Manuel
Souto
a,
Andrea
Santiago-Portillo
b,
Miguel
Palomino
c,
Iñigo J.
Vitórica-Yrezábal
d,
Bruno J. C.
Vieira
e,
João C.
Waerenborgh
e,
Susana
Valencia
c,
Sergio
Navalón
b,
Fernando
Rey
c,
Hermenegildo
García
bc and
Guillermo
Mínguez Espallargas
*a
aInstituto de Ciencia Molecular (ICMol), Universitat de València, c/Catedrático José Beltrán, 2, 46980 Paterna, Spain. E-mail: guillermo.minguez@uv.es
bDepartamento de Química, Universitat Politècnica de València, c/Camino de Vera, s/n, 46022, Valencia, Spain
cInstituto de Tecnología Química (UPV-CSIC), Universitat Politècnica de València-Consejo Superior de Investigaciones Científicas, Av. De los Naranjos s/n, 46022, Valencia, Spain
dSchool of Chemistry, University of Manchester, Oxford Road, Manchester, M139PL, UK
eCentro de Ciências e Tecnologias Nucleares, Instituto Superior Técnico, Universidade de Lisboa, 2695-066 Bobadela LRS, Portugal
First published on 24th January 2018
Herein we report the synthesis of a tetrathiafulvalene (TTF)-based MOF, namely MUV-2, which shows a non-interpenetrated hierarchical crystal structure with mesoporous one-dimensional channels of ca. 3 nm and orthogonal microporous channels of ca. 1 nm. This highly stable MOF (aqueous solution with pH values ranging from 2 to 11 and different organic solvents), which possesses the well-known [Fe3(μ3-O)(COO)6] secondary building unit, has proven to be an efficient catalyst for the aerobic oxidation of dibenzothiophenes.
On the other hand, tetrathiafulvalene (TTF) and its numerous derivatives are among the most versatile molecules which exhibit interesting redox properties, electron–donor character and potential application as molecular conductors.17 The use of TTF as a ligand for the design of porous coordination polymers can give rise to multifunctional materials combining different physical properties.18–20 For example, Dincă and coworkers have recently reported the use of the ligand tetrathiafulvalene tetrabenzoic acid (H4TTFTB) with various transition metals (II) obtaining a family of isostructural and microporous TTF-based MOFs exhibiting tunable electrical conductivity.21,22 More recently, the same ligand has also been used for the preparation of a TTF-based MOF with Mg(II), which exhibits permanent mesopores.23
Herein, we report the synthesis, structure determination and physical properties of MUV-2 (MUV: Materials of the University of Valencia), a highly stable TTF-based MOF with a unique non-interpenetrated hierarchical crystal structure with the mesoporous channels orthogonal to the microporous channels. Moreover, the advantages of MUV-2 with respect to widely used MOF catalysts will be clearly demonstrated for a reaction of great interest in the field, illustrating the advantages of having a hierarchical MOF with large mesopores and high stability.
Single-crystal X-ray diffraction data were collected with up to 1 Å resolution at the I19 beamline facilities at the Diamond Light Source (UK). MUV-2 crystallises in the space group P2m and the unit cell parameters are a = b = 33.3 Å and c = 12.4 Å and it consists of 6-connected [Fe3(μ3-O)(COO)6] SBUs and tetratopic TTFTB ligands. Considering each TTFTB ligand as a four-connected node and each Fe3O(COO)6 unit as a six-connected node, MUV-2 can be simplified as a 4,6-connected network with ttp topology (Fig. S2†), an unusual topology previously observed in two lanthanoid-based MOFs.25,26 The non-interpenetrated crystal structure reveals large hexagonal mesoporous 1-D channels of ca. 3 nm along the c-axis, which are formed by six TTFTB ligands and six clusters [Fe3(μ3-O)(COO)6] (Fig. 1a). TTF moieties are significantly twisted around the central CC bond with a dihedral angle of 20°, whereas the planes formed by the two dithiole rings (planes S1–C1–C2–S2 and S3–C5–C6–S4) have a dihedral angle of 41° (Fig. S3†). The torsion angles of S2–C3–S1–C1 and C1–C2–S2–C3 are 17° and 11°, respectively, which are typical for neutral TTFs. The phenyl rings exhibit large distortions with respect to the TTF core with dihedral angles of 62° with the latter. In contrast to NU-1000, where the microporous channels run parallel to the mesoporous ones,15 the crystal structure of MUV-2 shows that microporous channels of ca. 1 nm (9.5 × 12 Å) are orthogonal to the mesoporous channels and are formed between two TTFTB ligands and two [Fe3(μ3-O)(COO)6] clusters (Fig. 1b). In addition, microporous cages consist of three TTFTB ligands and two [Fe3(μ3-O)(COO)6] SBUs (Fig. 1c) leading to a remarkable open structure with a calculated free volume of ca. 82%. Note that the crystal structure of MUV-2 contains the precursor [Fe3O(CH3COO)6]ClO4 within the pores since it was determined from the as-synthesised material without the washing and activation procedure.
The results obtained from Mössbauer spectroscopy, magnetic measurements, solid-state cyclic voltammetry and Raman spectroscopy are consistent with the [Fe3(μ3-O)(COO)6] cluster being formed by three S = 5/2 Fe(III) ions in octahedral environments, and the TTF cores being neutral, thus yielding a material with the formula (TTFTB)3[(Fe3O)(H2O)2(OH)]2, which is in agreement with the EDAX analysis of MUV-2 (see Fig. S4–S9†). It is important to note that one of the three coordinated H2O molecules in the cluster is present as a negatively charged hydroxide (OH−) in order to maintain the charge balance.
Thermogravimetric analysis (TGA) of washed MUV-2 exhibited a sharp mass loss of 20% between 25 and 100 °C, which corresponds to the elimination of solvent molecules (Fig. S10†). TGA shows a large plateau above 200 °C until the final decomposition at 350 °C. Activation of MUV-2 was performed by heating the washed material at 150 °C for 2 h. Its crystallinity was confirmed by powder X-ray diffraction (PXRD) and it was observed that the principal peak was slightly shifted to 3.4° upon heating (Fig. 2) and recovered to the initial PXRD pattern upon resolvation (Fig. S11†). Additionally, MUV-2 shows extraordinary chemical stability in aqueous solution with pH values ranging from 2 to 11 and in different organic solvents for 24 h. The PXRD patterns showed that crystallinity is maintained under these conditions (Fig. S12†) and the CO2 adsorption capacity is well preserved, for example, after treatment with pH = 2 and 11 aqueous solutions (Fig. S20†) with only a minor reduction.
The N2 adsorption isotherm at 77 K revealed a combination of Type I and IV isotherms, resulting from the presence of micropores and mesopores, respectively. Thus, a steep N2 adsorption occurs at low p/p0, while a slight secondary uptake was also found due to the mesopores filling. A plateau was observed in the N2 uptake of 16 mmol g−1 (Fig. S13†). MUV-2 has a BET surface area of 1220 m2 g−1, which is higher than those for other reported mesoporous TTF-based MOFs.23 A micropore volume of 0.52 cm3 g−1 was found using the Dubinin–Radushkevich equation and the pore size analysis by means of the Barrett–Joyner–Halenda (BJH) method revealed a pore size of 38.7 Å (Fig. S15†). Fig. 3 shows the CO2 and CH4 isotherms at 298 K, revealing a high sorption capacity for both gases comparable to that of MIL-100.27
Fig. 3 Gas adsorption isotherms of CO2 (black) and CH4 (red) on MUV-2 at 298 K (lines correspond to the best fits). Data at other temperatures are shown in the ESI.† |
The isosteric heat of adsorption (qst) of CO2 decreases from 30 to 20 kJ mol−1, and remains constant in the case of CH4 at around 16 kJ mol−1 within the studied loading range (Fig. S18†). These values clearly indicate the higher affinity of MUV-2 for CO2 than for CH4. The isosteric heat of adsorption of CO2 at zero coverage (q0st) of MUV-2 is comparable to that of an LTA zeolite with a Si/Al ratio of around 6,28 and to that of MIL-101, and is in the same range as a wide variety of MOFs.29
The superior catalytic activity of MUV-2 due to the presence of mesopores with respect to widely used MIL MOFs as heterogeneous catalysts was clearly evidenced for the aerobic oxidation of dibenzothiophene (DBT) using long chain alkanes as solvents (Scheme 2). DBT is a model compound of the harmful aromatic sulphur compounds present in diesel.16 Legal regulations require diminishing the sulphur content in diesel down to the ppb scale. One possibility is to perform fuel oxidation to convert the sulphur-containing organic compounds to the corresponding sulfones (generally soluble in water) that can be removed from the fuel by washing. It has recently been reported that DBT can be oxidized by molecular oxygen to the corresponding sulfone (DBTO2) using MIL-101(Cr or Fe) as the solid catalyst,16 although an induction period, probably related to diffusion problems, was observed.
Fig. 4 shows the time conversion plots for DBT disappearance and DBTO2 formation in n-dodecane (plotted as sulphur content) comparing the temporal profile using MUV-2, MIL-101 (Fe) and MIL-100 (Fe), and it shows that MUV-2 is the best performing catalyst. Since all three MOFs contained the same type of Fe3-μ3O active centre, the higher catalytic activity of MUV-2 can be attributed to the more favourable diffusion due to the presence of large pores in this material (see Table S3†), as demonstrated with three different control experiments. To gain understanding on the origin of the induction period and its dependency on the pore size of MUV-2, this solid was contacted with n-dodecane containing DBT in the absence of O2 for 2 h, and then O2 was introduced into the flask, whereby an immediate oxidation of DBT without an induction period was observed (see Fig. S21a†). A similar observation, i.e. a lack of induction period, was also noted when MUV-2 was boiled in n-dodecane containing O2 and DBT was added two hours later (see Fig. S21b†). Finally, no reduction of the induction period is observed if MUV-2 is heated in n-dodecane for two hours before introducing O2 and DBT to the system (see Fig. S21c†). This rationalisation is in agreement with the fact that, besides higher reaction rates, the induction period is remarkably shortened to about 1 h using MUV-2.
From an application point of view, desulfuration of diesel should be carried out in the presence of a mixture of hydrocarbons from C13 to C18. Since diffusion is a limiting factor under this condition, the changes from model n-dodecane as the solvent to real diesel were accompanied by a considerable decrease in activity in the case of MIL-101.16 It is of interest, therefore, to determine the performance of MUV-2 for DBT aerobic oxidation using diesel as the solvent. The results presented in Fig. 4b and d provide a comparison of the time–conversion plots for DBT oxidation using MIL-101 (Fe), MIL-100 (Fe) and MUV-2. As is observed, the difference in catalytic activity when using diesel as the solvent remarkably favors MUV-2, and shows the advantages of this MOF under these conditions. Reusability and productivity tests also show the high stability of MUV-2 as a solid catalyst (Fig. S22 and S23†), which is also active for different DBT derivatives (4-MeDBT and 4,6-Me2DBT) (Fig. S24†).
Hot-filtration tests were performed by filtering MUV-2 out from the hot reaction mixture after 2 h. At this point, the sulphur content was about 125 ppm (cf. the initial 200 mg L−1), and the clear supernatant was allowed to continue to react in the absence of solid particles, and we observed a very minor progress of about 20 mg L−1 sulphur content decrease in the subsequent 3 h (Fig. S25c†). In contrast, a twin reaction where MUV-2 was not filtered achieved complete sulphur removal from the initial 200 mg L−1 in 5 h. These results indicate that after initiation of the reaction only a very minor contribution of leaching and homogeneous oxidation is present. In addition, control experiments that used either chromium(III) acetate (0.6 mg of Cr), the preformed [Fe3O(CH3COO)6]ClO4 at the loadings corresponding to those present in MUV-2 or leached out during the reaction, or H4TTFTB as homogeneous catalysts showed negligible conversion in all cases, indicating that Cr and Fe transition metals at these concentrations and the ligand are not able to promote DBT oxidation (Fig. S25†).
A combination of quenching experiments and spectroscopic studies has been used to address the reaction mechanism and, in particular, to determine that the primary reactive oxygen species responsible for oxidation is HOO˙ (see Fig. S26 and S27†). Thus, performing the oxidation in the presence of DMSO, a selective hydroxyl radical scavenger, does not influence the time–conversion plot much, while, in contrast, the presence of p-benzoquinone, a selective quencher of superoxide and hydroperoxyl radicals, strongly inhibits DBT oxidation to DBTO2. In addition, admission of oxygen into thermally dehydrated MUV-2 (220 °C, 5 h) at 140 °C led to the appearance of two new vibration bands in the Raman spectra at 1502 and 1161 cm−1 that can be attributed to physisorbed O2 and Fe–O–O, respectively (Fig. S27†). This metal-peroxo could abstract a hydrogen atom from the medium (n-dodecane), generating hydroperoxyl radicals that would initiate DBT oxidation. This hypothesis is supported by the observation of some very minor undetermined oxidation products from n-dodecane.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 1579606 (MUV-2). For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c7sc04829g |
This journal is © The Royal Society of Chemistry 2018 |