actually exhibits more complex binding modes than suggested by the X-ray diffraction data.
Figure 3.
(a) Different possible binding modes of carboxylate ligands to Zn2+ ions. (b) Left: 13C CP MAS NMR spectra recorded at 14.1 T with a spinning speed of 15 kHz under a regulated temperature (0 °C) (in black). Right: 13C CP MAS NMR spectra recorded at 14.1 T with a spinning speed of 4.0 or 4.2 kHz under a regulated temperature (0 °C) (in black). Black-filled squares indicate the isotropic chemical shifts. The parameters used for fitting the spectra (in red) are gathered in Table 4.
Here, we decided to perform 13C ssNMR analyses to obtain more information on the binding mode of the carboxylates. The 13C ssNMR spectra of the four samples obtained by mechanochemistry are presented in Figure 3b (see Supporting Information Figure S5 for data on H2BDC). Spectra were recorded with different spinning speeds, such as to focus on the isotropic chemical shifts (Figure 3b, left) or to derive the chemical shift anisotropy (CSA) parameters (Figure 3b, right).
Based on the reported crystal structures (see Table 1), two main binding modes were expected for the Zn-BDC compounds of interest: monodentate and bridging. For ZTA1, BDC ligands exhibit a monodentate configuration, and only one crystallographically inequivalent carboxylate carbon is present in the structure. For this phase, a single resonance was observed in the carboxylate region of the 13C ssNMR spectrum (Figure 3b), with δiso(13COO) = 176.6 ppm. On the other hand, ZTA2b and ZTA3 structures have different types of bridging terephthalate ligands based on the reported X-ray diffraction data. While the former clearly exhibits a single carboxylate environment (δiso(13COO) = 172.1 ppm), as expected from the crystal structure, the latter shows two main carboxylate resonances centered at 174.0 and 172.1 ppm (with an additional shoulder in the low-frequency area, which could perhaps be due to the presence of domains within the structure, in which terephthalate ligands exhibit very slight changes in symmetry/local environment). Finally, for the ZTA2a phase, while only one carbon signal was expected from the crystallographic structure, two carboxylate resonances were observed, centered at 175.7 and 172.5 ppm, and present in an ∼5/1 ratio. The weaker low-frequency resonance was systematically observed for this phase (for different synthetic batches) and with the same approximate ratio. Although its 13C chemical shift is relatively close to that of the H2BDC precursor (173.5 vs 172.5 ppm), no traces of residual H2BDC could be detected by IR spectroscopy in the C=O stretching region and no acidic proton signal was observed by 1H ssNMR, meaning that this small signal cannot arise from the presence of residual H2BDC (see Supporting Information Figures S6 and S7). Moreover, given that no diffraction peaks ascribable to known impurities could be detected on the X-ray diffraction pattern of ZTA2a (see Figure 1c and Supporting Information Figure S3), the weak resonance must correspond to systematic defects in the crystal structure, which could not be resolved nor identified by X-ray diffraction.96
In the literature, a few articles report 13C ssNMR data for zinc-carboxylates and more specifically Zn-BDC phases.37,49−51 In all cases, δiso(13COO) was found to vary between 171.8 and 177.7 ppm. Moreover, when comparing the reported δiso(13C) values of bridging carboxylates versus monodentate ones, a trend can be extracted. Indeed, the δiso(13COO) values of monodentate ligands are nearly systematically more deshielded than the δiso(13COO) of bridging ones. For example, Yuan et al. reported values of δiso(13COO) = 176.7 and 177.3 ppm for two Zn-BDC phases exhibiting monodentate binding modes (the first phase was actually the same as ZTA1, while the second one was Zn(BDC)(H2O)), and for a Zn-BDC·DMF MOF with bridging terephthalates, they reported δiso(13COO) = 172.7 ppm.50 Similarly, Julien et al. showed that the Zn(H2O)2(2,5-dihydroxyterephthalate) structure, which displays a monodentate configuration, exhibits a 13C resonance at δiso(13COO) = 176.0 ppm, while MOF-74, with bridging 2,5-dihydroxyterephthalate ligands, shows signals at δiso(13COO) = 173.2 and 173.9 ppm (for compounds obtained by LAG using either H2O or DMF, respectively).51 Habib et al. also studied two Zn-dicarboxylate complexes, [Zn2(μ2-BDC)2(μ4-btre)] and [Zn2(μ3-ip)2(μ2-btre)(H2O)2]·2H2O (with btre = 1,2-bis(1,2,4- triazol-4-yl)ethane, ip = benzene-1,3-dicarboxylate). Each structure exhibits two distinct δiso(13COO) by 13C ssNMR, corresponding to two monodentate binding modes (δiso(13COO) = 174.4 and 173.3 ppm) for the first one, while the second complex has a bridging and a monodentate ligands (δiso(13COO) = 174.6 and 176.7 ppm).49 In line with this literature survey, which shows that the δiso(13COO) values for monodentate binding modes are globally deshielded compared to those for bidentate ones, it was found that the isotropic 13C shifts of the carboxylates in ZTA1, ZTA2b, and ZTA3 also appeared to follow this trend. In contrast, for ZTA2a, there seemed to be discrepancies with what one would have expected from the reported crystal structure, not only in the number of 13C resonances (due to the additional small peak—vide supra) but also in the 13C isotropic shift of the main resonance, which is rather high for a “bridging mode”. Hence, overall, this points once again to the issues related to the X-ray data available for this phase.
To go further in the study of these, we conducted various DFT optimizations of the different structures studied here. Starting from the reported crystal structures, H-atoms were positioned (when missing). Then, keeping cell parameters constant, geometry optimizations were performed by optimizing either H positions only or all atomic positions. The H-relaxed structures did not lead to satisfactory results after the computation of NMR parameters, except for ZTA1 (as illustrated for δiso(13COO) in Table 4). A more complete geometry optimization was thus needed to refine the various crystal data. Moreover, for some of the structures, it was necessary to test various initial H orientations for the hydroxyls before geometry optimization to converge toward a model that matches best the experimental data; this will be further detailed for ZTA3 later in this article. In all cases, after these additional geometry optimization steps, the GIPAW-calculated NMR parameters for 13C were found to be in relatively good agreement with the experimentally extracted values. Notably, δiso(13COO)calc of the monodentate ZTA1 ligand was the most deshielded of the series, with a calculated 13C chemical shift at 181.8 ppm, while δiso(13COO)calc from bridging ligands were ∼6–10 ppm lower in frequency. Moreover, the calculated chemical shift anisotropy (CSA) parameters after full geometry optimization were also found to be in relatively good agreement with the experimental values, as further discussed below. These DFT-optimized structures can thus be considered as more reliable structural models of the four Zn-BDC phases studied here (see Supporting Information Figure S8).
Table 4. 13C NMR Parameters of Carboxylates Extracted from Fitting Spectra Recorded at Multiple Spinning Speeds and Calculated Ones Using DFT from Optimized Structural Models of the Phases.
| experimental | calculated (GIPAW-DFT)a | ||||||
|---|---|---|---|---|---|---|---|
| H-relaxed | fully optimized structures |
||||||
| compound | δiso (ppm) | Ω (ppm) | κ | δiso (ppm) | δiso (ppm) | Ω (ppm) | κ |
| ZTA3 | 172.1 ± 0.3 | 136.9 ± 0.4 | 0.05 ± 0.01 | 167.6 | 172.4 | 130.7 | 0.15 |
| 174.0 ± 0.1 | 139.9 ± 2.4 | 0.13 ± 0.01 | 168.0 | 175.6 | 132.3 | 0.31 | |
| ZTA2b | 172.1 ± 0.1 | 137.4 ± 0.2 | 0.22 ± 0.01 | 168.0 | 173.4 | 136.2 | 0.15 |
| ZTA2a | 172.5 ± 0.2 | 140.0 ± 8.2 | 0.34 ± 0.15 | ||||
| 175.7 ± 0.1 | 137.5 ± 0.4 | 0.20 ± 0.01 | 199.4 | 177.6 | 141.7 | 0.23 | |
| ZTA1 | 176.6 ± 0.1 | 154.7 ± 6.3 | 0.03 ± 0.02 | 176.5 | 181.8 | 146.4 | 0.35 |
A systematic study of 13C chemical shifts was performed in 2009 on 14 molecules (aromatics and carbohydrates) whose crystal structures have been precisely determined by neutron diffraction. This study showed that the GIPAW 13C predictions are more accurate than a cluster approach, with an error on the isotropic shift of ∼1 ppm.95
Following this, the structures after full relaxation of atom positions were analyzed in detail, and relationships between calculated 13C NMR parameters and local environments of the carboxylates (in the geometry-optimized structures) were looked into (see Supporting Information Figure S9). Significant structural modifications were observed upon full geometry optimization of some of the phases. For example, the ZTA3 crystallographic structure was found to undergo various changes, the major one being the modification of one zinc coordination from a 5-fold O-coordination environment (Zn2 in Table 1) to a 4-fold one. In the case of ZTA2a, significant variations in the carboxylate binding modes were observed during the relaxation of all atom positions, which can explain the ∼20 ppm decrease in the δiso(13COO)calc value between H-relaxed and fully relaxed models (see Table 4). Still for this phase, the calculated 13C NMR parameters were found to be in good agreement with those of the most intense 13C resonance (i.e., δiso(13COO)exp = 176.4 ppm vs δiso(13COO)calc = 175.7 ppm), again with a relatively high chemical shift compared to what one may have expected for bridging carboxylates. Yet, when looking more closely at the local environment of the carboxylate after full geometry optimization, the coordination mode was actually found to be closer to a bridging syn,syn/anti configuration (see Figure 3a) than to monodentate or bridging syn,syn ones. In this configuration, one of the dZn···O distances was found to be shorter (in the range of dZn···O for syn,syn bridging modes, i.e., ∼2.0 Å) than the two others (which were measured between 2.2 and 2.3 Å). This may be one of the reasons for the higher 13C shift of this phase, compared to bridging modes involving only shorter Zn···O bonds.
To go even further in trying to relate 13C NMR parameters to structural/geometric information about carboxylate bindings, the calculated data collected from the numerous geometry optimizations on Zn-BDC phases (especially the ZTA3 structure) were compiled. The most obvious trend we found was between the δiso(13COO)calc and the longest C–O distance of the carboxylate, (dC–O)l: the highest chemical shifts are clearly related to structures with the longest C–O distances (dC–O)l (see Supporting Information Figure S9). A contrario, no unambiguous relationship relating δiso(13COO) and the type of carboxylate coordination could be extracted, showing that this parameter actually depends on several factors.
As 13C CSA parameters can also be used to describe the local environment of carbon atoms, these were extracted from the experimental 13C NMR spectra recorded at two different spinning speeds (Table 4). The most noticeable difference between compounds was in the CSA span (Ω), which was found to be experimentally higher by approximately 15–20 ppm for ZTA1 (monodentate binding) than for the carboxylates of the three other compounds (155 vs 135–140 ppm, respectively) (see Supporting Information Figure S10 and Table S1). DFT-calculated values of spans also followed the observed trend, the highest value being for ZTA1, with, however, the calculated span of ZTA2a now only about 5 ppm lower than that calculated for ZTA1. Using the data library obtained by optimizing the Zn-BDC structures and also different structural models of ZTA3, attempts were made to relate the various CSA parameters stemming from the calculations to carboxylate binding modes. Here, a general trend was also observed between the longest C–O distance (dC–O)l and Ωcalc (see Supporting Information Figure S11), but no direct relationship with the carboxylate coordination.
Overall, although hints about carboxylate local environments could be derived from 13C ssNMR spectra, no complete insight into the carboxylate coordination mode could be obtained. Hence, we then considered analyzing local environments of oxygen atoms because they are directly coordinated to the Zn2+ ions in the Zn-BDC structures and may help gain insights into the various binding modes.
17O NMR Analyses of Carboxylate and Hydroxyl Binding Modes in Zn-BDC Structures
With the aim of directly probing the binding modes of the terephthalate ligands by 17O ssNMR, 17O-labeled terephthalic acid was used as a starting material to prepare Zn-BDC phases enriched on the carboxylate groups. Concerning the bridging hydroxyls, a direct enrichment during the mechanochemical synthesis of the Zn-BDC phases was envisaged using 17O-enriched water as a starting reactant (instead of normal water).
Both labeling strategies were tested for each compound (ZTA1, 2a, 2b, and 3) to isolate Zn-BDC phases selectively enriched on the carboxylates or on the hydroxyls/water and then perform high-resolution NMR spectroscopy. All samples could be prepared by mechanochemistry, with selective enrichment on the carboxylate part (phases noted ZTA-*BDC) or the hydroxyl/water part (phases noted ZTA-*OH), with the exception of ZTA2b, for which the carboxylate-enriched phase (ZTA2b-*BDC) could not be obtained with sufficient purity to be analyzed by 17O ssNMR. In total, six different enriched samples were thus isolated with good purity, using the BM synthetic procedure developed above thanks to the operando Raman setup. In all cases, preliminary evidence of the success of the enrichment could be assessed by FTIR, with the slight shifts of vibration modes such as νas(COO–) and νs(COO–) (see Supporting Information Figure S4c). The corresponding 1D 17O MAS NMR spectra are presented in Figure 4, which could all be recorded with good sensitivity in just a few hours.
Figure 4.
17O MAS NMR spectra of 17O-labeled Zn-BDC compounds prepared by mechanochemistry, B0 = 14.1 T, νrot = 18 kHz, and T = 0 °C. (a) ZTA-*OH samples corresponding to the enrichment of the hydroxyl/water groups. (b) ZTA-*BDC samples corresponding to the enrichment of the BDC moiety. For all spectra, experimental spectra are in black, and the red dotted line corresponds to the results of the fitting process.
Because oxygen-17 is a quadrupolar nucleus, the NMR resonances recorded on the 1D spectra show characteristic second-order quadrupolar line shapes. To extract the 17O NMR parameters associated with each oxygen environment (δiso, CQ, and ηQ), spectra were fitted at multiple fields (see Supporting Information Figures S14–S17), since the second-order quadrupolar broadening decreases as the magnetic field increases. Moreover, additional 2D high-resolution MQMAS (multiple-quantum MAS) experiments were also recorded for some of the phases, allowing the fits to be further refined (see Supporting Information Figures S12 and S13).
Regarding the ZTA-*OH phases labeled on the hydroxyl/water positions, the NMR spectra were found to be generally easier to analyze (Figure 4a). Only one *OH site was observed for both ZTA2a-*OH and ZTA2b-*OH, as expected from the published crystal structures. In contrast, ZTA3-*OH was found to exhibit a more complex line shape consisting of several inequivalent sites (Figure 4a). These were resolved using a 17O MQMAS NMR experiment (see Supporting Information Figure S12). Moreover, the OH network of this compound was further investigated through a 1H–17O dipolar correlation experiment, as detailed below (see Figure 5).
Figure 5.
(a) 1H–17O D-HMQC of ZTA3-*OH recorded at 20.0 T under a spinning speed of 60 kHz. The 17O and 1H 1D spectra recorded in the same condition as the D-HMQC are shown in blue. Extracted slices (in black) are fitted (red lines) with the experimental parameters presented in Table 5. The inset displays the FTIR spectrum centered on the OH stretching region. H-bonded OH groups are highlighted in orange, and isolated OH groups are shown in green. (b) Close-up views of the DFT-optimized ZTA3 structure of the different hydroxyl environments; color code: brown for carbon, red for oxygen, light pink for hydrogen, and purple for zinc. (c) Comparison of experimental 1H MAS spectrum (recorded at B0 = 20.0 T and νrot = 60 kHz, in black) and simulated spectrum from calculated 1H NMR parameters in dotted gray.
The D-HMQC experiment allows 1H···17O proximities to be probed and turned out to be highly useful to help position the hydroxyl protons (which were absent from the initial X-ray diffraction data). Here, a short recoupling time was used (τrec = 100 μs) to observe the shortest O···H contacts (i.e., the O–H bonds). In these conditions, no correlation between hydroxyl oxygen atoms and aromatic protons was observed (blue-shaded zone) but four environments could be resolved for the hydroxyls, one per inequivalent OH, in line with the published crystal structure (three being μ3-OH and one being μ2-OH). More interestingly, when looking at the FTIR spectrum of ZTA3 (see the inset of Figure 5a), four environments of the hydroxyls could also be detected and separated into two groups: two for isolated OH species (green-shaded zone) and two for OH species involved in H-bonding (orange-shaded zone). A similar distinction can be made using the D-HMQC data. Indeed, the most shielded oxygen site (δiso(17O) = −27.9 ppm) and the most deshielded one (δiso(17O) = 12.8 ppm) are correlated to 1H resonances with low chemical shift (close to 0 ppm), such 1H shifts being characteristic of isolated OH groups (zone shaded in green).97,98 On the other hand, the two remaining OH sites are correlated to 1H resonances, with δiso(1H) close to 5 ppm, implying that they are involved in hydrogen bonds (orange-shaded zone).87−99
Using these high-resolution 1H–17O experiments as a guideline, different configurations were tested to position the four hydroxyl hydrogens in the ZTA3 structure, which were then geometry-optimized by DFT. The resulting GIPAW-DFT-computed 1H ssNMR spectrum of the best model is highly comparable to the high spinning speed 1H NMR spectrum (Figure 5c). The resulting refined ZTA3 structure exhibits two isolated hydroxyls: one (O5-H13) is shared between Zn3–Zn2–Zn3 (μ3-OH), while the second one (O7–H10) is located between Zn1–Zn2 (μ2-OH); both display a low δiso(1H) chemical shift (1.34 and 0.80 ppm, respectively). In contrast, the hydroxyl protons denoted H11 and H12 have a higher chemical shift, indicative of the presence of a hydrogen-bonding network, as can be seen in the geometry-optimized structure (see Figure 5b): the H11···O7 and H12···O2 distances were found to be only ∼1.7 Å.
Generally speaking, the calculated 17O NMR parameters were found to be in satisfactory agreement with those extracted from experimental data. The overestimation of CQ, often observed in Table 5, is consistent with what has been recently reported in the 17O NMR literature on MOFs.59,100 However, unlike the calculated 13C NMR parameters, no clear trend between 17O NMR parameters and structural geometric parameters (dC–O, dZn···O, θZn···O–C...) could be derived (see Supporting Information Figures S18 and S19). This is probably due to the multifactorial dependence between the 17O NMR parameters and the local environment and geometry around the oxygen in these structures.
Table 5. 17O ssNMR Parameters Extracted from Experimental Spectra (Figure 4) Recorded at Multiple Magnetic Fields and Calculated Using DFT from Structural Models of the Phases.
| experimental | calculated | |||||||
|---|---|---|---|---|---|---|---|---|
| enrichment | compound | δiso(17O) (ppm) | CQa (MHz) | ηQ | Int (%) | δiso(17O) (ppm) | CQ (MHz) | ηQ |
| *OH | ZTA3 | 12.8 ± 0.6 | 5.59 ± 0.02 | 0.30 ± 0.02 | 27.2 | 6.1 | –7.88 | 0.64 |
| –1.9 ± 0.1 | 4.71 ± 0.04 | 0.43 ± 0.10 | 17.9 | 3.7 | 6.81 | 0.98 | ||
| –8.5 ± 0.1 | 5.22 ± 0.01 | 0.23 ± 0.02 | 23.6 | 0.4 | –4.31 | 0.53 | ||
| –27.9 ± 0.2 | 5.92 ± 0.02 | 0.47 ± 0.01 | 31.3 | –27.3 | –7.98 | 0.51 | ||
| ZTA2b | –5.6 ± 0.4 | 5.10 ± 0.10 | 0.36 ± 0.01 | 100 | –9.9 | –5.92 | 0.33 | |
| ZTA2a | 5.5 ± 1.3 | 5.90 ± 0.12 | 0.25 ± 0.02 | 100 | 1.8 | –6.89 | 0.34 | |
| ZTA1 | –3.2 ± 3.0 | 7.66 ± 0.20 | 0.61 ± 0.04 | 100 | –15.8 | 9.32 | 0.49 | |
| *BDC | ZTA3 | 258.4 ± 0.1 | 7.23 ± 0.01 | 0.56 ± 0.01 | 28.5 | 276.3 | 7.73 | 0.80 |
| 249.0 ± 0.7 | 6.97 ± 0.02 | 0.74 ± 0.03 | 25.3 | 273.4 | 7.93 | 0.84 | ||
| 237.3 ± 0.5 | 7.06 ± 0.01 | 0.75 ± 0.06 | 19.9 | 254.7 | –6.77 | 0.97 | ||
| 205.2 ± 0.8 | 6.39 ± 0.05 | 0.90 ± 0.10 | 26.3 | 210.2 | –7.26 | 0.63 | ||
| ZTA2a | 278.3 ± 1.6 | 7.07 ± 0.10 | 0.51 ± 0.11 | 32.3 | 298.1 | 7.37 | 0.69 | |
| 261.3 ± 1.0 | 7.04 ± 0.02 | 0.56 ± 0.13 | 10.5 | |||||
| 255.7 ± 1.5 | 6.70 ± 0.21 | 0.88 ± 0.12 | 5.2 | |||||
| 240.3 ± 5.3 | 6.74 ± 0.30 | 0.85 ± 0.15 | 8.7 | |||||
| 223.0 ± 1.1 | 7.85 ± 0.05 | 0.60 ± 0.08 | 43.3 | 253.0 | 8.79 | 0.62 | ||
| ZTA1 | 286.6 ± 1.3 | 6.90 ± 0.05 | 0.46 ± 0.02 | 50.9 | 320.6 | 7.46 | 0.69 | |
| 221.7 ± 2.1 | 5.83 ± 0.49 | 0.85 ± 0.01 | 49.1 | 251.3 | –6.71 | 0.67 | ||
Experimental CQ values are given in absolute values.
Concerning the phases enriched on the carboxylates, the experimental δiso(17O) arising from the BDC* ligands of the different structures were found to span between 205.2 and 286.6 ppm (see Figure 4b). ZTA1-*BDC exhibits two clear inequivalent oxygen resonances, as expected from the crystal structure. On the contrary, ZTA2a-*BDC and ZTA3-*BDC gave more complex signatures due to the presence of several overlapping resonances. The 17O MAS NMR spectra of ZTA3-*BDC could, for example, be fitted considering the presence of four different 17O signals, as expected from the crystal structure and resolved with the MQMAS experiments (see Supporting Information S12). Concerning the ZTA2a-*BDC phase, only two carboxylate oxygen resonances were expected according to the reported crystal structure. However, it was not possible to obtain a satisfactorily simulated spectrum with only two signals. This result supports the observation made previously by 13C NMR for ZTA2a, where an additional 13C carboxylate resonance of weak intensity was detected at 172.5 ppm (see Figure 3b).
To learn more about the minor carboxylate environments in ZTA2a, the 17O MQMAS spectrum of ZTA2a-*BDC was recorded (Figure 6). The two most intense resonances, which are centered at δiso(17O) = 278.3 and 223.0 ppm, belong to the main carboxylate group (i.e., with the 13C signal at 175.7 ppm). Yet, several signals of weaker intensity can also be observed (red-shaded circle), corresponding to multiple and slightly different carboxylate local environments (such regions of signals with lower intensity have been observed on two different batches of ZTA2a-*BDC with two different NMR spectrometers). Three examples of these additional oxygen environments were extracted from this area for further study (Figure 6, right).
Figure 6.
17O MQMAS spectrum of ZTA2a-*BDC recorded at B0 = 18.8 T under νrot = 16 kHz. Extracted slices (in black) are fitted (red dotted lines) with experimental parameters presented in Table 5. The red-shaded circle enlightens some of the additional oxygen sites corresponding to structural defects.
Despite their low signal-to-noise ratio, the signals from the three different slices could be tentatively fitted using quadrupolar line shapes, suggesting that they correspond to carboxylate oxygens in a relatively well-ordered local environment. Using the NMR parameters derived from the different 17O environments resolved by MQMAS, it was then possible to propose a fit of the 17O NMR spectra of ZTA2a-*BDC in the carboxylate region (see Figure 4b and Table 5). However, it should be kept in mind that additional weak-intensity carboxylate resonances may also be overlapping with the two main resonances (which remained unresolved at this stage). Interestingly, the 17O NMR data also show that the main resonances for ZTA2a-*BDC (δiso(17O) = 278.3 and 223.0 ppm) are relatively close in shift to those of ZTA1-*BDC (δiso(17O) = 286.6 and 221.7 ppm), further underscoring the similarities between 17O ssNMR and 13C ssNMR parameters (δiso(17O), δiso(13COO), and 13C CSA) for the two different binding modes, i.e., syn,syn/anti bridging (ZTA2a) and monodentate (ZTA1). Regarding the weaker signals (red circle in Figure 6), they are likely to arise from the presence of the systematic defects inside the crystallographic structure, as already hinted from the 13C NMR data.
Thanks to the selective 17O-labeling of the various Zn-BDC compounds by BM, it was possible to gain novel insights into the local environment of the oxygen-containing ligands, as well as refine one of the crystallographic structures (ZTA3) by DFT calculations. Although no straightforward trend relating 17O NMR parameters to carboxylate binding modes could be derived, most probably because of the influence of several geometrical features (Zn···O and O–C distances, bond angles, positioning of neighboring ligands, etc.), high-resolution 17O ssNMR spectroscopy, in conjunction with DFT calculations and GIPAW calculations of NMR parameters, clearly appears as a crucial asset to help confirm the structural models of these types of MOFs. In particular, they will be useful in the future to help elucidate the nature of the defect sites in ZTA2a in an NMR crystallography type of approach.101
Outlook: Isotope Labeling for Simplified In Situ Observations of Evolutions in the Reaction Media
Because structural rearrangements occur around the water molecules/hydroxyls during the Zn-BDC transformations described above, being able to follow this by Raman spectroscopy appeared to us as an important goal to increase our insight into the transformations taking place. Unfortunately, it was not possible to follow the evolution of the O–H stretching modes upon synthesis by BM with our Raman setup because it only allowed spectra to be recorded from 220 to 3120 cm–1, while ν(OH) bands most often appears at higher wavenumbers.102−104
To circumvent this issue, we thus decided to perform isotope labeling of Zn-BDC phases using deuterium oxide rather than normal water as a starting material in the syntheses. In doing so, the stretching modes arising from −OH groups would correspond to -OD groups, which can be observed between 2400 and 2800 cm–1. Figure 7 presents as an example the result of a synthesis followed by operando Raman spectroscopy and performed using an initial Zn/BDC ratio of 3/1 (i.e., one additional equivalent of ZnO compared to the experiment presented in Figure 1) and D2O instead of H2O. The corresponding temperature measurement through the thermal imaging camera is placed on the side of the Raman spectra (Figure 7a), and the focus of the Raman data is set in the zone between 2600 and 2700 cm–1, where O–D stretching vibrations appear.
Figure 7.
(a) Top: starting compounds used for the operando experiment with their respective equivalents. Bottom: time-resolved operando Raman spectra of two selected areas: 2600–2700 cm–1 range on the left and 1380–1480 cm–1 range on the right, following the Raman spectroscopy convention for displaying the data (high wavenumbers on the right). The graph on the far right corresponds to the temperature measurement recorded using the thermal imaging camera. (b) Experimental ex situ FTIR spectra of ZTA2b (in blue), ZTA2b-OD (in light blue), ZTA3 (in purple), and ZTA3-OD (in light purple), following the FTIR spectroscopy convention for displaying the data (high wavenumbers on the left).
With the starting experimental conditions used here, it is expected to obtain a different final product (i.e., ZTA3) compared to the experiment presented in Figure 1 (final product was ZTA2b). The formation and progressive transformation of the intermediates into ZTA3 are indeed noticeable through the change in temperature, as well as the modifications of the Raman spectra. It should be noted that no signal from ZTA1 was observed in the 2600–2700 cm–1 area (Figure 7) because of the lack of isolated OH environments in this phase, the water molecules being involved in hydrogen bonds. The broad signal arising from these water molecules is indeed too weak to be detected during the operando Raman experiment. Hence, in the 2600–2700 cm–1 range, when working with a 3/1 Zn/BDC ratio in D2O, we observed first the formation of the ZTA2a-OD, through the presence of the ν(OD) band (2660 cm–1). As with the synthesis presented in Figure 1, it then transforms into the ZTA2b-OD structure (2618 cm–1). Then, the excess of zinc oxide allows the formation of the ZTA3-OD compound as the final form (ν(OD) = 2630 and 2671 cm–1). This final phase has a Zn/BDC stoichiometry of 3/1, consistent with the one between the reagents introduced in the jar. The transformation from H2BDC to ZTA3 can also be observed looking at νs(COO–) band between 1380 and 1480 cm–1, the only difference with Figure 1 being the final transformation into the ZTA3 phase (ZTA3-OD here) in the later synthesis. It is worth noting that we were able to observe the two distinct sharp O–D bands in the final compound, corresponding to the expected two isolated hydroxyls (Zn–O(H)–Zn).
The interest of using deuterated precursors for operando Raman mechanochemistry was further demonstrated here, using ring-d4 terephthalic acid as a starting reagent. In doing so, the ν(CH) Raman signals become ν(CD) signals, which, in the case of ZTA1, for example, corresponds to a shift of the stretching vibrations from 3064 and 3105 cm–1 to 2287 and 2321 cm–1 (see Supporting Information Figure S20).105 Therefore, isotope labeling by deuteration can be seen as a way of shifting Raman vibrations in an area of the spectrum where no signal from the jar is expected, allowing a more straightforward observation of the evolving vibration modes as well as an easier postsynthesis data treatment. Both methods, using D2O or d4-BDC, can also be combined to study the change of kinetics for such reactions. Overall, this demonstrates, beyond the study of reaction mechanisms in ball-milling,106 how isotopic labeling by deuteration can be particularly useful also for following the kinetics in mechanochemical reactions by shifting vibration bands into regions where they become detectable (e.g., OD vs OH) and/or to avoid overlaps with other vibration frequencies (including from the jar). Such strategies may also turn out to be particularly valuable to refine kinetic data, which may be ambiguous in other regions (as highlighted at the beginning of this manuscript) and thereby help elucidate complex mechanochemical reactions.
Conclusions
In this contribution, we have performed the first in-depth study of the formation and structure of four coordination polymers involving Zn2+ cations and terephthalate ligands, noted ZTA1, ZTA2a, ZTA2b, and ZTA3. Although each of these phases had been previously reported independently (for instance, as part of investigations aiming at studying the formation of the well-known MOF-5 structure), the details of their structure, and in the case of ZTA2a, the possibility to isolate it as phase-pure, were missing. Here, it is shown for the first time how using mechanochemistry, in conjunction with operando Raman spectroscopy and thermal imaging, the evolution of the reaction medium could be followed. Each of the phases could be isolated as pure, either by adapting the stoichiometry of the Zn and BDC precursors involved in the milling, or by stopping the milling synthesis at a specific time point based on knowledge of the kinetics of the reactions. The study of the formation of ZTA3 by mechanochemistry was shown to proceed through the successive formation of the ZTA1, ZTA2a, and ZTA2b intermediates (see Figure 8).
Figure 8.
Summary of the structural evolution of the Zn-BDC compounds presented in this article upon milling synthesis. The molecular schemes arise from the DFT-optimized structures obtained in this article.
For each of the phases, an NMR crystallography approach was then used, combining high-resolution NMR spectroscopy with ab initio DFT calculations of NMR parameters, to help refine the crystal structures, as much information was lacking in the crystallographic data available to date. Regarding 13C ssNMR, a systematically lower δiso(13COO) was observed for syn,syn bridging ligands as opposed to monodentate and weaker syn,syn/anti bridging modes (∼4 ppm difference in shifts). Moreover, δiso(13COO) and the CSA span (Ω) parameter were found to correlate reasonably well with the longest C–O distance of the carboxylate, making the measurement of these parameters valuable for extracting structural information in future studies of other unknown Zn-terephthalate coordination networks. Furthermore, the 13C ssNMR study allowed an additional carboxylate resonance to be detected for the ZTA2a compound reflecting the likely presence of “defect” in that structure.
Regarding 17O ssNMR, the key point to highlight is that using the protocols developed here by mechanochemistry, it was possible to produce, at low cost, selectively 17O-labeled phases in high yield, with either enriched hydroxyls (Zn–O*(H)–Zn) or enriched carboxylates (CO*O*–Zn) and no isotopic scrambling during the milling (as shown in Figures S14 and S15). Such selective labeling was possible because reactions were performed here by BM under “ambient” temperature and pressure, and in short times (less than 3 h). Moreover, it is important to highlight that the syntheses of these 17O-enriched compounds would have been far too expensive if not for the BM procedures optimized through the use of real-time monitoring Raman spectroscopy. Indeed, the extent of enriched H2O* employed here was very low (for ZTA-*OH, ∼45 μL for ∼300 mg of sample recovered and ∼20 μL for ∼110 mg of sample recovered for ZTA-*BDC) and could not have been matched for syntheses carried out using the hydrothermal or microwave procedures proposed so far to prepare such compounds. Thanks to this isotopic labeling, high-resolution 17O ssNMR spectra could be recorded for each of the phases in just a few hours, allowing different carboxylate oxygen and hydroxyl environments to be resolved. This enabled, for example, a very precise positioning of hydroxyl hydrogens in the case of the ZTA3 phase. Regarding terephthalate ligands, however, no straightforward correlation between 17O NMR parameters and geometric features around the carboxylate oxygens could be derived, showing that the 17O data in these phases are very sensitive to the local environment of oxygen and depend on many different factors (e.g., bond distances, angles). This implies that combined experimental–computational approaches involving various NMR-active nuclei, including 17O, are a very valuable tool for validating structural models (and notably carboxylate binding modes) of these materials.
Among the four phases studied, the ZTA2a compound was proven to present a more complex structure than the one reported, with the probable presence of “defect” sites, as evidenced from both 13C and 17O ssNMR analyses. The information derived from these spectra can serve as the basis for proposing structural models of these defects, which can then be validated using DFT calculations of NMR parameters. More generally speaking, considering the increasing number of studies on MOF structures, which highlight the importance of further refining structural data (beyond what is possible by using powder X-ray diffraction) and to establish the nature and role of defects on reactivity, the possibility of using approaches such as the ones proposed herein, including selective 17O isotopic enrichment, appears very valuable.
Lastly, through the study of the mechanochemical formation of the four Zn-BDC phases using operando Raman spectroscopy, we also demonstrate here for the first time how the use of deuterated precursors can be useful to help follow the course of the reactions by enabling the observation of specific vibration modes (e.g., ν(OD) and/or ν(CD)) and thereby avoiding cutoff of high wavenumbers of the ν(OH) and/or ν(CH) modes due to instrumentation, and/or avoiding their overlap with signals arising from the milling jar in the Raman spectra. We foresee that on a more general perspective, the study of deuterated compounds in operando Raman investigations by BM will be particularly useful to refine the understanding of the reactions occurring in the jars and may enable the more straightforward investigation and optimization of reactions, which not only concern the formation of MOFs but may actually more specifically involve the formation of C–D (instead of C–H) bonds, as it is the case in reduction reactions in organic chemistry.
Acknowledgments
This project has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 772204; 2017 ERC-COG, MISOTOP project). DFT calculations were performed using HPC resources from GENCI-IDRIS (Grant 097535 and 2020-A0090807394). The authors are grateful to Prof. Franziska Emmerling for discussions regarding the Raman operando setup and for providing the Perspex jars used for the real-time monitoring synthesis. Sébastien Mittelette as well as the Mecanox summer school participants are thanked for preparing the 17O-enriched H2BDC. The authors thank Drs. Ieva Goldberga and Philippe Gaveau for assistance in running some of the ssNMR experiments. A portion of this work was performed at the National High Magnetic Field Laboratory, which is supported by the National Science Foundation Cooperative Agreement No. DMR-1644779 and the State of Florida. Financial support from the IR-RMN-THC FR 3050 CNRS for conducting part of the NMR research at the CEMHTI facility in Orléans is gratefully acknowledged.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.chemmater.1c04132.
Experimental details and additional supporting tables and figures: selected ssNMR data (1H, 13C, and 17O NMR spectra, together with their fits); pXRD, FTIR, and Raman data; and graphical plots related to DFT calculations and structural parameters (PDF)
Author Contributions
The project was conducted by C.L. in close interaction with D.L., T.-X.M., and C.G. T.-X.M., C.L., and D.L. developed together the operando Raman spectroscopy setup coupled to mechanochemistry. C.L. performed all BM syntheses, all characterizations by IR, Raman, and pXRD, and the vast majority of ssNMR analyses. I.H. and Z.G. performed part of the high-field 17O NMR experiments reported herein (MagLab). C.G. carried out the computational simulations and GIPAW-DFT calculations. C.L. performed the data analyses and prepared all figures and tables. C.L. and D.L. wrote the initial draft of the manuscript. All authors discussed the results and contributed to the final preparation of the manuscript.
The authors declare no competing financial interest.
Supplementary Material
References
- Friščić T.; Mottillo C.; Titi H. M. Mechanochemistry for Synthesis. Angew. Chem., Int. Ed. 2020, 59, 1018–1029. 10.1002/anie.201906755. [DOI] [PubMed] [Google Scholar]
- Boulineau S.; Courty M.; Tarascon J.-M.; Viallet V. Mechanochemical Synthesis of Li-Argyrodite Li6PS5X (X = Cl, Br, I) as Sulfur-Based Solid Electrolytes for All Solid State Batteries Application. Solid State Ionics 2012, 221, 1–5. 10.1016/j.ssi.2012.06.008. [DOI] [Google Scholar]
- Do J.-L.; Friščić T. Mechanochemistry: A Force of Synthesis. ACS Cent. Sci. 2017, 3, 13–19. 10.1021/acscentsci.6b00277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tan D.; García F. Main Group Mechanochemistry: From Curiosity to Established Protocols. Chem. Soc. Rev. 2019, 48, 2274–2292. 10.1039/C7CS00813A. [DOI] [PubMed] [Google Scholar]
- Gracin D.; Štrukil V.; Friščić T.; Halasz I.; Užarević K. Laboratory Real-Time and In Situ Monitoring of Mechanochemical Milling Reactions by Raman Spectroscopy. Angew. Chem. 2014, 126, 6307–6311. 10.1002/ange.201402334. [DOI] [PubMed] [Google Scholar]
- Užarević K.; Halasz I.; Friščić T. Real-Time and In Situ Monitoring of Mechanochemical Reactions: A New Playground for All Chemists. J. Phys. Chem. Lett. 2015, 6, 4129–4140. 10.1021/acs.jpclett.5b01837. [DOI] [PubMed] [Google Scholar]
- Batzdorf L.; Fischer F.; Wilke M.; Wenzel K.-J.; Emmerling F. Direct In Situ Investigation of Milling Reactions Using Combined X-Ray Diffraction and Raman Spectroscopy. Angew. Chem., Int. Ed. 2015, 54, 1799–1802. 10.1002/anie.201409834. [DOI] [PubMed] [Google Scholar]
- Fischer F.; Wenzel K.-J.; Rademann K.; Emmerling F. Quantitative Determination of Activation Energies in Mechanochemical Reactions. Phys. Chem. Chem. Phys. 2016, 18, 23320–23325. 10.1039/C6CP04280E. [DOI] [PubMed] [Google Scholar]
- Kulla H.; Greiser S.; Benemann S.; Rademann K.; Emmerling F. In Situ Investigation of a Self-Accelerated Cocrystal Formation by Grinding Pyrazinamide with Oxalic Acid. Molecules 2016, 21, 917 10.3390/molecules21070917. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Haferkamp S.; Fischer F.; Kraus W.; Emmerling F. Mechanochemical Knoevenagel Condensation Investigated in Situ. Beilstein J. Org. Chem. 2017, 13, 2010–2014. 10.3762/bjoc.13.197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lukin S.; Stolar T.; Tireli M.; Blanco M. V.; Babić D.; Friščić T.; Užarević K.; Halasz I. Tandem In Situ Monitoring for Quantitative Assessment of Mechanochemical Reactions Involving Structurally Unknown Phases. Chem. - Eur. J. 2017, 23, 13941–13949. 10.1002/chem.201702489. [DOI] [PubMed] [Google Scholar]
- Štrukil V.; Gracin D.; Magdysyuk O. V.; Dinnebier R. E.; Friščić T. Trapping Reactive Intermediates by Mechanochemistry: Elusive Aryl N-Thiocarbamoylbenzotriazoles as Bench-Stable Reagents. Angew. Chem., Int. Ed. 2015, 54, 8440–8443. 10.1002/anie.201502026. [DOI] [PubMed] [Google Scholar]
- Lukin S.; Tireli M.; Lončarić I.; Barišić D.; Šket P.; Vrsaljko D.; di Michiel M.; Plavec J.; Užarević K.; Halasz I. Mechanochemical Carbon–Carbon Bond Formation That Proceeds via a Cocrystal Intermediate. Chem. Commun. 2018, 54, 13216–13219. 10.1039/C8CC07853J. [DOI] [PubMed] [Google Scholar]
- Kulla H.; Wilke M.; Fischer F.; Röllig M.; Maierhofer C.; Emmerling F. Warming up for Mechanosynthesis – Temperature Development in Ball Mills during Synthesis. Chem. Commun. 2017, 53, 1664–1667. 10.1039/C6CC08950J. [DOI] [PubMed] [Google Scholar]
- Kulla H.; Haferkamp S.; Akhmetova I.; Röllig M.; Maierhofer C.; Rademann K.; Emmerling F. In Situ Investigations of Mechanochemical One-Pot Syntheses. Angew. Chem., Int. Ed. 2018, 57, 5930–5933. 10.1002/anie.201800147. [DOI] [PubMed] [Google Scholar]
- Carta M.; Colacino E.; Delogu F.; Porcheddu A. Kinetics of Mechanochemical Transformations. Phys. Chem. Chem. Phys. 2020, 22, 14489–14502. 10.1039/D0CP01658F. [DOI] [PubMed] [Google Scholar]
- Sović I.; Lukin S.; Meštrović E.; Halasz I.; Porcheddu A.; Delogu F.; Ricci P. C.; Caron F.; Perilli T.; Dogan A.; Colacino E. Mechanochemical Preparation of Active Pharmaceutical Ingredients Monitored by In Situ Raman Spectroscopy. ACS Omega 2020, 5, 28663–28672. 10.1021/acsomega.0c03756. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kuppler R. J.; Timmons D. J.; Fang Q.-R.; Li J.-R.; Makal T. A.; Young M. D.; Yuan D.; Zhao D.; Zhuang W.; Zhou H.-C. Potential Applications of Metal-Organic Frameworks. Coord. Chem. Rev. 2009, 253, 3042–3066. 10.1016/j.ccr.2009.05.019. [DOI] [Google Scholar]
- Sakamaki Y.; Ozdemir J.; Heidrick Z.; Watson O.; Shahsavari H. R.; Fereidoonnezhad M.; Khosropour A. R.; Beyzavi M. H. Metal–Organic Frameworks and Covalent Organic Frameworks as Platforms for Photodynamic Therapy. Comments Inorg. Chem. 2018, 38, 238–293. 10.1080/02603594.2018.1542597. [DOI] [Google Scholar]
- Ghanbari T.; Abnisa F.; Wan Daud W. M. A. A Review on Production of Metal Organic Frameworks (MOF) for CO2 Adsorption. Sci. Total Environ. 2020, 707, 135090 10.1016/j.scitotenv.2019.135090. [DOI] [PubMed] [Google Scholar]
- Stock N.; Biswas S. Synthesis of Metal-Organic Frameworks (MOFs): Routes to Various MOF Topologies, Morphologies, and Composites. Chem. Rev. 2012, 112, 933–969. 10.1021/cr200304e. [DOI] [PubMed] [Google Scholar]
- Lee Y.-R.; Kim J.; Ahn W.-S. Synthesis of Metal-Organic Frameworks: A Mini Review. Korean J. Chem. Eng. 2013, 30, 1667–1680. 10.1007/s11814-013-0140-6. [DOI] [Google Scholar]
- Zhao N.; Cai K.; He H. The Synthesis of Metal–Organic Frameworks with Template Strategies. Dalton Trans. 2020, 49, 11467–11479. 10.1039/D0DT01879A. [DOI] [PubMed] [Google Scholar]
- Stolar T.; Užarević K. Mechanochemistry: An Efficient and Versatile Toolbox for Synthesis, Transformation, and Functionalization of Porous Metal–Organic Frameworks. CrystEngComm 2020, 22, 4511–4525. 10.1039/D0CE00091D. [DOI] [Google Scholar]
- Lin C.; Zhou W.; Xiong X.; Xuan W.; Kitson P. J.; Long D.; Chen W.; Song Y.; Cronin L. Digital Control of Multistep Hydrothermal Synthesis by Using 3D Printed Reactionware for the Synthesis of Metal–Organic Frameworks. Angew. Chem. 2018, 130, 16958–16962. 10.1002/ange.201810095. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jones C. L.; Hughes C. E.; Yeung H. H. M.; Paul A.; Harris K. D. M.; Easun T. L. Exploiting in Situ NMR to Monitor the Formation of a Metal–Organic Framework. Chem. Sci. 2021, 12, 1486–1494. 10.1039/D0SC04892E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karadeniz B.; Žilić D.; Huskić I.; Germann L. S.; Fidelli A. M.; Muratović S.; Lončarić I.; Etter M.; Dinnebier R. E.; Barišić D.; Cindro N.; Islamoglu T.; Farha O. K.; Friščić T.; Užarević K. Controlling the Polymorphism and Topology Transformation in Porphyrinic Zirconium Metal–Organic Frameworks via Mechanochemistry. J. Am. Chem. Soc. 2019, 141, 19214–19220. 10.1021/jacs.9b10251. [DOI] [PubMed] [Google Scholar]
- Qin Y.; Wen P.; Voshage M.; Chen Y.; Schückler P. G.; Jauer L.; Xia D.; Guo H.; Zheng Y.; Schleifenbaum J. H. Additive Manufacturing of Biodegradable Zn-XWE43 Porous Scaffolds: Formation Quality, Microstructure and Mechanical Properties. Mater. Des. 2019, 181, 107937 10.1016/j.matdes.2019.107937. [DOI] [Google Scholar]
- Arabbaghi E. K.; Mokhtari J.; Naimi-Jamal M. R.; Khosravi A. Zn-MOF: An Efficient Drug Delivery Platform for the Encapsulation and Releasing of Imatinib Mesylate. J. Porous Mater. 2021, 28, 641–649. 10.1007/s10934-020-01027-3. [DOI] [Google Scholar]
- Karmakar A.; Guedes da Silva M. F. C.; Pombeiro A. J. L. Zinc Metal–Organic Frameworks: Efficient Catalysts for the Diastereoselective Henry Reaction and Transesterification. Dalton Trans. 2014, 43, 7795–7810. 10.1039/C4DT00219A. [DOI] [PubMed] [Google Scholar]
- Wang L.; Zou J.; Chen S.; Yang J.; Qing F.; Gao P.; Li J. Zinc Terephthalates ZnC8H4O4 as Anodes for Lithium Ion Batteries. Electrochim. Acta 2017, 235, 304–310. 10.1016/j.electacta.2017.03.095. [DOI] [Google Scholar]
- Qin X.; Yang W.; Yang Y.; Gu D.; Guo D.; Pan Q. A Zinc Metal–Organic Framework for Concurrent Adsorption and Detection of Uranium. Inorg. Chem. 2020, 59, 9857–9865. 10.1021/acs.inorgchem.0c01072. [DOI] [PubMed] [Google Scholar]
- Friščić T.; Reid D. G.; Halasz I.; Stein R. S.; Dinnebier R. E.; Duer M. J. Ion- and Liquid-Assisted Grinding: Improved Mechanochemical Synthesis of Metal-Organic Frameworks Reveals Salt Inclusion and Anion Templating. Angew. Chem., Int. Ed. 2010, 49, 712–715. 10.1002/anie.200906583. [DOI] [PubMed] [Google Scholar]
- Getachew N.; Chebude Y.; Diaz I.; Sanchez-Sanchez M. Room Temperature Synthesis of Metal Organic Framework MOF-2. J. Porous Mater. 2014, 21, 769–773. 10.1007/s10934-014-9823-6. [DOI] [Google Scholar]
- Xing J.; Schweighauser L.; Okada S.; Harano K.; Nakamura E. Atomistic Structures and Dynamics of Prenucleation Clusters in MOF-2 and MOF-5 Syntheses. Nat. Commun. 2019, 10, 3608 10.1038/s41467-019-11564-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schweighauser L.; Harano K.; Nakamura E. Experimental Study on Interconversion between Cubic MOF-5 and Square MOF-2 Arrays. Inorg. Chem. Commun. 2017, 84, 1–4. 10.1016/j.inoche.2017.07.009. [DOI] [Google Scholar]
- Edgar M.; Mitchell R.; Slawin A. M. Z.; Lightfoot P.; Wright P. A. Solid-State Transformations of Zinc 1,4-Benzenedicarboxylates Mediated by Hydrogen-Bond-Forming Molecules. Chem. - Eur. J. 2001, 7, 5168–5175. . [DOI] [PubMed] [Google Scholar]
- Hirai Y.; Furukawa K.; Sun H.; Matsushima Y.; Shito K.; Masuhara A.; Ono R.; Shimbori Y.; Shiroishi H.; White M. S.; Yoshida T. Microwave-Assisted Hydrothermal Synthesis of ZnO and Zn-Terephthalate Hybrid Nanoparticles Employing Benzene Dicarboxylic Acids. Microsyst. Technol. 2018, 24, 699–708. 10.1007/s00542-017-3392-y. [DOI] [Google Scholar]
- Geranmayeh S.; Abbasi A.; Zarnani A.-H.; Skripkin M. Y. A Novel Trinuclear Zinc Metal–Organic Network: Synthesis, X-Ray Diffraction Structures, Spectroscopic and Biocompatibility Studies. Polyhedron 2013, 61, 6–14. 10.1016/j.poly.2013.05.030. [DOI] [Google Scholar]
- Zhu L.-N.; Zhang L. Z.; Wang W.-Z.; Liao D.-Z.; Cheng P.; Jiang Z.-H.; Yan S.-P. [Zn(BDC)(H2O)2]n: A Novel Blue Luminescent Coordination Polymer Constructed from BDC-Bridged 1-D Chains via Interchain Hydrogen Bonds (BDC = 1,4-Benzenedicarboxylate). Inorg. Chem. Commun. 2002, 5, 1017–1021. 10.1016/S1387-7003(02)00634-2. [DOI] [Google Scholar]
- Guilera G.; Steed J. W. Topological Control in Coordination Polymers by Non-Covalent Forces. Chem. Commun. 1999, 2, 1563–1564. 10.1039/a903163d. [DOI] [Google Scholar]
- Kaduk J. A.CCDC deposition number 1520407; Database identifier: ECATIO, Experimental Crystal Structure Determination, 2016.
- Carton A.; Mesbah A.; Aranda L.; Rabu P.; François M. New Metastable Hybrid Phase, Zn2(OH)2(C8H4O4), Exhibiting Unique Oxo-Penta-Coordinated Zn(II) Atoms. Solid State Sci. 2009, 11, 818–823. 10.1016/j.solidstatesciences.2008.12.011. [DOI] [Google Scholar]
- Carton A.; Abdelouhab S.; Renaudin G.; Rabu P.; François M. Structure of Zinc Hydroxy-Terephthalate: Zn3(OH)4(C8H4O4). Solid State Sci. 2006, 8, 958–963. 10.1016/j.solidstatesciences.2006.02.043. [DOI] [Google Scholar]
- Rodríguez N. A.; Parra R.; Grela M. A. Structural Characterization, Optical Properties and Photocatalytic Activity of MOF-5 and Its Hydrolysis Products: Implications on Their Excitation Mechanism. RSC Adv. 2015, 5, 73112–73118. 10.1039/C5RA11182J. [DOI] [Google Scholar]
- Thirumurugan A.; Rao C. N. R. 1,2-, 1,3- and 1,4-Benzenedicarboxylates of Cd and Zn of Different Dimensionalities: Process of Formation of the Three-Dimensional Structure. J. Mater. Chem. 2005, 15, 3852. 10.1039/b504666a. [DOI] [Google Scholar]
- Zeleňák V.; Vargová Z.; Györyová K. Correlation of Infrared Spectra of Zinc(II) Carboxylates with Their Structures. Spectrochim. Acta, Part A 2007, 66, 262–272. 10.1016/j.saa.2006.02.050. [DOI] [PubMed] [Google Scholar]
- Hadjiivanov K. I.; Panayotov D. A.; Mihaylov M. Y.; Ivanova E. Z.; Chakarova K. K.; Andonova S. M.; Drenchev N. L. Power of Infrared and Raman Spectroscopies to Characterize Metal-Organic Frameworks and Investigate Their Interaction with Guest Molecules. Chem. Rev. 2021, 121, 1286–1424. 10.1021/acs.chemrev.0c00487. [DOI] [PubMed] [Google Scholar]
- Habib H. A.; Hoffmann A.; Höppe H. A.; Janiak C. Crystal Structures and Solid-State CPMAS 13C NMR Correlations in Luminescent Zinc(II) and Cadmium(II) Mixed-Ligand Coordination Polymers Constructed from 1,2-Bis(1,2,4-Triazol-4-Yl)Ethane and Benzenedicarboxylate. Dalton Trans. 2009, 32, 1742. 10.1039/b812670d. [DOI] [PubMed] [Google Scholar]
- Yuan W.; Friščić T.; Apperley D.; James S. L. High Reactivity of Metal-Organic Frameworks under Grinding Conditions: Parallels with Organic Molecular Materials. Angew. Chem., Int. Ed. 2010, 49, 3916–3919. 10.1002/anie.200906965. [DOI] [PubMed] [Google Scholar]
- Julien P. A.; Užarević K.; Katsenis A. D.; Kimber S. A. J.; Wang T.; Farha O. K.; Zhang Y.; Casaban J.; Germann L. S.; Etter M.; Dinnebier R. E.; James S. L.; Halasz I.; Friščić T. In Situ Monitoring and Mechanism of the Mechanochemical Formation of a Microporous MOF-74 Framework. J. Am. Chem. Soc. 2016, 138, 2929–2932. 10.1021/jacs.5b13038. [DOI] [PubMed] [Google Scholar]
- Lucier B. E. G.; Chen S.; Huang Y. Characterization of Metal Organic Frameworks: Unlocking the Potential of Solid-State NMR. Acc. Chem. Res. 2018, 51, 319–330. [DOI] [PubMed] [Google Scholar]
- Al-Terkawi A.-A.; Scholz G.; Prinz C.; Emmerling F.; Kemnitz E. Ca-, Sr-, and Ba-Coordination Polymers Based on Anthranilic Acid via Mechanochemistry. Dalton Trans. 2019, 48, 6513–6521. 10.1039/C9DT00991D. [DOI] [PubMed] [Google Scholar]
- O’Keefe C. A.; Mottillo C.; Vainauskas J.; Fábián L.; Friščić T.; Schurko R. W. NMR-Enhanced Crystallography Aids Open Metal–Organic Framework Discovery Using Solvent-Free Accelerated Aging. Chem. Mater. 2020, 32, 4273–4281. 10.1021/acs.chemmater.0c00894. [DOI] [Google Scholar]
- Hou J.; Ashling C. W.; Collins S. M.; Krajnc A.; Zhou C.; Longley L.; Johnstone D. N.; Chater P. A.; Li S.; Coulet M.-V.; Llewellyn P. L.; Coudert F.-X.; Keen D. A.; Midgley P. A.; Mali G.; Chen V.; Bennett T. D. Metal-Organic Framework Crystal-Glass Composites. Nat. Commun. 2019, 10, 2580 10.1038/s41467-019-10470-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Muratović S.; Karadeniz B.; Stolar T.; Lukin S.; Halasz I.; Herak M.; Mali G.; Krupskaya Y.; Kataev V.; Žilić D.; Užarević K. Impact of Dehydration and Mechanical Amorphization on the Magnetic Properties of Ni(ii)-MOF-74. J. Mater. Chem. C 2020, 8, 7132–7142. 10.1039/D0TC00844C. [DOI] [Google Scholar]
- Stolar T.; Prašnikar A.; Martinez V.; Karadeniz B.; Bjelić A.; Mali G.; Friščić T.; Likozar B.; Užarević K. Scalable Mechanochemical Amorphization of Bimetallic Cu–Zn MOF-74 Catalyst for Selective CO2 Reduction Reaction to Methanol. ACS Appl. Mater. Interfaces 2021, 13, 3070–3077. 10.1021/acsami.0c21265. [DOI] [PubMed] [Google Scholar]
- Kong X.; Terskikh V. V.; Khade R. L.; 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, 4753–4757. 10.1002/anie.201409888. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bignami G. P. M.; Davis Z. H.; Dawson D. M.; Morris S. A.; Russell S. E.; McKay D.; Parke R. E.; Iuga D.; Morris R. E.; Ashbrook S. E. Cost-Effective 17O Enrichment and NMR Spectroscopy of Mixed-Metal Terephthalate Metal–Organic Frameworks. Chem. Sci. 2018, 9, 850–859. 10.1039/C7SC04649A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Špačková J.; Fabra C.; Mittelette S.; Gaillard E.; Chen C.-H.; Cazals G.; Lebrun A.; Sene S.; Berthomieu D.; Chen K.; Gan Z.; Gervais C.; Métro T.-X.; Laurencin D. Unveiling the Structure and Reactivity of Fatty-Acid Based (Nano)Materials Thanks to Efficient and Scalable 17O and 18O-Isotopic Labeling Schemes. J. Am. Chem. Soc. 2020, 142, 21068–21081. 10.1021/jacs.0c09383. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ashbrook S. E.; Davis Z. H.; Morris R. E.; Rice C. M. 17O NMR Spectroscopy of Crystalline Microporous Materials. Chem. Sci. 2021, 12, 5016–5036. 10.1039/D1SC00552A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martins V.; Xu J.; Wang X.; Chen K.; Hung I.; Gan Z.; Gervais C.; Bonhomme C.; Jiang S.; Zheng A.; Lucier B. E. G.; Huang Y. Higher Magnetic Fields, Finer MOF Structural Information: 17O Solid-State NMR at 35.2 T. J. Am. Chem. Soc. 2020, 142, 14877–14889. 10.1021/jacs.0c02810. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Métro T.; Gervais C.; Martinez A.; Bonhomme C.; Laurencin D. Unleashing the Potential of 17O NMR Spectroscopy Using Mechanochemistry. Angew. Chem., Int. Ed. 2017, 56, 6803–6807. 10.1002/anie.201702251. [DOI] [PubMed] [Google Scholar]
- Chen C.-H.; Gaillard E.; Mentink-Vigier F.; Chen K.; Gan Z.; Gaveau P.; Rebière B.; Berthelot R.; Florian P.; Bonhomme C.; Smith M. E.; Métro T.-X.; Alonso B.; Laurencin D. Direct 17O Isotopic Labeling of Oxides Using Mechanochemistry. Inorg. Chem. 2020, 59, 13050–13066. 10.1021/acs.inorgchem.0c00208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen C.-H.; Goldberga I.; Gaveau P.; Mittelette S.; Špačková J.; Mullen C.; Petit I.; Métro T.; Alonso B.; Gervais C.; Laurencin D. Looking into the Dynamics of Molecular Crystals of Ibuprofen and Terephthalic Acid Using 17O and 2H Nuclear Magnetic Resonance Analyses. Magn. Reson. Chem. 2021, 59, 975–990. 10.1002/mrc.5141. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schiffmann J. G.; Emmerling F.; Martins I. C. B.; Van Wüllen L. In-Situ Reaction Monitoring of a Mechanochemical Ball Mill Reaction with Solid State NMR. Solid State Nucl. Magn. Reson. 2020, 109, 101687 10.1016/j.ssnmr.2020.101687. [DOI] [PubMed] [Google Scholar]
- Kentgens A. P. M.; Verhagen R. Advantages of Double Frequency Sweeps in Static, MAS and MQMAS NMR of Spin I = 3/2 Nuclei. Chem. Phys. Lett. 1999, 300, 435–443. 10.1016/S0009-2614(98)01402-X. [DOI] [Google Scholar]
- Iuga D.; Schäfer H.; Verhagen R.; Kentgens A. P. M. Population and Coherence Transfer Induced by Double Frequency Sweeps in Half-Integer Quadrupolar Spin Systems. J. Magn. Reson. 2000, 147, 192–209. 10.1006/jmre.2000.2192. [DOI] [PubMed] [Google Scholar]
- Siegel R.; Nakashima T. T.; Wasylishen R. E. Signal-to-Noise Enhancement of NMR Spectra of Solids Using Multiple-Pulse Spin-Echo Experiments. Concepts Magn. Reson., Part A 2005, 26A, 62–77. 10.1002/cmr.a.20038. [DOI] [Google Scholar]
- Brinkmann A.; Kentgens A. P. M. Proton-Selective 17O–1H Distance Measurements in Fast Magic-Angle-Spinning Solid-State NMR Spectroscopy for the Determination of Hydrogen Bond Lengths. J. Am. Chem. Soc. 2006, 128, 14758–14759. 10.1021/ja065415k. [DOI] [PubMed] [Google Scholar]
- Massiot D.; Fayon F.; Capron M.; King I.; Le Calvé S.; Alonso B.; Durand J.-O.; Bujoli B.; Gan Z.; Hoatson G. Modelling One- and Two-Dimensional Solid-State NMR Spectra. Magn. Reson. Chem. 2002, 40, 70–76. 10.1002/mrc.984. [DOI] [Google Scholar]
- Jameson C. J. Reply to conventions for Tensor Quantities Used in Nuclear Magnetic Resonance, Nuclear Quadrupole Resonance and Electron Spin Resonance Spectroscopy. Solid State Nucl. Magn. Reson. 1998, 11, 265–268. 10.1016/S0926-2040(98)00029-0. [DOI] [PubMed] [Google Scholar]
- Kresse G.; Hafner J. Ab Initio Molecular-Dynamics Simulation of the Liquid-Metal–Amorphous-Semiconductor Transition in Germanium. Phys. Rev. B 1994, 49, 14251–14269. 10.1103/PhysRevB.49.14251. [DOI] [PubMed] [Google Scholar]
- Giannozzi P.; Baroni S.; Bonini N.; Calandra M.; Car R.; Cavazzoni C.; Ceresoli D.; Chiarotti G. L.; 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 A. P.; Smogunov A.; Umari P.; Wentzcovitch R. M. QUANTUM ESPRESSO: A Modular and Open-Source Software Project for Quantum Simulations of Materials. J. Phys.: Condens. Matter 2009, 21, 395502 10.1088/0953-8984/21/39/395502. [DOI] [PubMed] [Google Scholar]
- Baroni S.; de Gironcoli S.; Dal Corso A.; Giannozzi P. Phonons and Related Crystal Properties from Density-Functional Perturbation Theory. Rev. Mod. Phys. 2001, 73, 515–562. 10.1103/RevModPhys.73.515. [DOI] [Google Scholar]
- Perdew J. P.; Burke K.; Ernzerhof M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
- Troullier N.; Martins J. L. Efficient Pseudopotentials for Plane-Wave Calculations. Phys. Rev. B 1991, 43, 1993–2006. 10.1103/PhysRevB.43.1993. [DOI] [PubMed] [Google Scholar]
- Kleinman L.; Bylander D. M. Efficacious Form for Model Pseudopotentials. Phys. Rev. Lett. 1982, 48, 1425–1428. 10.1103/PhysRevLett.48.1425. [DOI] [Google Scholar]
- Pickard C. J.; Mauri F. All-Electron Magnetic Response with Pseudopotentials: NMR Chemical Shifts. Phys. Rev. B 2001, 63, 245101 10.1103/PhysRevB.63.245101. [DOI] [Google Scholar]
- Lejaeghere K.; Bihlmayer G.; Björkman T.; Blaha P.; Blügel S.; Blum V.; Caliste D.; Castelli I. E.; Clark S. J.; Dal Corso A.; de Gironcoli S.; Deutsch T.; Dewhurst J. K.; Di Marco I.; Draxl C.; Dułak M.; Eriksson O.; Flores-Livas J. A.; Garrity K. F.; Genovese L.; Giannozzi P.; Giantomassi M.; Goedecker S.; Gonze X.; Grånäs O.; Gross E. K. U.; Gulans A.; Gygi F.; Hamann D. R.; Hasnip P. J.; Holzwarth N. A. W.; Iuşan D.; Jochym D. B.; Jollet F.; Jones D.; Kresse G.; Koepernik K.; Küçükbenli E.; Kvashnin Y. O.; Locht I. L. M.; Lubeck S.; Marsman M.; Marzari N.; Nitzsche U.; Nordström L.; Ozaki T.; Paulatto L.; Pickard C. J.; Poelmans W.; Probert M. I. J.; Refson K.; Richter M.; Rignanese G.-M.; Saha S.; Scheffler M.; Schlipf M.; Schwarz K.; Sharma S.; Tavazza F.; Thunström P.; Tkatchenko A.; Torrent M.; Vanderbilt D.; van Setten M. J.; Van Speybroeck V.; Wills J. M.; Yates J. R.; Zhang G.-X.; Cottenier S. Reproducibility in Density Functional Theory Calculations of Solids. Science 2016, 351, 1415. 10.1126/science.aad3000. [DOI] [PubMed] [Google Scholar]
- Stievano L.; Tielens F.; Lopes I.; Folliet N.; Gervais C.; Costa D.; Lambert J.-F. Density Functional Theory Modeling and Calculation of NMR Parameters: An Ab Initio Study of the Polymorphs of Bulk Glycine. Cryst. Growth Des. 2010, 10, 3657–3667. 10.1021/cg100525h. [DOI] [Google Scholar]
- Colas H.; Bonhomme-Coury L.; Diogo C. C.; Tielens F.; Babonneau F.; Gervais C.; Bazin D.; Laurencin D.; Smith M. E.; Hanna J. V.; Daudon M.; Bonhomme C. Whewellite, CaC2O4·H2O: Structural Study by a Combined NMR, Crystallography and Modelling Approach. CrystEngComm 2013, 15, 8840. 10.1039/c3ce41201f. [DOI] [Google Scholar]
- Pyykkö P. Year-2017 Nuclear Quadrupole Moments. Mol. Phys. 2018, 116, 1328–1338. 10.1080/00268976.2018.1426131. [DOI] [Google Scholar]
- Friščić T.; Childs S. L.; Rizvi S. A. A.; Jones W. The Role of Solvent in Mechanochemical and Sonochemical Cocrystal Formation: A Solubility-Based Approach for Predicting Cocrystallisation Outcome. CrystEngComm 2009, 11, 418–426. 10.1039/B815174A. [DOI] [Google Scholar]
- Bowmaker G. A. Solvent-Assisted Mechanochemistry. Chem. Commun. 2013, 49, 334–348. 10.1039/C2CC35694E. [DOI] [PubMed] [Google Scholar]
- Téllez S C. A.; Hollauer E.; Mondragon M. A.; Castaño V. M. Fourier Transform Infrared and Raman Spectra, Vibrational Assignment and Ab Initio Calculations of Terephthalic Acid and Related Compounds. Spectrochim. Acta, Part A 2001, 57, 993–1007. 10.1016/S1386-1425(00)00428-5. [DOI] [PubMed] [Google Scholar]
- Varghese H. T.; Panicker C. Y.; Philip D.; Sreevalsan K.; Anithakumary V. IR, Raman and SERS Spectra of Disodium Terephthalate. Spectrochim. Acta, Part A 2007, 68, 817–822. 10.1016/j.saa.2006.12.065. [DOI] [PubMed] [Google Scholar]
- Embrechts H.; Kriesten M.; Ermer M.; Peukert W.; Hartmann M.; Distaso M. In Situ Raman and FTIR Spectroscopic Study on the Formation of the Isomers MIL-68(Al) and MIL-53(Al). RSC Adv. 2020, 10, 7336–7348. 10.1039/C9RA09968A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stroh J.; Ali N. Z.; Maierhofer C.; Emmerling F. Ettringite via Mechanochemistry: A Green and Rapid Approach for Industrial Application. ACS Omega 2019, 4, 7734–7737. 10.1021/acsomega.9b00560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Harris R. K.; Wasylishen R. E.; Duer M. J.. NMR Crystallography; John Wiley & Sons, Ltd.: Chichester, UK, 2009. [Google Scholar]
- Bryce D. L. NMR Crystallography: Structure and Properties of Materials from Solid-State Nuclear Magnetic Resonance Observables. IUCrJ 2017, 4, 350–359. 10.1107/S2052252517006042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smith M. E. Recent Progress in Solid-state Nuclear Magnetic Resonance of Half-integer Spin Low-γ Quadrupolar Nuclei Applied to Inorganic Materials. Magn. Reson. Chem. 2021, 59, 864–907. 10.1002/mrc.5116. [DOI] [PubMed] [Google Scholar]
- Leroy C.; Szell P. M. J.; Bryce D. L. On the Importance of Accurate Nuclear Quadrupole Moments in NMR Crystallography. Magn. Reson. Chem. 2019, 57, 265–267. 10.1002/mrc.4787. [DOI] [PubMed] [Google Scholar]
- Leroy C.; Bryce D. L. Recent Advances in Solid-State Nuclear Magnetic Resonance Spectroscopy of Exotic Nuclei. Prog. Nucl. Magn. Reson. Spectrosc. 2018, 109, 160–199. 10.1016/j.pnmrs.2018.08.002. [DOI] [PubMed] [Google Scholar]
- Johnston J. C.; Iuliucci R. J.; Facelli J. C.; Fitzgerald G.; Mueller K. T. Intermolecular Shielding Contributions Studied by Modeling the 13C Chemical-Shift Tensors of Organic Single Crystals with Plane Waves. J. Chem. Phys. 2009, 131, 144503 10.1063/1.3225270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- No information regarding ″defect″ sites could be found in the X-ray diffraction data for the CCDC structure named ″ECATIO″.
- Pourpoint F.; Gervais C.; Bonhomme-Coury L.; Azaïs T.; Coelho C.; Mauri F.; Alonso B.; Babonneau F.; Bonhomme C. Calcium Phosphates and Hydroxyapatite: Solid-State NMR Experiments and First-Principles Calculations. Appl. Magn. Reson. 2007, 32, 435–457. 10.1007/s00723-007-0040-1. [DOI] [Google Scholar]
- Reinholdt M.; Croissant J.; Di Carlo L.; Granier D.; Gaveau P.; Bégu S.; Devoisselle J.-M.; Mutin P. H.; Smith M. E.; Bonhomme C.; Gervais C.; van der Lee A.; Laurencin D. Synthesis and Characterization of Crystalline Structures Based on Phenylboronate Ligands Bound to Alkaline Earth Cations. Inorg. Chem. 2011, 50, 7802–7810. 10.1021/ic200961a. [DOI] [PubMed] [Google Scholar]
- Sene S.; Reinholdt M.; Renaudin G.; Berthomieu D.; Zicovich-Wilson C. M.; Gervais C.; Gaveau P.; Bonhomme C.; Filinchuk Y.; Smith M. E.; Nedelec J.-M.; Bégu S.; Mutin P. H.; Laurencin D. Boronate Ligands in Materials: Determining Their Local Environment by Using a Combination of IR/Solid-State NMR Spectroscopies and DFT Calculations. Chem. - Eur. J. 2013, 19, 880–891. 10.1002/chem.201203560. [DOI] [PubMed] [Google Scholar]
- Martins V.; Xu J.; Hung I.; Gan Z.; Gervais C.; Bonhomme C.; Huang Y. 17O Solid-state NMR at Ultrahigh Magnetic Field of 35.2 T: Resolution of Inequivalent Oxygen Sites in Different Phases of MOF MIL-53(Al). Magn. Reson. Chem. 2021, 59, 940–950. 10.1002/mrc.5122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bennett T. D.; Todorova T. K.; Baxter E. F.; Reid D. G.; Gervais C.; Bueken B.; Van de Voorde B.; De Vos D.; Keen D. A.; Mellot-Draznieks C. Connecting Defects and Amorphization in UiO-66 and MIL-140 Metal–Organic Frameworks: A Combined Experimental and Computational Study. Phys. Chem. Chem. Phys. 2016, 18, 2192–2201. 10.1039/C5CP06798G. [DOI] [PubMed] [Google Scholar]
- Johansson U.; Frost R. L.; Forsling W.; Kloprogge J. T. Raman Spectroscopy of the Kaolinite Hydroxyls at 77 K. Appl. Spectrosc. 1998, 52, 1277–1282. 10.1366/0003702981942780. [DOI] [Google Scholar]
- Kolesov B. Raman Investigation of H2O Molecule and Hydroxyl Groups in the Channels of Hemimorphite. Am. Mineral. 2006, 91, 1355–1362. 10.2138/am.2006.2179. [DOI] [Google Scholar]
- Yang M.; Han L.; Xu Y.; Ke H.; Zhou N.; Dong H.; Liu S.; Qiao G. Near Infrared Spectroscopic Study of Trioctahedral Chlorites and Its Remote Sensing Application. Open Geosci. 2019, 11, 815–828. 10.1515/geo-2019-0063. [DOI] [Google Scholar]
- Flakus H. T.; Hachuła B.; Hołaj-Krzak J. T.; Al-Agel F. A.; Rekik N. “Long-Distance” H/D Isotopic Self-Organization Phenomena in Scope of the Infrared Spectra of Hydrogen-Bonded Terephthalic and Phthalic Acid Crystals. Spectrochim. Acta, Part A 2017, 173, 65–74. 10.1016/j.saa.2016.08.051. [DOI] [PubMed] [Google Scholar]
- Lukin S.; Tireli M.; Stolar T.; Barišić D.; Blanco M. V.; di Michiel M.; Užarević K.; Halasz I. Isotope Labeling Reveals Fast Atomic and Molecular Exchange in Mechanochemical Milling Reactions. J. Am. Chem. Soc. 2019, 141, 1212–1216. 10.1021/jacs.8b12149. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.








