Antonia M. Rasero-Almansa
a,
Marta Iglesias*a and
Félix Sánchezb
aInstituto de Ciencia de Materiales de Madrid, ICMM-CSIC, Sor Juana Inés de la Cruz, 3, Cantoblanco, 28049 Madrid, Spain. E-mail: marta.iglesias@icmm.csic.es
bInstituto de Química Orgánica, IQOG-CSIC, Juan de la Cierva, 3, 28006 Madrid, Spain
First published on 3rd November 2016
Bimetallic Zr(Ti)-NDC based metal–organic frameworks (MOFs) have been prepared by incorporation of titanium(IV) into zirconium(IV)-NDC-MOFs (UiO family). The resulting materials maintain thermal (up to 500 °C), chemical and structural stability with respect to parent Zr-MOFs as can be deduced from XRD, N2 adsorption, FTIR and thermal analysis. The materials have been studied in Lewis acid catalyzed reactions, such as, domino Meerwein–Ponndorf–Verley (MPV) reduction–etherification of p-methoxybenzaldehyde with butanol, isomerization of α-pinene oxide and cyclization of citronellal.
A titanium MOF is one of the most interesting candidates due to a large abundance of the metal, cheap price, low toxicity and photocatalytic properties. However, to date, scarce porous titanium MOFs have been reported, except MIL-125 and NTU-9. The difficulties to synthesize titanium MOFs directly prompted us to search for new synthetic strategies and PSE could be an alternative to obtain certain Ti-MOFs that cannot or are difficult to be synthesized directly.
Recently, it has been reported that a series of porous titanium MOFs can be successfully synthesized by the PSE strategy.13 The incorporation of the smaller Ti atom in the water and high temperature stable Zr-based UiO-66, enhanced adsorption capacity14 and maintains the very high, thermal, chemical15 and structural stability of UiO-66 during water adsorption/desorption cycles.16 Ti-UiO-66 also shows exceptional gas permeability in mixed matrix membranes.17 Moreover Cohen et al. have reported that a mixed-ligand (terephthalic acid/2-aminoterephthalic acid), mixed-metal (Zr–Ti)-UiO-66-derivative results an effective photocatalyst for CO2 reduction under visible light irradiation.18 Ti-substituted UiO-66(Ti)-NH2 also showed enhanced photocatalytic performance via a Ti-mediated electron transfer mechanism.19,20 In the present manuscript, we described a facile preparation of bimetallic Zr(Ti)-NDC-MOFs (NDC = 2,6-naphthalendicarboxylate, NDC-NH2 = 4-aminonaphthalen-2,6-dicarboxylic acid) by introduction of Ti(IV) ions into the preformed Zr-NDC framework. The resultant materials are good Lewis acid catalysts for cascade reactions involving the Meerwein–Ponndorf–Verley (MPV) reduction followed by etherification, reactions of cyclization of citronellal, and isomerization of α-pinene oxide.21
O) band at 1660 cm−1 (Fig. S6†).
The statistical incorporated Ti(IV) (Zr6−xTixO4(OH)4(NDC)6) was quantified by total X-ray fluorescence (TXRF) spectroscopy being the Zr/Ti ratio 2.6 for Zr(Ti)-NDC and 2.5 for Zr(Ti)-NDC-NH2 (Table S1†) and the final metal compositions were Zr4.4Ti1.6 for Zr(Ti)-NDC and Zr4.3Ti1.7 for Zr(Ti)-NDC-NH2.
For mixed NDC-NH2-MOFs the ratio between ligands was obtained by elemental analysis of nitrogen and confirmed by UV-visible spectroscopy and corresponds to a 20% of amino ligand content. 1H NMR of digested Zr-NDC-MOFs confirmed the presence of both NDC and NDC-NH2 linkers (Fig. S13†). Hydrofluoric acid was employed to digest the materials, because of the high affinity of Zr for fluoride. Electrospray ionization-mass spectrometry (ESI-MS) of digested Zr-MOFs was also performed (Fig. S14†).
The Zr(Ti)-NDC-MOFs show excellent crystallinity, as evidenced by the powder X-ray diffraction (PXRD) (Fig. 1), the good agreement between the XRD patterns of Zr(Ti)-derivative and parent-Zr indicates that the framework of parent Zr-MOF is not collapsed and rules out the possibility of the formation of Ti-based impurities; moreover, no diffraction peaks belonged to other titanium species could be observed. UiO66(Ti)-NH2 and UiO67(Ti)-NH2 were also prepared following the same procedure for comparative purposes. As observed, peaks in XRD patterns of Zr(Ti)-NDC shifted to higher reflection angles (2θ = 6.55) as compared to original Zr-NDC (2θ = 6.43) (from 6.30 to 6.36 for Zr(Ti)-NDC-NH2), the 2θ value in UiO66(Ti)-NH2 (from 7.28 to 7.41)19 and UiO67(Ti)-NH2 shifts from 5.77 to 5.80 (Fig. S2†). This fact is also observed over inorganic solid solutions indicating that Ti(IV) ion (0.605 Å) substitutes a larger Zr(IV) ion (0.72 Å) in Zr–O oxo-clusters.23,24
The X-ray photoelectron spectroscopy (XPS) of the elements of interest, Zr 3d and Ti 2p, of Zr(Ti)NDC-MOF are shown in Fig. 2 (general survey can be observed in Fig. S12†). Fig. 2a presents the Ti 2p region showing two peaks at 458.7 eV and 464.5 eV, corresponding to Ti 2p3/2 and Ti 2p1/2 respectively, providing another confirmation of the successful incorporation of the Ti moiety in Zr-NDC. The spin energy separation is 5.8 eV. The peak position of Ti4+ is consistent with Ti4+ in an octahedral coordination environment.25–27 The Zr 3d was shown in Fig. 2b with two peaks at 182.7 and 185 eV, due to Zr 3d5/2 and Zr 3d3/2. These data are ver similar to that previously described for UiO66(Ti)-NH2 (ref. 19) and UiO66(Ti).20
The thermogravimetric analysis (TGA) of Zr(Ti)-NDC showed a high thermal stability for these compounds with a thermal decomposition temperature similar to that of Zr-NDC (>400 °C, Fig. S3†).
The porosity was investigated by the BET analysis (N2, 77k) of Zr(Ti)-NDC (1068 m2 g−1 which is comparable to that of the parent Zr-NDC) (∼1062 m2 g−1) (Table S2, Fig. S1†). PXRD patterns collected after the N2 sorption experiments indicate that all other samples remain intact after activation; although some peak broadening was observed (Fig. S4†).
Similar to parent Zr-NDC and Zr-NDC-NH2, the UV-visible diffuse-reflectance spectra of Zr(Ti)-NDC and Zr(Ti)-NDC-NH2 also show two main absorption peaks at around 290 and 350 nm (Fig. 3). However, the peak at 350 nm, corresponding to the absorption of Zr–O clusters is broader,19,20 that could be attributed to the absorption of new Ti–O oxo-groups (Fig. S5†). In addition to the peak at 350 nm, a shoulder to ca. 400 nm also appears in Zr-NDC-NH2 and Zr(Ti)-NDC-NH2 which can be assigned to the amino group. A new band at 650 nm is also observed for Zr(Ti)-NDC-NH2 probably due to a charge transfer. These results indicate that Ti(IV) has been successfully incorporated in Zr-NDC-MOF.
Zr(Ti)-NDCs show in the FTIR spectra the characteristic ν(Zr–O) bands at 682, 642 cm−1 and ν(Ti–O) at 665 cm−1 (Fig. S6†).
The solid-state photoluminescence spectra of the Zr-NDC compounds were studied at room temperature. As shown in Fig. 3, with excitation at 350 nm, the emission wavelength of the Zr-NDC28 is red-shifted to 405 nm (475 nm for Zr-NDC-NH2), while in the case of Zr(Ti)-NDC the emission wavelength appears at 424 nm (504 nm for Zr(Ti)-NDC-NH2). Thus, the introduction of a functional group (–NH2) on the organic ligand, and other metal as Ti improved the luminescent behavior.
13C CP/MAS-NMR spectra of evacuated Zr(Ti)-NDC samples (Fig. 4 and S8†) showed clearly all signals corresponding to the linkers at 172–171 (COO), 134.8–134.6 (CCOO), 132.8–132.5, 130.2–130.0, 127.4–127.1, and 125.7–125.4 ppm which are characteristic of the unique carbon atoms of NDC and NDC-NH2 linkers, respectively, C–NH2 resonance (for the mixed linker compounds) do not appears possibly due to their poor relaxation.
The textural morphology of Zr-NDC and Zr-NDC-NH2 reveal some changes on the surface of crystals (from the defects by replacing a 8-coordinated ZrIV ion for an 6-coordinated TiIV ion) although the reaction was performed at a temperature below 85 °C (Fig. S7†).
![]() | ||
| Scheme 2 Domino reaction of aldehydes with 2-butanol: MPV reduction of the aldehyde to alcohol with successive etherification. | ||
To explore the Lewis acid sites, we investigated the ammonia temperature-programmed desorption (NH3-TPD) (commonly used to determine the acidity of zeolites)35 of the Zr-NDC and Zr(Ti)-NDC, indicating the presence of acid sites in the catalysts (Fig. 5). Zr-NDC and Zr(Ti)-NDC showed weak to medium strength acidity.36,37
Brønsted acids can catalyze the formation of ethers when starting from two alcohol molecules,42 and sulfuric acid has been applied in homogeneous phase;43 but are limited to primary ethers. In contrast, water stable Lewis acid catalysts as Sn- and Zr-containing silicate molecular sieves with framework single, isolated metal sites are active and selective for catalyzing MPV reduction of aldehydes as well as etherification of alcohols.44 Thus, considering that Zr-NDC-MOFs show Lewis acidity, it could be of interest to find if they are able to promote efficiently the tandem reaction involving the MPV reduction of p-methoxybenzaldehyde with 2-butanol to p-methoxybenzyl alcohol, followed by its etherification with an excess of 2-butanol (see Scheme 2), and how they compare with other solid Lewis acids such as Sn-, Zr-, Ta-beta zeolites44,45 Results in Table 1 show that the desired fragrance was obtained with Zr-NDC and Zr(Ti)-NDC, being the Zr(Ti) catalyst more active for the global process (yield > 90%). The linker NDC-NDC-NH2 also has a beneficial effect on the final activity. The reaction proceeds under mild reaction conditions and all catalysts proved to be active for yielding the ether compound with good yield, and easy. The selectivity for the one-pot process is very high, and the desired ether was the only product observed. In this case, the intermediary alcohol rapidly reacts on the catalyst to give the corresponding ether. Substrate scope for the etherification reaction was explored by using Zr(Ti)-NDC as catalyst. Benzaldehyde readily reacted with 2-butanol to form the ether product 2, while aldehydes with electron withdrawing groups at the aromatic ring (see Table 1, entry 7) can be more difficult to react with Zr(Ti)-NDC.
| Catalyst | Aldehyde R | Conv.b (h) | Selectivityc (%) | ||
|---|---|---|---|---|---|
| Alcohol | Ether | ||||
| a Reaction conditions: aldehyde (0.16 mmol), 2-butanol (1 mL), catalyst (5 mg), MW oven (150 °C).b Analysed by GC-MS analysis.c Normalized to 100%.d Conventional reflux. | |||||
| 1 | Zr(Ti)-NDC | OCH3 | 90 (4) | 10 | 90 |
| 2 | Zr(Ti)-NDC-NH2 | OCH3 | 95 (4) | — | 100 |
| 3 | Zr(Ti)-NDC-NH2d | OCH3 | 65 (4), 98 (21) | 1.5 | 98.5 |
| 4 | Zr-NDC | OCH3 | 19 (4) | — | 100 |
| 5 | Zr-NDC-NH2 | OCH3 | 30 (4) | 7 | 93 |
| 6 | Zr(Ti)-NDC | H | 97 (4) | — | 100 |
| 7 | Zr(Ti)-NDC | F | 85 (4) | 58 | 42 |
| 8 | Zr-beta44 | OCH3 | 100 (8) | — | 100 |
| 9 | Sn-beta44 | OCH3 | 71 (8) | — | 100 |
| 10 | TiO2 | OCH3 | 0 (4) | — | — |
Zr(Ti) is the most active catalyst for the both isolated reactions MPV reaction and etherification. When Zr-NDC was tested for their catalytic activity in the MPV reduction of cyclohexanone with 2-butanol it shows practically no activity and conversions of 5% were obtained after 1 h of reaction time. On the other hand, conversions of 20 and 60% were obtained with Zr(Ti) after 1 and 4 h of reaction time. When the etherification of 1-butanol with p-methoxybenzyl alcohol is carried out, 15% conversion was observed for the Zr-NDC-catalyzed reaction and 83% for Zr(Ti)-NDC. These results together with our previous work,44,46 indicate that in the case of domino reactions catalyzed by Lewis acid sites, MOFs offer the possibility to maximize the final yield by introducing more than one type of Lewis acid site in the framework. Indeed, a catalyst with Zr and Ti in the MOF maximize each one of two steps of the process to give the highest yield of the desired ether (see Table 1). A blank experiment with TiO2 shows that it is inactive for this reaction and our experimental conditions (Table 1, entry 10).
Recycling experiments with Zr(Ti)-NDC-NH2 catalyst show that the activity decreases after three cycles, and longer reaction times were necessary to obtain a better selectivity to the ether (Fig. S11†).
Here, the citronellal cyclization to isopulegol has been studied with Zr-NDC and Zr(Ti)-NDC catalysts and the results compared with those obtained with other MOFs and inorganic solid acids as MCM-41(Ti), and Sn-beta zeolite. (±) Citronellal was submitted to a Zr-NDCs-catalyzed Prins reaction at different temperatures and catalyst loadings in toluene (each catalyst was previously activated at 493 K). Results in Table 2, show that in all cases, the selectivity for the four diastereomeric pulegols with respect to other products was >98% and selectivity for the desired isopulegol diastereomer was in the order of 65–75% with respect to the other three diastereomers.
| Entry | Catalyst | T (°C) | Cat. (mg) | Conv. (%) | Sel.b (%) | Diast.c (%) |
|---|---|---|---|---|---|---|
| a Reactions were carried out in toluene with citronellal (0.3 mmol, 58 μL) added to 5, 10, 20 mg of catalyst preactivated at 200 °C, reaction time: 24 h.b Selectivity for the four diastereomeric pulegols with respect to other products.c Selectivity for the isopulegol diastereomer with respect to the other three diastereomers.d This work. | ||||||
| 1 | Zr-NDC | 100 | 10 | 47 | 99 | 68 |
| 2 | Zr(Ti)-NDC | 100 | 10 | 48 | 99 | 75 |
| 3 | Zr-NDC | 100 | 20 | 60 | 99 | 62 |
| 4 | Zr(Ti)-NDC | 100 | 20 | 90 | 99 | 69 |
| 5 | Zr-NDC | 150 | 5 | 58 | 98 | 69 |
| 6 | Zr(Ti)-NDC | 150 | 5 | 82 | 98 | 71 |
| 7 | Zr-NDC-NH2 | 150 | 5 | 56 | 99 | 60 |
| 8 | Zr(Ti)-NDC-NH2 | 150 | 5 | 65 | 99 | 62 |
| 9 | Zr-NDC (untreated) | 150 | 5 | 45 | 99 | 74 |
| 10 | Zr(Ti)-NDC (untreated) | 150 | 5 | 57 | 99 | 76 |
| 11 | UiO-66-NH2 | 150 | 5 | 15d | 99 | 72 |
| 12 | Cu3(BTC)2 | 150 | 5 | 50d | 99 | 73 |
| 13 | MCM-41(Ti) | 150 | 5 | 90d | 99 | 70 |
| 14 | Sn-beta49 | 80 | 50 | >99 | >98 | 83 |
| 15 | Ti-beta49 | 80 | 50 | 35 | >98 | 56 |
| 16 | Cu3(BTC)2 (ref. 51) | 110 | 100 | 80 | — | 65 |
| 17 | UiO-66 (ref. 53) | 100 | — | 20 | — | 75 |
| 18 | UiO-66-NH2 (ref. 53) | 100 | — | 20 | — | 75 |
| 19 | MIL-101(Cr)55 | 80 | — | >99 | >99 | 74 |
| 20 | MOF-808-1.3SO4 (ref. 56) | 60 | — | 97 | — | 67 |
| 21 | Zr-TUD-1 (ref. 57) | 80 | — | 99.6 | 99.6 | 65 |
Zr(Ti)-NDC shows better performance than Zr-NDC at different temperatures and catalyst loading. For these heterogeneous catalysts, the diastereoselectivity for (±)-isopulegol decreased slightly at higher temperatures, although the overall selectivity to cyclization improved (Table 2). Zr-MOFs with NDC-NH2 linker are less active than NDC-derivatives. The cyclization of citronellal also proceeds smoothly without any solvent being the rate of reaction higher without solvent, which could be explained by a dilution effect. The diastereoselectivity for (±)-isopulegol was similar without solvent and in toluene, 63 and 66%, respectively. Under the same reaction conditions Cu3(BTC)2 shows a similar selectivity than Zr(Ti)-NDC catalysts. A comparison with data from literature for other MOF catalysts (Cu3(BTC)2, UiO-66, MIL-101(Cr), MOF-808-SO4) reveals that Zr(Ti)-NDC displays similar rates and selectivity (Table 2). A comparison with other solid molecular sieve inorganic catalyst such as MCM-41(Ti) and Sn-beta shows that the inorganic solids are more active and Sn-beta presents better diastereoselectivity (83%).
Finally, for preparative purposes, an experiment was scaled up by a factor of 10 and a conversion of 80% and diastereoselectivity of 60% were reached.
Cycle times of 24 h were chosen in order to ensure that a possible loss of activity of Zr(Ti)-NDC could be detected. Although the conversion decreased by about 15% in each consecutive batch experiment, the diastereoselectivity for (±)-isopulegol remained constant. The loss of activity might be partly attributed to successive treatments or blockage of pores. The catalyst was calcined at 300 °C (heating rate 1 °C min−1) for 2 h to remove any possible inhibitors. Activity of the Zr(Ti)-NDC was partially recovered up to 90–95%. Moreover, no leaching of zirconium or titanium could be detected (ICP-OES, detection limit 0.002 ppm). A hot filtration experiment was also performed and no reaction was observed in the filtrate upon removal of the catalyst after 8 h of reaction.
Comparison of X-ray diffraction patterns and SEM images of Zr-NDC and Zr(Ti)-NDC catalysts before and after several reaction runs did not reveal significant differences (Fig. S10†).
Isomerization of α-pinene oxide. The third test reaction for Lewis acid catalysis of Zr(Ti)-NDCs is the isomerization of α-pinene oxide, a highly sensitive substrate towards acids, which reacts to give a mixture of campholenic aldehyde (5), isopinocamphone (6), pinocarveol (7) and others products (Scheme 4). Campholenic aldehyde is a fragrance compound prepared in high yield when a suitable Lewis acid is used.58–61 It has been seen that MOFs as Cu3(BTC)2,51,52,62 UiO-66,63 MIL-100(Fe),64,65 Fe(BTC)62 with well-defined Lewis acid exhibited moderate activity and selectivity for this reaction.
The acid-catalyzed isomerization of α-pinene oxide was investigated using here Zr-NDC and Zr(Ti)-NDCs as catalysts and 1,2-dichloroethane (DCE) as solvent, at 70 °C (data in other solvents and temperatures were included in ESI, Table S3†). The added-value campholenic aldehyde product was formed in 50% yield at 80% conversion. The by-products included isopinocamphone, pinocarveol and traces of trans-carveol (Scheme 4). Without catalyst the reaction is slow and selective to campholenic aldehyde is poor, indicating that Zr(Ti)-NDCs promote the reaction. All reported yields and selectivities were reproducible for at least three experiments and were constant until complete conversion. The catalytic performance of Zr(Ti)s for the reaction is strongly dependent on the type of solvent used as shown in Table S3.†
Reactions were also carried out with other MOFs (Cu3(BTC)2, Fe-MIL-101) under the same conditions to compare the results with those obtained with Zr-NDC-MOFs. According to Table 3, the conversion and selectivity towards campholenic aldehyde depends on the type of Zr-MOF showing Zr(Ti)-NDC the best activity and selectivity (entry 2). It is well known that Lewis acid sites favor the formation of campholenic aldehyde; therefore, it is reasonable to suggest that the better catalytic properties of Zr(Ti)-catalysts are determined by titanium. Another aspect for the catalytic properties may be the difference in the structure of Zr(Ti)-MOFs, which can affect the reagents accessibility to active sites. This could explain the better activity and selectivity observed, in the presence of Zr(Ti)-NDC with pores of 2.06 nm (1.48 nm for Zr-NDC). Zr(Ti)-NDC give better results than MIL101 (Fe) (entry 10), MIL100 (Fe) (entries 15, 16) but better selectivity was found when Cu2(BTC)3 was the catalyst (entries 9, 14). A comparison with inorganic solids as Ti-beta shows that Zr(Ti)-NDC has a better catalytic activity but a lower selectivity towards campholenic aldehyde. MCM-41(Si/Al = 15) shows a similar behavior than Zr(Ti)-MOF.
| Entry | Catalyst | Conv. (%) (h) | Selectiv. (%) | ||
|---|---|---|---|---|---|
| 5 | 6 | 7 | |||
| a Reaction conditions: 5 mg cat, α-pinene oxide (0.13 mmol, 20 mg), 70 °C in 1,2-dichloroethane; yields and selectivity correspond to and overage value of three experiments.b This work.c Reactions were carried out at room temperature with 0.1 g of α-pinene oxide in 5 mL of solvent added to 0.1 g of Cu3(BTC)2.d Reaction conditions: 0.5 mL of α-pinene oxide, 50 mg catalyst activated at 150 °C for 2 h under vacuum before use, 70 °C without solvent.e Reaction conditions: 1,2-dichloroethane, 5 mg catalyst, 30 °C. | |||||
| 1 | Zr-NDC | 74 (24) | 51 | 19 | 30 |
| 2 | Zr(Ti)-NDC | 88 (24) | 58 | 20 | 22 |
| 3 | Zr-NDC-NH2 | 56 (24) | 43 | 28 | 29 |
| 4 | Zr(Ti)-NDC-NH2 | 55 (24) | 50 | 17 | 33 |
| 5 | UiO67-NH2 | 68 (24) | 44 | 26 | 30 |
| 6 | UiO67(Ti)-NH2 | 75 (24) | 50 | 22 | 28 |
| 7 | UiO66-NH2 | 46 (24) | 42 | 24 | 34 |
| 8 | UiO66(Ti)-NH2 | 74 (24) | 31 | 27 | 42 |
| 9 | Cu2(BTC)3b | 73 (24) | 82 | 3.5 | 7 |
| 10 | MIL101(Fe)b | 62 (24) | 48 | 10 | 34 |
| 11 | MCM-41(Si/Al = 15)b | 94 (21) | 60 | — | — |
| 12 | UiO66 (ref. 63) | 100 | 45 | — | — |
| 13 | Cu2(BTC)3 (ref. 62) | 8 (6) | 48 | 41 | — |
| 14 | Cu2(BTC)3 (ref. 51) | 70 (40)c | 80 | — | — |
| 15 | MIL100(Fe)62 | 22 (6)d | 45 | 40 | — |
| 16 | MIL100(Fe)64 | 96 (0.5)e | 56 | — | — |
| 17 | Ti-beta61 | 29 (24) | 81 | — | — |
Recycling experiments show that the activity of Zr(Ti)-NDC decreases about 30% upon recycling, probably due to absorption of subproducts on the active sites; regeneration by washing with ethanol or calcination of catalyst at 200 °C lead to a slightly recovering of catalytic activity (80% than original ones).
Footnote |
| † Electronic supplementary information (ESI) available: Experimental details on the preparation and characterization of materials, additional tables and figures are included. See DOI: 10.1039/c6ra23143h |
| This journal is © The Royal Society of Chemistry 2016 |