Next Article in Journal
Highly Efficient Cobalt Sulfide Heterostructures Fabricated on Nickel Foam Electrodes for Oxygen Evolution Reaction in Alkaline Water Electrolysis Cells
Previous Article in Journal
Understanding the Effects of Post-Deposition Sequential Annealing on the Physical and Chemical Properties of Cu2ZnSnSe4 Thin Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Selective Chemical Filters for VOF3: Tailoring MgF2 Filter Selectivity through Surface Chemistry

1
Institut de Chimie de Clermont-Ferrand, CNRS, Université Clermont Auvergne, F-63000 Clermont-Ferrand, France
2
Clermont Auvergne INP, Université Clermont Auvergne, CNRS UMR6296, 24 Av. Blaise Pascal, F-63178 Aubière, France
3
Orano Cycle Tricastin, Hall de Recherche de Pierrelatte, F-26701 Pierrelatte, France
*
Author to whom correspondence should be addressed.
Surfaces 2023, 6(4), 480-492; https://doi.org/10.3390/surfaces6040032
Submission received: 22 September 2023 / Revised: 30 October 2023 / Accepted: 8 November 2023 / Published: 19 November 2023

Abstract

:
In order to synthesize chemical filters for the selective removal of volatile fluorides, commercial magnesium fluoride MgF2 with high specific surface area (HSA) was investigated. The amount of -OH groups substituting fluorine is not negligible, partly due to the high surface area, but also due to the synthesis route. These hydroxyl groups induce a Lewis basicity on the surface of metal fluorides. The amount of these Lewis basic sites has been tailored using fluorination with F2 gas. The sorption of VOF3, used as model gas, onto these fluorides was investigated. The versatility of surface chemistry as a function of a number of Lewis basic sites opens the way to filter selectivity mixture of volatile fluorides depending on their Lewis acidity. HSA MgF2 acts as a stable matrix towards the gas to be purified, and the selectivity may be achieved by a higher Lewis acidity of the gaseous impurity.

Graphical Abstract

1. Introduction

Volatile fluorides are involved in numerous industrial applications, either as reagents or pollutants. Most of them are Lewis acids and react with a wide variety of compounds; for example, in organic synthesis, gaseous VOF3 is often used for oxidative coupling of phenolic rings [1]. Due to their ability to release atomic fluorine, volatile fluorides can also be used as fluorinating agents. Nanoparticles, thin films or atomic layers of tungsten and molybdenum metal are synthesized using WF6 or MoF6 as precursors [2,3]; metal deposition occurs through the reaction of MF6 into M plus 3 F2. A reactant such as Si2H6 is used together with MoF6 to form Mo layers. Dense films of molybdenum oxide may be deposited by plasma-enhanced chemical vapor deposition using mixtures of MoF6, H2, and O2 [4].
When the volatile fluorides are pollutants, however, chemical filters are needed to remove them. The most important application concerns UF6, which plays a key role in the nuclear industry. UF6, as a volatile uranium compound, allows uranium enrichment to 235U whatever the process, i.e., gaseous diffusion, centrifugation, or laser excitation. This particular case highlights the necessity for selectivity of the chemical filter. To provide high-purity nuclear fuel, international standards have been established and the quantity of pollutants allowed in nuclear pellets is restricted. Pollutants originate from both uranium ore, the chemical agents used in the conversion process, and/or the fission products of spent uranium. All these elements are fluorinated during the synthesis of UF6 and their volatility and solubility in UF6 vary depending on the element [5]. When the volatility of the impurity is close to that of UF6, its removal is complicated. Selective chemical filters that do not react with UF6 are thus highly needed. Ca(OH)2, Mg(OH)2, KOH, either separately or in combination, cannot be used because they react easily with UF6 [6].
Our strategy involves metal fluorides which contain just a small part of basic -OH groups. Considering the reactivity of the volatile fluorides which must be trapped, the criteria for an effective chemical filter are a high surface area, high porosity to increase the interface with the target gas, chemical stability in order to avoid their decomposition, but also enough Lewis basic sites on the surface to react with volatile fluorides as Lewis acids. To reach sufficient selectivity, the surface chemistry or porosity of the filter must be able to match the chemisorption or physisorption of the gas onto the filter surface.
To fulfil these criteria, our choice was to go towards fluorides with high specific surface area (HSA). Usually, such fluorides are obtained via the sol–gel route [7,8,9,10,11] or by microwave-assisted solvothermal synthesis [12,13,14,15,16,17]. HSA metal fluorides are prepared using metal alkoxides in an organic medium or various solvents and metal precursors in an aqueous-HF medium. The synthesis conditions using microwave-assisted solvothermal routes, i.e., the choice of precursors, solvent, HF concentration and reaction temperature, strongly influence the formation of various networks with different chemical compositions. Magnesium difluoride MgF2, better known for its catalytic [18,19] and optical properties [20,21,22], has been selected as a chemical filter based on previous data on the removal of impurities such as fluorides of technetium [23], molybdenum [24], ruthenium, neptunium or plutonium [25].
The presence of weak basic sites was evidenced in MgF2 prepared by sol–gel [26] and weak basic sites may prevent reaction with UF6. Basic sites coexist with a large amount of Lewis acid sites, which explains the unique catalytic properties of MgF2 prepared by sol–gel [8]. F and O2− are both intrinsically Lewis bases; however, fluorine atoms reduce the basicity of oxygen atoms. Moreover, the basicity of MgF2 is much lower than that of MgO. We propose to adapt the OH content in order to reach the selectivity, VOF3 versus UF6 for instance, but in a general way by using the Lewis acidity differences of the gases in the mixture. Moreover, the aim is to design a chemical filter that may be regenerated; metal fluoride acts as the support and the OH content changes during reaction with the impurity to be removed.

2. Materials and Methods

2.1. Materials

In order to test the tuning of the OH content, three MgF2−x(OH)x were used: a commercial MgF2 received in pellet form (Nippon Puretec, Nagoya, Japan), which is used as either as received or after fluorination treatment, and a locally synthesized oxygen-free MgF2. The commercial sample was chosen for its ease of use, in pellet rather than powder form, for all the filtering and regeneration operations as well as for the quantities available for future industrial uses. VOF3 was synthesized locally and used as is.

2.2. Filtering Capacities

All experiments were performed in polytetrafluoroethylene (PTFE) bottles sealed in a nitrogen dry box due to the hygroscopic nature of fluorides. MgF2−x(OH)x pellets and VOF3 powders were put in separate nickel baskets. The volatility of VOF3 at 80 °C allowed the exposure of its gas on MgF2−x(OH)x samples. After exposure to VOF3 for 24 h, the MgF2−x(OH)x pellets were crushed for characterization. The vanadium content on the filter surface was estimated by weight uptake and ICP analysis, both of which gave consistent data.

2.3. Fluorination

Pure molecular fluorine (Solvay, 99%+) was used. A chemical trap filled with soda lime scrubbed F2 molecules from the exhaust in order to avoid their release into the atmosphere. The gas flow was set at 20 mL·min−1. The treatment was performed at a fluorination temperature TF = 300 °C. The temperature profile of the treatment is a heating ramp of 5 °C · min−1, with a stabilization of temperature for 4 h. After this, the reactor was flushed with nitrogen in order to remove all reactive gases.

2.4. Characterization

2.4.1. X-ray Powder Diffraction

X-ray powder diffraction patterns were recorded with a Panalytical X’Pert powder diffractometer in Θ- Θ Bragg Bentano geometry. The samples were transferred to a sealed cell to avoid exposure to moisture, with an aluminum sample holder for some. All patterns were recorded between 5° and 70° in 2Θ with a step of 0.015° and a counting time of 60 min using a back graphite monochromated CuKα radiation (Kα1 = 1.54056 Å and Kα2 = 1.54439 Å). Profile matching refinements were performed using the FULLPROF software [27].

2.4.2. Adsorption/Desorption Isotherms

The nitrogen adsorption/desorption isotherms were measured using the Micromeritics ASAP 2020 instrument. Prior to each adsorption experiment, the samples were degassed at 473 K under primary vacuum and then under secondary vacuum. Pore volume, specific surface area, and pore size distribution were extracted from the N2 adsorption/desorption isotherms at 77 K using the BET (Brunauer, Emmett and Teller) and BJH (Barrett, Joyner and Halenda) models for specific surface area (SSA) and pore size distribution for mesoporous materials, respectively.

2.4.3. Raman Spectroscopy

Raman spectra were collected at room temperature using a Bruker RFS 100/S apparatus with a Nd-YAG (aluminum-doped yttrium garnet) laser source at 1064 nm. A total of 500 scans were recorded between 4000 and 25 cm−1 Raman shift. Samples were prepared in a sealed fluorinated ethylene-propylene tube (FEP, La Mothe-aux-Aulnais, Saint Gobain) that resulted in the presence of additional Raman bands (marked on the spectra). For the Raman analyses, the MgF2 single-crystal (Sigma Aldrich, 99%) was used as the reference for oxygen-free magnesium difluoride.

2.4.4. NMR Spectroscopy

Multinuclear 19F, 1H and 51V NMR measurements were carried out with a Bruker Advance Spectrometer with working frequency of 282.2, 300.0 and 78.8 MHz, respectively. A Magic Angle Spinning (MAS) probe operating with 2.5 mm rotors was used allowing a 30 kHz spinning rate. A sequence with a single π/2 pulse duration of 4.0 μs was used. The 19F, 1H and 51V NMR chemical shifts were externally referenced to CFCl3, tetramethylsilane (TMS) and solution of vanadium phosphate (1 M), respectively.

3. Results and Discussion

3.1. Fluorination to Tailor OH/F Ratio in MgF2−x(OH)x

In order to test the tuning of the OH content, three MgF2−x(OH)x were used. All the characterizations highlight the presence of -OH groups in the commercial sample, and synthesis via the sol–gel route is strongly suspected. The chemical composition may be written as MgF2−x(OH)x. The aim of post-fluorination of the commercial source is to tailor its OH/F ratio and surface chemistry, but to obtain pure oxygen-free MgF2, the choice was made to fluorinate an oxygen-free source.

3.1.1. Conversion of MgB2 into Oxygen-Free MgF2

In order to select the precursor for the synthesis of oxygen-free MgF2 via an etching during the fluorination of the element other than Mg, i.e., B, N; Si or P in MgB2 (BF3 evolution), Mg3N2 (NF3), Mg2Si (SiF4) and Mg3P2, the following criteria were used:
  • The expected pore size, in accordance with the size of the released molecules (0.243, 0.320, 0.377 and 0.377 nm) for MgB2 (BF3), Mg3N2 (NF3), Mg2Si (SiF4) and Mg3P2 (PF5), respectively.
  • The toxicity of the gases released at the completion of the first reaction item.
  • The presence of solid products other than MgF2.
According to (3) MgC2 cannot be retained because fluorocarbons may be formed. MgH2 results after fluorination in the narrowest pores and has not been selected because HF is also undesirable according to (2). A toxic gaseous mixture S2F2/SF6/F2 is also formed from MgS, excluding this precursor according to (2). Mg3N2 is not considered because a thermal post-treatment is necessary to remove NH4F from as-prepared MgF2, that should lead to decrease the surface area. A narrow pore size being preferred, MgB2 is selected rather than Mg3P2. The fluorination of this MgB2 precursor at 300 K and 1 atm occurred in two steps: for the addition of F2 in between 0 and 4 moles, the initial precursor is totally consumed to form solid boron and MgF2; when F2 is further added, the boron is then fluorinated as gaseous BF3. From 4 moles of fluorine gas, the only solid product is MgF2 and the gaseous mixture consists of BF3 and F2. A solid reaction yield slightly higher than 100% is explained by the presence of the intermediate product Mg(BF4)2 found at the completion of the reaction.
While the XRD pattern of the final product (Figure 1a) reveals only magnesium difluoride MgF2 [28,29] with traces of MgB2, the presence of the intermediate product Mg(BF4)2 is unambiguously revealed by FTIR spectroscopy (Figure 1b) considering the B-F bond vibration bands of at 1111, 1080, 461 cm−1 [30].
The bands related to the hydroxyl groups (3430 and 1647 cm−1) are absent contrary to the other MgF2 samples (main peak at 435 cm−1 associated with Mg-F vibration mode) [31]. Two bands are identified in the 19F NMR spectrum (Figure 1c). The main band at −196 ppm is assigned to Mg-F bonds. A small shoulder centered at −145 ppm is due to BF4 present in very few amounts [32]. BF4 reveals the presence of an intermediate compound between MgB2 and BF3 following the mechanism:
M g B 2 ( s , black ) Δ , + 4 F 2 M g B F 4 2 ( s ) Δ , 2 B F 3 M g F 2 ( s , white )
It is worthwhile to note that no shoulders indicating the presence of hydroxyl groups are observed in the final product. As a matter of fact, OH groups result in a change in electronic density around the 19F nuclei in their neighboring and consequently a small band should appear at higher chemical shift due to the decrease in the Mg-F bond ionicity (increase in the covalence of the Mg-F bond). A band or shoulder is then observed in the −160/−180 ppm range for other MgF2 samples; its intensity is related to the synthesis and post-treatment but mainly to the nature and concentration of the hydroxyl groups substituting fluorine. The higher the intensity of this band, the higher the content of hydroxyl groups, (Figure 1c). The 1H NMR spectrum of MgF2 without oxygen (Figure 1d) confirms the absence of hydroxyl groups. The objective of preparing oxygen-free MgF2 is thus reached. The sample obtained by fluorination of the magnesium diboride precursor exhibits a type II profile for the N2 isotherm (Figure 1e), typical of non-porous or macroporous materials.
The absence of micropores indicates that the lattice is totally rebuilt during the chemical etching and BF3 evolution. One should note the SSA (specific surface area) of 35 m2 · g−1 for the oxygen-free MgF2. BF3 gaseous molecules which are produced during the fluorination of MgB2 allow a relatively high specific surface area to be maintained, that is unusual with gas/solid fluorination synthesis of fluorides.

3.1.2. Tailoring of OH/F Ratio in Conventional MgF2

Considering the 19F NMR spectra (Figure 1c), the main band at −196 ppm is assigned to 19F nuclei in the F-Mg-F groups. This chemical shift is in accordance with the literature data [32]. A shoulder also appears for the raw compound and its intensity decreases after post-fluorination treatment. This shoulder is relative to the presence of hydroxyl groups (OH-Mg-F) in MgF2. As mentioned previously, the position of the band gives information on the fluorine–oxygen environment of magnesium in MgF2. By fitting the spectra using two Lorentzian lines, the amount of hydroxyl groups is obtained for each compound. The OH/F ratio is 0.14 for raw MgF2 and 0.04 for the MgF2 post-fluorinated at 240 °C. The composition of the two samples can be written MgF1.75(OH)0.25 and MgF1.925(OH)0.075. Regarding 1H NMR spectra (Figure 1d), only one band is observed (the others being related to the rotor cap) corresponding to the hydroxyl groups in MgF2. After fluorination, its intensity, i.e., the amount of OH groups, decreases, confirming the efficiency of the treatment in removing OH.
Conversely, oxygen-free MgF2 exhibits the characteristics of a non-porous or macroporous compound, raw and post-fluorinated MgF2 present a type IV hysteresis which is typical of a mesoporous material (Figure 1e). The BJH method indicates an average pore size of 8.3 and 20.3 nm for raw and post-fluorinated pellets of commercial MgF2, respectively. The increase in pore diameter is due to the coalescence phenomenon induced both by the fluorination temperature and by the release of OH groups. It is not possible to extract an average pore diameter using the BJH method for the sample obtained from the boride precursor because the BJH method is only suitable for mesoporous materials. The BET surfaces are 100 and 72 m2 · g−1 for raw and post-fluorinated MgF2, respectively. The post-fluorination treatment decreases the surface area by substituting fluorine atoms for hydroxyl groups.

3.2. Sorption of VOF3 in MgF2−x(OH)x

At this step, three different kinds of magnesium difluoride with various contents of OH groups were available for VOF3 sorption tests: commercial (MgF1.75(OH)0.25), post-fluorinated (MgF1.925(OH)0.075) and oxygen-free MgF2 (MgF2). After exposure to VOF3, the relative quantity of vanadium trapped was 3.6 w.% for the raw MgF2, 2.1 w.% for the post-fluorinated sample and 1.3 w.% for the oxygen-free synthesized product. All weight uptakes were confirmed by ICP analysis (Table 1). The higher the OH/F ratio, the higher the sorption rate. The similitudes between the XRD patterns before and after exposure to VOF3 indicate no new crystalline phase formed nor any change in crystallinity, the width of diffraction peak being the same as that of MgF2 before the sorption of VOF3.
The FTIR spectra of the samples (Figure 2a) reacting with vanadium oxyfluoride reveal the occurrence of magnesium—fluorine and vanadium—oxygen bonds. The V=O and V-F stretching bands of vanadium oxyfluoride are identified as raw and treated commercial MgF2 after sorption at 1000 cm−1 (V=O) and 820 cm−1 (V-O) on the IR spectra (Figure 2a) [33,34]. The intensities of the vanadium–oxygen vibration bands are higher for raw MgF2 than for treated MgF2 in accordance with the amount of vanadium trapped (3.6 and 2.1 w.%, respectively); see Table 1. Only low bands are detected at 1000 cm−1 (V=O) and 720 cm−1 (V-F) in the case of oxygen-free samples.
Gas phase IR and pressure measurements were also performed to follow the nature of the gas generated or consumed during the sorption of VOF3 on MgF2. The pressure drops in a first sorption step as expected due to the trapping of VOF3 molecules onto the surface of the filter, but increases rapidly after, indicating the release of other gases into the IR chamber. The gas-phase FTIR spectra pointed out a massive group of bands between 4500 and 3500 cm−1 (Figure 2c). These bands are characteristic of gaseous HF. Their intensity increased during the sorption (Figure 2b).
HF is unambiguously the product of a chemical reaction occurring between MgF2−x(OH)x and VOF3. This constitutes another proof of a chemisorption process. V=O and V-F vibration bands characteristic of vanadium oxyfluoride were also detected with Raman spectroscopy (Figure 3).
Data from the literature indicate that the first band at 1018 cm−1 is assigned to the vibration of the V=O bond in VOF3 whereas the second at 798 cm−1 is related to the V-F bond [35]. As in the FTIR data of MgF2 exposed to VOF3, the peak intensity was higher for the raw MgF2 than for the post-fluorinated sample, once again in accordance with the sorption rates. No bands were observed with the oxygen-free crystalline MgF2 sample as the quantity of vanadium trapped (close to 1 w.%) was not sufficient to detect any sorption product.
After sorption of VOF3, the 19F NMR spectra show a single line due to Mg-F bonds (Figure 4a). The shoulder indicating the presence of hydroxyl groups disappears after the sorption. A chemical reaction occurs that removes OH groups from the surface of MgF2 and involves the volatile compound. No further lines due to fluorinated vanadium compounds are observed in the final product. For the 1H NMR spectra (Figure 4b), the peak intensity assigned to hydroxyl groups in MgF2 decreases for another contribution at a chemical shift of +9 ppm. This band is assigned to the interaction between vanadium oxyfluoride and protons and confirms the chemical reaction between the OH groups and VOF3. 51V NMR spectra recorded at different spinning rates in order to distinguish between isotropic and spinning bands (Figure 4c) reveal the presence of 3 isotropic bands for raw and post-fluorinated MgF2 exposed to VOF3. The isotropic bands at −561, −615 and −791 ppm refer to VO2F2, VOF3 and VOF4, respectively [36,37,38].
The presence of VO2F2 ions evidences that VOF3 reacts with the OH groups on the surface of the chemical filter to form MgF2−x(VO2F2)x. Assuming that the vanadium oxyfluoride anion coordinates Mg2+ via oxygen and occupies a distorted tetrahedral site, the steric hindrance to hydroxyl groups is radically different and VO2F2 can substitute OH groups only at the surface and not in the rutile network. From an electronic point of view, V5+ is a second–order Jahn-Teller ion exhibiting a strong polyhedral distortion. This implies that this molecular species can accommodate a high distortion which permits the stabilization of MgF2−x(VO2F2)x compositions [39].
The generated HF increases the pressure in the reactor and can react with the VOF3. The resulting product is HVOF4 (trapped on the surface of the filter) in accordance with the isotropic band observed at −791 ppm. As expected, no isotropic band is detected for the oxygen-free MgF2 because of the low quantity of vanadium trapped. VOF3 sorption can be summarized as:
M g F 2 x O H x ( s ) + x V O F 3 ( g ) M g F 2 x V O 2 F 2 x ( s ) + x H F ( g )
In addition to this expected reaction on the metal fluoride surface (chemisorption), a physisorption mechanism may also occur because some of the vanadium is trapped by oxygen-free MgF2 (1.3 w.%).
Chemisorption involves the Lewis basicity of MgF2 through OH groups. This basicity may be tailored both via the nature, i.e., the coordination number of OH/F anions, the content of OH groups and the cation associated with fluorine. To go further in the discussion, the combination of polarizable cations with low electronegativity (K+ in KMgF3, Mg2+, Ca2+) and OH groups substituting F ions is another route to control the strength and number of Lewis basic sites keeping in mind the Lewis acidity of the gas that must be removed. The surface concentration and strength of the Lewis basic sites of the filter can be fitted to the Lewis acidity of the target gas in a gaseous mixture. Considering the structural features of the fluoride series: CaF2 with a fluorite-type structure and fluorine atoms coordinated fourfold to Ca2+, MgF2 with a rutile-type structure with fluorine atoms coordinated threefold to Mg2+ and KMgF3 with a perovskite-type structure and fluorine atoms twice coordinated to Mg2+, a large variation in the concentration and strength of the Lewis basicity sites is expected. Furthermore, the concentration of OH groups can be adjusted by post-fluorination.

3.3. Regeneration of the Chemical Filter

The hydroxyl groups were removed upon exposure to VOF3 for MgF2. With the aim to regenerate the chemical filter a two-step process was investigated: the first step consisted of the removal of vanadium with a solvent for V species (cleaning), whereas the second involved a rinsing of the surface. The investigations were performed on a sample with V content of 11,000 ppm after exposure of VOF3 (ICP data). The extraction rate is given as a function of the cleaning/rinsing agent pair (Figure 5a). A high yield extraction (95%) was achieved with HNO3 as both cleaning and rinsing agent. It is to note that this removal occurred without significant losses of MgF2 (only 4 w.%, Figure 5b). After removal of the vanadium, the number of basic sites (OH) must be recovered via the regeneration process. To reach this goal, MgF2 may be treated with NaOH.
19F MAS NMR is once again a powerful tool to evidence and quantify the presence of OH groups (Figure 6b) in addition to XRD that proves the unchanged presence of the MgF2 structure without significant change in Mg(OH)2 (Figure 6a).
The chemical compositions extracted from the fit of the NMR spectra according to the method described before (Figure 6c) are: MgF1.79(OH)0.21, MgF1.72(OH)0.28, MgF1.65(OH)0.35 as a function of the NaOH concentration, i.e., 0.25, 0.5 and 1 M, respectively. The number of basic sites can then be tailored. In order to nearly recover the initial O/F ratio of 0.25 (MgF1.6(OH)0.4), the concentration must be 1 M (O/F = 0.21); the duration of the treatment is 1 h at 60 °C. Two VOF3 filtering/regeneration cycles were carried out and the change in specific surface area was studied. Table 2 shows the slight decrease in the BET surface after a full filtering/regeneration cycle.
Whereas the specific surface area decreased after the exposure to VOF3, the initial value was nearly recovered after regeneration with NaOH (1 M concentration, 1 h at 60 °C). Such characteristics prove the possibility of regeneration of the selective filter for both reuse and recovery of vanadium. When the selective filter is used for the purification of UF6, some uranium species will be present on the surface of the filter; the regeneration aims to remove these species too, underlining its primary importance.

4. Conclusions

Metal fluorides have been investigated as selective chemical filters for the removal of VOF3, a model gas for volatile fluorides. HSA MgF2 containing different contents of OH groups and an oxygen-free MgF2 with a rather high specific surface area (35 m2 · g−1) were tested. It is worth noting that such fluorination synthesis using MgB2 precursor is reported for the first time. The sorption mechanism identified for MgF2 consists of a chemical reaction between VOF3 and Lewis basic sites, i.e., OH groups. The higher the amount of OH groups, the higher the quantity of vanadium trapped. Without hydroxyl groups (free-oxygen MgF2), physisorption is possible but the amount of vanadium is lower than that of a raw and post-fluorinated commercial MgF2. It was expected that hydroxyl groups are involved in the reaction with VOF3 but our data evidence physisorption too. Moreover, understanding the sorption mechanism using complementary techniques allowed us to select the most promising selective and regenerable filter. Both the amounts and strength of the Lewis basic sites may be tailored using a post-fluorination treatment with F2 gas. This versatility opens the route for the selectivity of filtering for mixtures of volatile fluorides according to the Lewis acidity of the target gas, and the present materials can be considered for active materials of gas sensors [40,41,42]. Since the reactions with the impurity to be removed occur at the surface of the HSA MgF2, the metal fluoride matrix is maintained and controlled regeneration of the hydroxyl groups with treatment in NaOH solution is possible.

Author Contributions

Conceptualization, L.J., A.S. and M.D.; methodology, L.J., A.S. and M.D.; investigation, L.J., J.-M.H., E.P. and M.D.; resources, A.S. and B.M.; data curation, L.J. and M.D.; writing—original draft preparation, L.J., J.-M.H. and M.D.; writing—review and editing, L.J. and M.D.; supervision, L.J., A.S., B.M. and M.D.; project administration, L.J. and M.D.; funding acquisition, L.J. and M.D. All authors have read and agreed to the published version of the manuscript.

Funding

This project was undertaken through a University of Clermont Auvergne PhD funded by ORANO Group under a partnership contract with the CNRS.

Data Availability Statement

Data will be made available on request.

Acknowledgments

We thank Alain Demourgues (ICMCB) for valuable discussion about acidity/basicity of fluoride compounds.

Conflicts of Interest

J.-M.H.’s thesis work was funded by ORANO group. A.S. and B.M. work for ORANO group. The funding sponsors had no role in the collection, analyses, or interpretation of data and in the decision to publish the results.

References

  1. Langeslay, R.R.; Kaphan, D.M.; Marshall, C.L.; Stair, P.C.; Sattelberger, A.P.; Delferro, M. Catalytic Applications of Vanadium: A Mechanistic Perspective. Chem. Rev. 2019, 119, 2128–2191. [Google Scholar] [CrossRef] [PubMed]
  2. Landström, L.; Lu, J.; Heszler, P. Size-Distribution and Emission Spectroscopy of W Nanoparticles Generated by Laser-Assisted CVD for Different WF6/H2/Ar Mixtures. J. Phys. Chem. B 2003, 107, 11615–11621. [Google Scholar] [CrossRef]
  3. Seghete, D.; Rayner, G.B., Jr.; Cavanagh, A.S.; Anderson, V.R.; George, S.M. Molybdenum Atomic Layer Deposition Using MoF6 and Si2H6 as the Reactants. Chem. Mater. 2011, 23, 1668–1678. [Google Scholar] [CrossRef]
  4. Wolden, C.A.; Pickerell, A.; Gawai, T.; Parks, S.; Hensley, J.; Way, J.D. Synthesis of β-Mo2C Thin Films. ACS Appl. Mater. Interfaces 2011, 3, 517–521. [Google Scholar] [CrossRef] [PubMed]
  5. Nikolaev, N.S.; Sadikova, A.T. Solubility of fluorides of certain elements in liquid uranium hexafluoride. Sov. Atom. Energy 1975, 39, 982–987. [Google Scholar] [CrossRef]
  6. Tananaev, I.V.; Nikolaev, N.S.; Luk’yanychev, Y.A.; Opalovskii, A.A. The chemistry of uranium fluorides. Russ. Chem. Rev. 1961, 30, 654–671. [Google Scholar] [CrossRef]
  7. Kemnitz, E.; Groß, U.; Rüdiger, S.; Shekar, C.S. Amorphous Metal Fluorides with Extraordinary High Surface Areas. Angew. Chem. Int. Ed. 2003, 42, 4251–4254. [Google Scholar] [CrossRef]
  8. Kemnitz, E.; Groß, U.; Rüdiger, S.; Shekar, C.S. Amorphe Metallfluoride mit außergewöhnlich großer spezifischer Oberfläche. Angew. Chem. 2003, 115, 4383–4386. [Google Scholar]
  9. Rüdiger, S.; Kemnitz, E. The fluorolytic sol–gel route to metal fluorides—A versatile process opening a variety of application fields. Dalton Trans. 2008, 9, 1117–1127. [Google Scholar] [CrossRef]
  10. Rüdiger, S.; Eltanamy, G.; Groß, U.; Kemnitz, E. Real sol-gel synthesis of catalytically active aluminium fluoride. J. Sol-Gel Sci. Technol. 2007, 41, 299–311. [Google Scholar] [CrossRef]
  11. Rüdiger, S.; Groß, U.; Kemnitz, E. Non-aqueous sol–gel synthesis of nano-structured metal fluorides. J. Fluorine Chem. 2007, 128, 353–368. [Google Scholar] [CrossRef]
  12. Demourgues, A.; Penin, N.; Dambournet, D.; Clarenc, R.; Tressaud, A.; Durand, E. About MX3 and MX2 (Mn+ = Mg2+, Al3+, Ti4+, Fe3+; Xp− = F, O2−, OH) nanofluorides. J. Fluorine Chem. 2012, 134, 35–43. [Google Scholar] [CrossRef]
  13. Dambournet, D.; Demourgues, A.; Martineau, C.; Durand, E.; Majimel, J.; Legein, C.; Buzaré, J.Y.; Fayon, F.; Vimont, A.; Leclerc, H.; et al. Microwave Synthesis of an Aluminum Fluoride Hydrate with Cationic Vacancies: Structure, Thermal Stability, and Acidic Properties. Chem. Mater. 2008, 20, 7095–7106. [Google Scholar] [CrossRef]
  14. Dambournet, D.; Demourgues, A.; Martineau, C.; Durand, E.; Majimel, J.; Vimont, A.; Leclerc, H.; Lavalley, J.C.; Daturi, M.; Legein, C.; et al. Structural investigations and acidic properties of high surface area pyrochlore aluminium hydroxyfluoride. J. Mater. Chem. 2008, 18, 2483–2492. [Google Scholar] [CrossRef]
  15. Dambournet, D.; Demourgues, A.; Martineau, C.; Majimel, J.; Feist, M.; Legein, C.; Buzaré, J.Y.; Fayon, F.; Tressaud, A. Nanostructured Al-based Fluoride-Oxide Materials with a Core-Shell Morphology. J. Phys. Chem. C 2008, 112, 12374–12380. [Google Scholar] [CrossRef]
  16. Dambournet, D.; Eltanamy, G.; Vimont, A.; Lavalley, J.C.; Goupil, J.M.; Demourgues, A.; Durand, E.; Majimel, J.; Rüdiger, S.; Kemnitz, E.; et al. Coupling sol-gel synthesis and microwave-assisted techniques: A new route from amorphous to crystalline high-surface-area aluminium fluoride. Chem. Eur. J. 2008, 14, 6205–6212. [Google Scholar] [CrossRef]
  17. Dambournet, D.; Demourgues, A.; Martineau, C.; Pechev, S.; Lhoste, J.; Majimel, J.; Vimont, A.; Lavalley, J.C.; Legein, C.; Buzaré, J.Y.; et al. Nanostructured Aluminium Hydroxyfluorides Derived from β-AlF3. Chem. Mater. 2008, 20, 1459–1469. [Google Scholar] [CrossRef]
  18. Wojciechchova, M.; Zielinski, M.; Pietrowski, M. MgF2 as a non-conventional catalyst support. J. Fluorine Chem. 2003, 120, 1–11. [Google Scholar] [CrossRef]
  19. Wang, Y.; Gohari-Bajestani, Z.; Lhoste, J.; Auguste, S.; Hémon-Ribaud, A.; Body, M.; Legein, C.; Maisonneuve, V.; Guiet, A.; Brunet, S. The Effects of Various Parameters of the Microwave-Assisted Solvothermal Synthesis on the Specific Surface Area and Catalytic Performance of MgF2 Nanoparticles. Materials 2020, 13, 3566. [Google Scholar] [CrossRef] [PubMed]
  20. Schreiber, H.; Wang, J.; Wilkinson, S.J. Durable MgO–MgF2 Composite Film for Infrared Anti-Reflection Coatings. European Patent EP2715412, 2012. [Google Scholar]
  21. Lisitsyn, V.M.; Lisitsyna, L.A.; Popov, A.I.; Kotomin, E.A.; Abuova, F.U.; Akilbekov, A.; Maier, J. Stabilization of primary mobile radiation defects in MgF2 crystals. Nucl. Instrum. Methods Phys. Res. Sect. B 2016, 374, 24–28. [Google Scholar] [CrossRef]
  22. Cappellini, G.; Furthmüller, J.; Bechstedt, F.; Botti, S. Electronic and Optical Properties of Alkaline Earth Metal Fluoride Crystals with the Inclusion of Many-Body Effects: A Comparative Study on Rutile MgF2 and Cubic SrF2. Symmetry 2023, 15, 539. [Google Scholar] [CrossRef]
  23. Golliher, W.R. Process for Separation and Recovery of Volatile Fluoride Impurities from Uranium Hexafluoride Containing the Same. U.S. Patent 3165376, 12 January 1962. [Google Scholar]
  24. Watanabe, D.; Sasakira, A.; Hoshino, K.; Kawamura, F. Adsorption of molybdenum hexafluoride on magnesium difluoride for uranium purification in FLUOREX reprocessing. J. Nucl. Sci. Technol. 2011, 48, 1413–1419. [Google Scholar] [CrossRef]
  25. Seregin, M.B.; Mikhalichenko, A.A.; Kuznetsov, A.Y.; Sokovin, G.S.; Chekmarev, A.M. Sorption of gaseous RuF5 on granulated fluoride sorbents. Radiochemistry 2011, 53, 288–291. [Google Scholar] [CrossRef]
  26. Wuttke, S.; Vimont, A.; Lavalley, J.-C.; Daturi, M.; Kemnitz, E. Infrared Investigation of the Acid and Basic Properties of a Sol-Gel Prepared MgF2. J. Phys. Chem. C 2010, 114, 5113–5120. [Google Scholar] [CrossRef]
  27. Rodriguez-Carvajal, J. Recent advances in magnetic structure determination by neutron powder diffraction. Phys. B Condens. Matter 1993, 192, 55–69. [Google Scholar] [CrossRef]
  28. Dhaoudadi, H.; Chaabane, H.; Touati, F. Nanorods Synthesized by A Facile. Hydrothermal Method in the Presence of CTAB. Nano-Micro Lett. 2011, 3, 153–159. [Google Scholar] [CrossRef]
  29. Hunt, G.R.; Perry, C.H.; Ferguson, J. Far-Infrared reflectance and transmittance of Potassium Magnesium fluoride and magnesium fluoride. Phys. Rev. A 1964, 134, A688–A691. [Google Scholar] [CrossRef]
  30. Heimer, N.E.; Del Sesto, R.E.; Meng, Z.; Wilkes, J.S.; Carper, W.R. Vibrational spectra of imidazolium tetrafluoroborate ionic liquids. J. Mol. Liq. 2006, 124, 84–95. [Google Scholar] [CrossRef]
  31. Wuttke, S.; Scholtz, G.; Rudiger, S.; Kemnitz, E. Variation of sol-gel synthesis parameters and their consequence for the surface area and structure of magnesium fluoride. J. Mat. Chem. 2007, 17, 4980–4988. [Google Scholar] [CrossRef]
  32. Scholtz, G.; Stosiek, C.; Noack, J.; Kemnitz, E. Local fluorine environments in nanoscopic magnesium hydr(oxide) fluorides studied by F-19 MAS NMR. J. Fluor. Chem. 2011, 132, 1079–1085. [Google Scholar] [CrossRef]
  33. Zidan, M.D.; Allaf, A.W. The gas-phase on-line production of vanadium oxytrihalides, VOX3 and their identification by infrared spectroscopy. Spectrochim. Acta Part A 2000, 56, 2693–2698. [Google Scholar] [CrossRef] [PubMed]
  34. Beattie, I.R.; Livingston, K.; Reynolds, D.J.; Ozin, G.A. Vibrational spectra of some oxide halides of the transition elements with particular reference to gas-phase and single-crystal Raman spectroscopy. J. Chem. Soc. A 1970, 1210–1216. [Google Scholar] [CrossRef]
  35. Clark, R.J.H.; Rippon, D.M. The vapour phase Raman spectra, Raman band contour analyses, Coriolis constants, force constants, and values for thermodynamic functions of the symmetric top molecules POF3, POCl3, VOF3, VOCl3, PSCl3 and FClO3. Mol. Phys. 1974, 28, 305–319. [Google Scholar] [CrossRef]
  36. Davis, M.F.; Jura, M.; Leung, A.; Levason, W.; Littlefield, B.; Reid, G.; Webster, M. Synthesis, chemistry and structures of complexes of the dioxovanadium(v) halides VO2F and VO2Cl. Dalton Trans. 2008, 44, 6265–6273. [Google Scholar] [CrossRef] [PubMed]
  37. Hibbert, R.C. A 51V and 19F n.m.r. study of vanadium(V) oxide fluoride complexes displaying vanadium–fluorine coupling. J. Chem. Soc. Chem. Commun. 1985, 317–318. [Google Scholar] [CrossRef]
  38. Howell, J.A.S.; Moss, K.C. Nuclear magnetic resonance studies on fluorine-containing compounds. Part III. The tetrafluoro-oxovandate(V) ion. J. Chem. Soc. A 1971, 270–272. [Google Scholar] [CrossRef]
  39. Senchyk, G.A.; Bukhanko, V.O.; Lysenko, A.B.; Krautscheid, H.; Rusanov, E.B.; Chernega, A.N.; Karbowiak, M.; Domasevitch, K.V. AgI/VV Heterobimetallic Frameworks Generated from Novel-Type {Ag2(VO2F2)2(triazole)4} Secondary Building Blocks: A New Aspect in the Design of SVOF Hybrids. Inorg. Chem. 2012, 51, 8025–8033. [Google Scholar] [CrossRef]
  40. Chen, C.; Jiang, M.; Luo, X.; Tai, H.; Jiang, Y.; Yang, M.; Xie, G.; Su, Y. Ni-Co-P hollow nanobricks enabled humidity sensor for respiratory analysis and human-machine interfacing. Sens. Actuators B Chem. 2022, 370, 132441. [Google Scholar] [CrossRef]
  41. Chen, C.; Xie, G.; Dai, J.; Li, W.; Cai, Y.; Li, J.; Zhang, Q.; Tai, H.; Jiang, Y.; Su, Y. Integrated core-shell structured smart textiles for active NO2 concentration and pressure monitoring. Nano Energy 2023, 116, 108788. [Google Scholar] [CrossRef]
  42. Zegebreal, L.T.; Tegegne, N.A.; Hone, F.G. Recent progress in hybrid conducting polymers and metal oxide nanocomposite for room-temperature gas sensor applications: A review. Sens. Actuators A Phys 2023, 359, 114472. [Google Scholar] [CrossRef]
Figure 1. XRD patterns (a), FTIR (b), MAS 30 kHz 19F (c) and 1H (d) spectra and N2 adsorption isotherms (e) of raw and post-fluorinated commercial MgF2 and oxygen-free MgF2. The (*) marks spinning side bands.
Figure 1. XRD patterns (a), FTIR (b), MAS 30 kHz 19F (c) and 1H (d) spectra and N2 adsorption isotherms (e) of raw and post-fluorinated commercial MgF2 and oxygen-free MgF2. The (*) marks spinning side bands.
Surfaces 06 00032 g001aSurfaces 06 00032 g001b
Figure 2. FTIR spectra in the solid (a) and gaseous phase (c) of adsorbed species by MgF2−x(OH)x after exposure to VOF3, evolution of pressure in the gas chamber as a function of time (b) and XRD (d) patterns of resulting powders.
Figure 2. FTIR spectra in the solid (a) and gaseous phase (c) of adsorbed species by MgF2−x(OH)x after exposure to VOF3, evolution of pressure in the gas chamber as a function of time (b) and XRD (d) patterns of resulting powders.
Surfaces 06 00032 g002
Figure 3. Raman spectra of varied magnesium difluoride samples exposed to VOF3.
Figure 3. Raman spectra of varied magnesium difluoride samples exposed to VOF3.
Surfaces 06 00032 g003
Figure 4. 19F (a), 1H (b) and 51V (c) MAS NMR spectra (spinning rate of 30 kHz) before and after exposure to VOF3. Arrows mark the isotropic lines.
Figure 4. 19F (a), 1H (b) and 51V (c) MAS NMR spectra (spinning rate of 30 kHz) before and after exposure to VOF3. Arrows mark the isotropic lines.
Surfaces 06 00032 g004
Figure 5. (a) Extraction rate of vanadium and (b) vanadium weight loss according to the cleaning and rinsing agents.
Figure 5. (a) Extraction rate of vanadium and (b) vanadium weight loss according to the cleaning and rinsing agents.
Surfaces 06 00032 g005
Figure 6. XRD patterns (a) and 19F MAS NMR spectra (b) of raw MgF2 treated with NaOH solutions; (c) 19F MAS NMR spectra of post-treated MgF2 exposed to VOF3, after cleaning (with HNO3) and rinsing (H2O) at 60 °C and regeneration with NaOH.
Figure 6. XRD patterns (a) and 19F MAS NMR spectra (b) of raw MgF2 treated with NaOH solutions; (c) 19F MAS NMR spectra of post-treated MgF2 exposed to VOF3, after cleaning (with HNO3) and rinsing (H2O) at 60 °C and regeneration with NaOH.
Surfaces 06 00032 g006
Table 1. Weight uptake after VOF3 sorption experiments.
Table 1. Weight uptake after VOF3 sorption experiments.
MgF2 TreatmentBulk Weight Uptake (%)Vanadium Weight Uptake (%)Vanadium ICP Analysis (%)
Raw8.73.63.4
Post-fluorinated at 240 °C5.12.12.0
Synthesized from MgB23.11.31.2
Table 2. Change in the specific surface area during two filtering/regeneration sequences.
Table 2. Change in the specific surface area during two filtering/regeneration sequences.
Process StepStarting MaterialAfter FilteringAfter NaOH Regeneration
SSABET (m²·g−1)First run723764
Second run643851
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jouffret, L.; Hiltbrunner, J.-M.; Petit, E.; Selmi, A.; Morel, B.; Dubois, M. Selective Chemical Filters for VOF3: Tailoring MgF2 Filter Selectivity through Surface Chemistry. Surfaces 2023, 6, 480-492. https://doi.org/10.3390/surfaces6040032

AMA Style

Jouffret L, Hiltbrunner J-M, Petit E, Selmi A, Morel B, Dubois M. Selective Chemical Filters for VOF3: Tailoring MgF2 Filter Selectivity through Surface Chemistry. Surfaces. 2023; 6(4):480-492. https://doi.org/10.3390/surfaces6040032

Chicago/Turabian Style

Jouffret, Laurent, Jean-Michel Hiltbrunner, Elodie Petit, Ania Selmi, Bertrand Morel, and Marc Dubois. 2023. "Selective Chemical Filters for VOF3: Tailoring MgF2 Filter Selectivity through Surface Chemistry" Surfaces 6, no. 4: 480-492. https://doi.org/10.3390/surfaces6040032

Article Metrics

Back to TopTop