Next Article in Journal
Evaluation of the Antifungal, Antioxidant, and Anti-Diabetic Potential of the Essential Oil of Curcuma longa Leaves from the North-Western Himalayas by In Vitro and In Silico Analysis
Next Article in Special Issue
Proton Affinity in the Chemistry of Beta-Octamolybdate: HPLC-ICP-AES, NMR and Structural Studies
Previous Article in Journal
Thermogravimetric Analysis and Kinetic Modeling of the AAEM-Catalyzed Pyrolysis of Woody Biomass
Previous Article in Special Issue
Organic/Inorganic Species Synergistically Supported Unprecedented Vanadomolybdates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Grafting of Anionic Decahydro-Closo-Decaborate Clusters on Keggin and Dawson-Type Polyoxometalates: Syntheses, Studies in Solution, DFT Calculations and Electrochemical Properties

1
Institut Lavoisier de Versailles, CNRS, UVSQ, Université Paris-Saclay, 78035 Versailles, France
2
Laboratory of Organometallic and Coordination Chemistry, LCIO, Faculty of Sciences I, Lebanese University, Hadath 6573, Lebanon
3
Institute of Chemical Research of Catalonia (ICIQ), The Barcelona Institute of Science and Technology, 43007 Tarragona, Spain
4
Equipe de Recherche et Innovation en Electrochimie pour l’énergie (ERIEE), Institut de Chimie Moléculaire et des Matériaux d’Orsay (ICMMO), UMR CNRS 8182, Université Paris-Sud, Université Paris-Saclay, 91405 Orsay, France
5
Institut de Chimie Physique, CNRS, UMR 8000, Université Paris-Saclay, 91405 Orsay, France
6
Institut de Chimie des Substances Naturelles, CNRS UPR2301, Université Paris-Sud, Université Paris-Saclay, 91198 Gif-sur-Yvette, France
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(22), 7663; https://doi.org/10.3390/molecules27227663
Submission received: 2 October 2022 / Revised: 29 October 2022 / Accepted: 2 November 2022 / Published: 8 November 2022
(This article belongs to the Special Issue Research on Polyoxometalate Materials)

Abstract

:
Herein we report the synthesis of a new class of compounds associating Keggin and Dawson-type Polyoxometalates (POMs) with a derivative of the anionic decahydro-closo-decaborate cluster [B10H10]2− through aminopropylsilyl ligand (APTES) acting as both a linker and a spacer between the two negatively charged species. Three new adducts were isolated and fully characterized by various NMR techniques and MALDI-TOF mass spectrometry, notably revealing the isolation of an unprecedented monofunctionalized SiW10 derivative stabilized through intramolecular H-H dihydrogen contacts. DFT as well as electrochemical studies allowed studying the electronic effect of grafting the decaborate cluster on the POM moiety and its consequences on the hydrogen evolution reaction (HER) properties.

1. Introduction

Polyoxometalates (POMs) and POM-based materials constitute a highly versatile class of compounds rich in more than several thousand inorganic compounds, which can be finely tuned at the molecular level. Because of their stunning compositions, diversified architectures and their rich electrochemical redox behaviors, they are known to display numerous properties or applications in many domains such as supramolecular chemistry [1,2,3], catalysis [4,5,6], electro-catalysis [7,8,9], and medicine, especially when POMs are functionalized with organic groups or complexes [10,11,12,13].
On their side, hydroborates represent a wide family of anionic clusters, for which many reports demonstrated their interest in different areas, especially in the biomedical domain [14,15,16,17,18]. This property thus makes the studies of borane derivatives of a great interest. In particular, the [B10H10]2− cluster offers the possibility of various selective functionalizations [19,20] leading for example to closo-decaborate-triethoxysilane precursor, which can be coordinated to luminescent dye doped silica nanoparticles, hence facilitating the tracing of the closo-decaborate drug pathway in BNCT (Boron Neutron Capture Therapy) [21,22].
Driven by the synthetic challenge that constitutes the association of two anionic species with two complementary redox characters, reductive for hydroborates and oxidative for POMs, and by the biomedical applications which could be reached by associating these two families of compounds, this study aims to find the right strategy to design such POM-borate adducts and to study their chemical properties.
In a previous paper, we demonstrated that it is possible to covalently graft decaborate clusters to an Anderson-type polyoxometalate functionalized with the well-known TRIS ligand (TRIS = tris(hydroxymethyl)aminomethane), namely [MnIIIMo6O18(TRIS)2]3− [23]. Nevertheless, the compound [MnIIIMo6O18(TRIS-B10)2]7− resulting from the coupling between both components revealed to be fragile, probably because of the rigidity of the linker and the close proximity of both anionic components. This weakness is confirmed by DFT calculations indicating an athermic or slightly exothermic process for the formation of the adducts with Anderson-TRIS hybrid POMs.
In the field of hybrid POMs, the organosilyl derivatives of vacant polyoxotungstates as [PW9O34]9−, [SiW10O36]8−, [PW11O39]7−, [SiW11O39]6−, or [P2W17O61]10− offer large diversities of compounds exhibiting a wide panel of applications [24,25]. Among them, the divacant POM Keggin [SiW10O36]8− (noted hereafter SiW10) and the monovacant POM Dawson [P2W17O61]10− (noted hereafter P2W17) derivatives are probably the most used because of their stability, their topology and the richness of their electrochemical properties in reduction. In particular, by reacting with aminopropyltri(ethoxy)silane (called APTES) they can provide two very useful platforms, noted respectively SiW10-APTES and P2W17-APTES (see Figure 1), for elaborating functional hybrid molecular architectures.
The aim of this study is to use these two different platforms to prepare new hybrid compounds associating an anionic decaborate boron cluster (denoted hereafter B10) with Keggin and Dawson POM derivatives. The choice of polyoxotungstate moieties rather than Mo-based POMs is based on its stability towards reduction. The employment of a long and flexible linker as APTES is essential to tackle the challenge of combining efficiently a reduced anionic boron cluster with an anionic oxidized polyoxometalate. The use of APTES linker should limit the repulsion between the two components, while its flexibility allows more easily accommodating the two entities. Finally, as shown in Figure 1, due to monovacant and divacant characters of P2W17 and SiW10, respectively, it is worth noting that the relative conformations of the chains are different. For SiW10-APTES, the two alkyl chains are oriented nearly in parallel, whereas the monovacancy of P2W17 imposes divergent directions for the two alkyl chains. This topology is well adapted for designing triangular or square molecular species as evidenced by Izzet et al. [2,26], and in our case, we expect that these two kinds of conformation could lead to different types of adducts incorporating B10 clusters. In this study, we thus report the synthesis, the full characterization in solution by various NMR techniques, the electronic, the electrochemical and the electrocatalytic properties of three new hybrid POMs. In the absence of XRD structures, DFT studies provide a fine structural description of these hybrids and rationalization of their properties.

2. Results and Discussion

2.1. Syntheses

The synthesis of hybrid POMs can be achieved through different strategies. In the present study, the best synthetic procedure to get the targeted hybrid POMs has been to react first the lacunary POMs “SiW10” and “P2W17” with two aminopropyltri(ethoxy)silane molecules (APTES) to give the two POM-APTES precursors (see Figure 2) of formulas (TBA)3H[(SiW10O36)(Si(CH2)3NH2)2O]·3H2O (denoted hereafter SiW10-APTES) and (TBA)5H[P2W17O61(Si(CH2)3NH2)2O]·6H2O, denoted hereafter P2W17-APTES. The syntheses of these two precursors were adapted from Mayer at al. [28] by reaction of k8(γ-SiW10O36)·12H2O or K10α–P2W17O61·20H2O with 3-aminopropyltriethoxy silane in presence of TBABr in H2O/CH3CN medium acidified by concentrated HCl (for more details see experimental section in Supplementary Materials). Note that for each, the proton usually written as counter-cation is in fact probably an ammonium arm R-NH3+.
The synthetic strategy to get POM-borate adducts is then to combine the amines of these POM-APTES precursors with the reactive carbonyl of the decaborate cluster [B10H9CO] (Figure 1C) to give an amide function connecting both components. Since the boron cluster can react with water for giving a carboxylic acid and since heating the synthetic mixture above 40–50 °C led to some degradation products or to some reduction in the Dawson derivative by the hydrodecaborate cluster, reactions have been conducted at room temperature and under nitrogen atmosphere. Furthermore, the coupling reaction needs the presence of a base both to help the deprotonation of the ammonium arm(s) of the POM-APTES precursors and to trap the proton produced by the coupling reaction. A moderate and a bulky organic base, diisopropylethylamine (DIPEA), was thus used to avoid the competition with APTES for the coupling reaction with [B10H9CO].
To quickly circumscribe the optimal conditions for the synthesis of the POM-borate adducts, 29Si, 31P and 1H NMR titrations were conducted by varying the ratios of the three reactants [B10H9CO]/POM-APTES/DIPEA (all details are given in the Supplementary Materials).
For the [B10H9CO]/SiW10-APTES/DIPEA system, the 29Si NMR studies in solution reveal that it is possible to modulate the coupling reaction between [B10H9CO] and POM-APTES precursors by playing on the amounts of DIPEA and of [B10H9CO]. For this tri-reactants system, the successive formation of two POM-borate species identified as mono- and di-adduct compounds was demonstrated thanks to their molecular symmetries (Cs versus C2v). Besides, the crucial role of DIPEA in the reaction of [B10H9CO] with POM-APTES precursors was clearly evidenced. No reaction occurs when no base is used. NMR titration studies allowed establishing that the optimal quantity of base was two equivalents for one equivalent of [B10H9CO]. The Figure 2 shows for instance the proportions of SiW10-derivatives determined by the integration of the different peaks obtained by 29Si NMR in the system SiW10-APTES/[B10H9CO]/DIPEA as a function of [B10H9CO]/SiW10-APTES ratio at fixed DIPEA/[B10H9CO] ratio of 2.
Starting from SiW10-APTES, it evidences first the formation of a mono-adduct, which predominates for ration B10/SiW10-APTES = 1, before being converted into a di-adduct. The NMR titrations studies allowed establishing that using proportions SiW10-APTES/B10H9CO/DIPEA = 1/3/6 lead to the pure di-adduct denoted SiW10-diB10, while using 1/1/2 ratios lead to around 80% of mono-adduct mixed with some unreacted starting POM and the di-adduct. The separation of compounds has not been possible but considering the effect of the added DIPEA amounts, we succeeded to reduce the formation of the di-adduct and thus to get the mono-adduct compound denoted SiW10-monoB10 with a good purity by decreasing the quantity of DIPEA in the proportions SiW10-APTES/B10H9CO/DIPEA = 1/1/1.5. The Figure 3 summarizes the experimental conditions used to isolate POM-borate adducts.
Similar NMR studies were also performed in solution with the Dawson derivative P2W17-APTES (see Supplementary Materials). In contrast to SiW10 derivatives, the formation of mono- and di-adduct of the Dawson derivative are not so separated as for SiW10. Therefore, we failed to isolate the mono-adduct as pure product. Nevertheless, we can obtain quantitatively the di-adduct compound in the reaction mixture when ratios P2W17-APTES/B10H9CO/DIPEA = 1/3/6 are used.
To summarize, the multistep coupling reactions have successfully been monitored by 29Si and 31P NMR, fully described in the Supplementary Materials, revealing that intermediate products can be followed and isolated. From these results, we established the experimental conditions allowing to selectively synthesize with good yields the mono adduct of SiW10 POM and the di-adducts of both POMs as mixed TBA+ and DIPEAH+ salts, namely (TBA)3(DIPEAH)3[(SiW10O36)(B10H9CONHC3H6Si)(NH2C3H6Si)O]·3H2O denoted SiW10-monoB10, (TBA)6.5(DIPEAH)1.5[(SiW10O36)(B10H9CONHC3H6Si)2O]·2H2O denoted SiW10-diB10, and (TBA)6(DIPEAH)4[(P2W17O61)(B10H9CONHC3H6Si)2O]·3H2O, denoted P2W17-diB10 (See Experimental Section in Supplementary Materials for more details). All adducts were isolated as powders and were characterized by FT-IR, TGA, elemental analysis, MALDI-TOF and NMR techniques. It should be noted that to our knowledge, SiW10-monoB10 is the first example of a POM-APTES monoadduct isolated so far from the direct synthesis. All studies in the literature usually reported di-adducts with such types of hybrid POMs [29,30,31].

2.2. FT-IR Spectroscopy

FT-IR spectra are given in Figures S11 and S12 in Supplementary Materials. The FT-IR spectra of SiW10-monoB10, SiW10-diB10 and P2W17-diB10 evidence that the integrity of the POM part is maintained compared to the POM-APTES precursors. Furthermore, the association of the [B10H9CO] cluster is demonstrated by the disappearance of the carbonyl CO band at 2098 cm−1 in the B10H9CO cluster, while the broad band located at 2464–2470 cm−1 typical for B-H vibration bands of the decaborate moiety within the three compounds SiW10-monoB10, SiW10-diB10 and P2W17-diB10 is significantly shifted from that observed at 2517 cm−1 for the [B10H9CO] precursor [22,23].

2.3. Characterizations by MALDI-TOF Mass Spectrometry

Mass spectrometry (MS) is a very efficient technique for the characterization of polyoxometalates in solution. In our case, we did not succeed in getting mass spectra with reasonable signal-to-noise ratio and exploitable data by the usual electrospray ESI-MS technique. On the contrary, Matrix-Assisted Laser Desorption/Ionization coupled to a Time-of-Flight mass spectrometer (MALDI-TOF) revealed to be an effective technique for hybrid POMs characterization, as shown for example by Mayer and coworkers on “SiW10” and “P2W17“ organosilyl derivatives [28,32]. MALDI-TOF technique is applied on samples which are diluted in a matrix solution (DCTB in our case, DCTB = Trans-2-[3-(4-ter-Butylphenyl)-2-propenylidene] malonitrile) and then co-crystallized on a conductive target. Thanks to a laser irradiation, it allows producing singly charged species (cationic or anionic) and presenting the great advantage to strongly limit the number of peaks in comparison with ESI-MS spectra, where multiply charged species are generated. In the present study, the experiments were performed in both negative and positive modes (see Figure S14 in the Supplementary Materials for the example of SiW10-diB10). According to previous works in this field, the best results were obtained in the positive mode, although the anionic character of the POM [28,32]. Indeed, as seen in the Supplementary Materials for SiW10-diB10, the intensity reached in the negative mode appears lower, but the number of peaks is higher as there are more degradation species. Even thought our systems are polyanionic, they are more efficiently analyzed as monocationic species resulting from adducts between POMs and counter cations such as TBA+ and H+ in our case (H+ coming notably from DIPEAH+ cations or protonated amines). Furthermore, the monocationic character of the species is confirmed in all cases by the shift between peaks in the isotopic massifs.
The precursor SiW10-APTES and the compounds SiW10-monoB10, SiW10-diB10 and P2W17-diB10, were thus analyzed by this technique in the positive mode. The results are gathered in Table S1 (see Supplementary Materials). The full spectra and a zoom on the target compounds with a spectrum simulated with IsoPro3 software are shown in Figure 4 for SiW10 derivatives and in Figures S15 and S16 for P2W17 ones.
As shown in Figure 4 the spectrum of the precursor SiW10-APTES (Figure 4a) displays a major peak centered at m/z 4087.3 and a minor peak at m/z 4328.3. The first peak is assigned to the monocationic species {(TBA)3H2[(SiW10O36)O(SiC3H6NH2)2](CH3CN)2(H2O)8(DCTB)2}+ (calculated m/z 4087.3), while the second peak is attributed to the species {(TBA)4H[(SiW10O36)O(SiC3H6NH2)2](CH3CN)2(H2O)8(DCTB)2}+ (calculated m/z 4328.7). The two peaks correspond to the expected hybrid POM associated with some TBA+ and H+ cations, some solvates and two molecules of the DCTB matrix. Note that the presence of amines on the APTES part of the POM could probably favor the formation of intermolecular interactions with solvates and DCTB molecules. Such an adduct with DCTB is also observed with the precursor P2W17-APTES (Figure S15, Supplementary Materials) but not seen with the other POMs functionalized with B10 clusters. The MALDI-TOF spectrum of P2W17-APTES indeed exhibits a major peak corresponding to the expected precursor associated with one molecule of the DCTB matrix at m/z 6058.7 (calculated m/z 6057.9 for (TBA)6H[(P2W17O61)O(SiC3H6NH2)2](DCTB)}+) and a minor peak at m/z 6300.0 (calculated m/z 6299.4 for (TBA)7[(P2W17O61)O(SiC3H6NH2)2](DCTB)}+). The attribution of the peaks is definitely confirmed thanks to the fitting of the isotopic distribution massifs. The latter are mainly due to the isotopic distribution of the 10 or 17 tungsten atoms of the POMs, which appears consistent with the experimental spectrum (see Figure 4a and Figure S15, respectively).
The spectrum of SiW10-monoB10 depicted in Figure 4b shows only one experimental peak at m/z 3843.2 which is perfectly consistent with the calculated mass for the monocationic product {(TBA)4H3[(SiW10O36)O(SiC3H6NH2)(SiC3H6NHCOB10H9)](CH3CN)(H2O)3}+ (calculated m/z 3843.1). It evidences the formation of the expected adduct SiW10-monoB10 and thus indirectly the grafting of one (B10H9CO) cluster to SiW10-APTES. The simulated spectrum agrees well with the experimental data, which supports this assumption although the presence of one B10 cluster does not modify significantly the isotopic massif.
The MALDI-TOF spectrum of SiW10-diB10 shown in Figure 4c displays a major peak centered at m/z 4085.8, which fully agrees with the expected di-grafted compound {(TBA)4H5[(SiW10O36)O(SiC3H6NHCOB10H9)2](CH3CN)2(H2O)6}+ (m/z calculated 4084.4) and a minor peak at m/z = 4330.2 consistent with the species {(TBA)5H4[(SiW10O36)O(SiC3H6NHCOB10H9)2](CH3CN)3(H2O)4}+ (m/z calculated 4330.8). This result confirms the formation of the expected di-grafted compound.
Finally, the case of P2W17-diB10, appears more complicated, certainly due to a higher charge of the hybrid POM (10-) and a larger surface, which both favor intermolecular interactions with solvent molecules and cations. For technical reasons, the MALDI-TOF spectrum shown in Figure S16 (see Supplementary Materials) was recorded in linear mode, which does not favor the high resolution in contrast with other compounds. The spectrum displays an intense and broad experimental peak centered at m/z 6048.1, while four smaller peaks are found, respectively, at m/z 6289.6, 6431.7, 6672.5 and 6813.8. All these peaks are consistent with di-grafted species of general formula {(TBA)xHy[(P2W17O61)O(SiC3H6NHCOB10H9)2](CH3CN)z(H2O)t}+ (x + y = 11, z = 0–3 and t = 5–6). Regarding the main peak, the latter appears much broader than expected for only one species. Moreover, the resolution of the isotopic massif is lost. In fact, the experimental spectrum likely corresponds to a spectra superimposition of monocationic species of general formula {(TBA)5H6[(P2W17O61)O(SiC3H6NHCOB10H9)2](CH3CN)x(H2O)y}+ with x ranging from 1 to 5 and y from 0 to 8 (m/z in the range 6035.70 to 6076.75). Some simulated spectra are given in Figure S16 in Supplementary Materials.

2.4. NMR Studies in Solution

In the absence of crystallographic data, the three obtained hybrid systems have been thoroughly characterized by multinuclear NMR spectroscopy in order to verify their structures in solution. 1H, 11B, 13C, 15N, 29Si, 31P, and 183W NMR spectra were recorded in CD3CN at room temperature. The data are gathered in Table S2 (Supplementary Materials), while selected spectra are given in Figure 5 and Figure 6 and in Figures S17–S32 (Supplementary Materials).
As shown in Figure 5a and in Figures S17–S20 (Supplementary Materials), 11B{1H} NMR spectrum of [B10H9CO] undergoes a significant change upon coupling with SiW10-APTES or P2W17-APTES. In particular, the signal at −44.4 ppm specific for the equatorial boron atom bearing the substituent CO in [B10H9CO] (B2 atom, see Figure 1c) is strongly shifted to ca. −25 ppm in the spectra of SiW10-monoB10, SiW10-diB10 and P2W17-APTES in agreement with the grafting of the cluster on the POM.
Concomitantly, the 1H NMR spectrum of the mono adduct SiW10-monoB10, exhibits a splitting of the signals for the three methylene groups –CH2- of the APTES linker, denoted a, b, c (see Figure 5b), because of the lowering of the symmetry of SiW10-APTES from C2v to Cs. In addition, a new peak at 6.12 ppm assigned to an amide function is observed. For the remaining amine function, a broad signal is observed at 7.4 ppm (d), but together with two other broad signals at 5.70 and 6.33 ppm (d’ and d’’), attributed to the amine function in a frozen configuration in which the interaction with B10 cluster generates two inequivalent protons as depicted in Figure 7a (DFT optimized structure). These assumptions are confirmed by 1H-15N HMBC (Heteronuclear Multiple Bond Correlation) NMR spectrum (Figure S25) revealing two 15N signals at −251 ppm (amide) correlated to the proton signal at 6.12 ppm and at −272 ppm (free amine) correlated to the two protons peaks at 5.70 and 6.33 ppm.
Grafting a second [B10H9CO] group on the SiW10-APTES platform allows recovering the C2v symmetry and thus one set of peaks was observed for the linker and especially the protons “a”, in addition to an amide peak at 5.94 ppm (Figure S24, Supplementary Materials). The signals of the amine at 7.4, 5.7, and 6.3 ppm disappear in agreement with the reaction of [B10H9CO] groups with this function. Similarly, the 1H NMR spectrum of P2W17-diB10 compared to that of P2W17-APTES (Figure S27, Supplementary Materials) evidences the appearance of a sharp peak at 5.94 ppm assigned to an amide function, while the signal of the free amine at 7.03 ppm in the precursor P2W17-APTES disappears.
To further confirm our assignments of the signal of the amide function, 1H-1H ROESY (Rotating frame Overhause Effect SpectroscopY) and 13C NMR experiments were performed on SiW10-diB10 and P2W17-diB10 (Figures S28–S32, Supplementary Materials). Cross REO peaks involving the amide proton (5.94 ppm in both compounds) and some equatorial B-H protons of the B10 cluster at 0.4 ppm and the protons “c” of the APTES chains can be seen in Figures S28 and S29 (Supplementary Materials). This demonstrates the spatial proximity between these protons that interact between each other through dipolar contacts. 13C NMR spectra of SiW10-monoB10, SiW10-diB10, and P2W17-diB10 (Figures S30–S32) notably exhibits a signal at 203 ppm assigned to a carbon atom from an amide group, which is confirmed by 2D 1H-13C HMBC NMR spectrum of SiW10-monoB10 evidencing a correlation between this 13C signal at 203 ppm and the 1H amide signal at 6.12 ppm. Besides, in both cases of SiW10-monoB10 and SiW10-diB10 this 13C signal appears as a poorly resolved quadruplet with a coupling constant of 95 Hz consistent with a 1J13C-11B coupling.
Therefore, 1H, 1H-15N HMBC, 1H-1H ROESY, 13C and 1H-13C HMBC NMR experiments unambiguously confirm the formation of an amide group in our three adducts by reaction of the amines of POM-APTES precursors with the carbonyl group of the cluster [B10H9CO]. The modification of the 11B{1H} NMR spectra of the boron cluster after its reaction with the POM-APTES precursors further confirms such results.
29Si, 31P and 183W NMR probe the POM part in compounds SiW10-APTES, SiW10-monoB10, SiW10-diB10, P2W17-APTES and P2W17-diB10 (see Figure 6 and Figures S21–S23 in Supplementary Materials). The unsymmetrical environment in the mono adduct SiW10-monoB10 is clearly confirmed through the appearance of two peaks for Si of the different linker arms at −61.9 and −63.3 ppm, while only one signal was observed at −62.3 ppm for the symmetrical di adduct SiW10-diB10 with only a small shift from the initial SiW10-APTES precursor (−62.5 ppm). In addition, for all the compounds, a single peak is observed for the Si atom in the central cavity of the SiW10 POM moiety which is almost not affected by the grafting of the boron clusters and the resulting changes of symmetry of the adducts (Figure S21, Supplementary Materials). In case of P2W17-APTES and P2W17-diB10, both compounds exhibit only one signal assigned to the two equivalent Si atoms of the APTES linker (Figure S22, Supplementary Materials).
The 183W NMR spectra of precursors and adducts are given in Figure 6. For SiW10-monoB10 the 183W NMR spectrum displays five peaks of intensities 2:2:2:2:2 in agreement with the expected low Cs symmetry, while three resonances are observed for the di adduct SiW10-diB10 and the initial precursor SiW10-APTES of intensities (4:2:4) consistent with their C2v symmetry (Figure 6a). Figure 6b shows the 183W NMR spectrum of P2W17-diB10 which differs significantly from its precursor. Both compounds exhibit nine NMR lines of integration 2:2:2:1:2:2:2:2:2 in agreement with the expected Cs symmetry, but their positions are slightly changed. This is due to the modification of the P2W17 moiety induced by the grafting of the two [B10H9CO] clusters. Additionally, 31P NMR spectra of P2W17-APTES and P2W17-diB10 display two signals (Figure S23, Supplementary Materials), wherein one of them showed a common chemical shift, while the second exhibited a small shift from −13.4 ppm in P2W17-APTES to −13.6 ppm in P2W17-diB10.
In conclusion, these experiments focused on the POM part fully agree in terms of molecular symmetries with the formation of the expected mono- or di-adducts with B10 clusters.

2.5. Computational Studies

The molecular geometries of SiW10-APTES, SiW10-monoB10, and SiW10-diB10, as well as those of P2W17-APTES, P2W17-monoB10 and P2W17-diB10 were fully optimized at a DFT level including implicit solvent effects (see Figure 7 and Supplementary Materials for computational details). We considered the most relevant plausible conformers. Firstly, regarding SiW10-APTES, it exhibits two main conformers as defined by the orientation of the two amine organic arms, which we called them open and closed forms. The small difference in their relative energy, less than 1 kcal·mol−1 in favor of the closed form (represented in Figure 1a), forecasted that further substitution would easily overcome any initial geometric preference in the reactants. Indeed, upon B10 incorporation a much more complex situation arises. For SiW10-monoB10 we characterized five conformers, two arising from the closed reactant and three species from the open reactant form. In the most stable conformer (Figure 7a), which arise from the closed form, the decaborate moiety interacts favorably with the amine hydrogens (d’ and d’’) of the unreacted arm through strong dihydrogen contacts. In the most stable open form (Figure 7b), although interaction between arms is almost neglected, the H amide atom develops other interactions. Overall, the most stable conformer given in Figure 7a is 11 kcal·mol−1 below the second one (Figure 7b). All five conformers lie in a narrow 20 kcal·mol−1 range. For the double substituted SiW10-diB10, since the additional repulsion arising from the negatively charged B10 groups, we could only characterize two forms: an open (not shown) and a closed one (two views on Figure 7c,d). The energy difference between both species was computed to be just only 5 kcal·mol−1. We highlight, as dashed lines in Figure 7, those hydrogen atoms of the organic arm and the B10 moiety that lie close in three-dimensional space.
Due to the monovacant character of P2W17, the Si-O-Si angle of the APTES moiety grafted to the POM strongly differs from that observed for the divacant SiW10 POM (see Figure 1). The topology of the two arms, and thus the connectivity of the two POM-APTES derivatives, strongly differs. Therefore, as seen in Figure 1, closed form is not possible for P2W17-APTES. Only one geometry could thus be considered. Then, for the mono-substituted P2W17-monoB10, two conformers were characterized, one open and one folded, the folded one being more stable by only 2.2 kcal·mol−1. For the di-substituted Dawson derivative, only one open shaped product could be characterized (see Figure 7e,f).
Hydrogen atoms of the decaborate moieties possess a hydride character. Consequently, they can establish hydrogen-hydrogen contacts with protic solvent or with functional groups like amines or amides. In the present structures, many H-H dihydrogen contacts between the amine organic arms from APTES moiety and hydrogen atoms from decaborate clusters were observed. For instance, for SiW10-monoB10 (Figure 7a), the hydrogen atoms d’ and d’’ from the «free» amino group are found 2.21 and 1.99 Å far from an H atom belonging to the B10 cluster, which are quite short distances. This fact agrees with the couplings observed in the NMR experiments.
The thermodynamics of the formation of the mono- and di-adducts of the Keggin and the Dawson species is computed exergonic in all cases as seen in Figure 8. For SiW10-APTES, both the formation of the mono- and bi-derivative were computed exergonic, 22.1 kcal·mol−1 for the SiW10-monoB10, and 13.0 kcal·mol−1 for the SiW10-diB10. The formation of SiW10-monoB10 is clearly favored and the strong dihydrogen contacts between amino group and the grafted B10 cluster undoubtedly strongly stabilize such a species compared to the di-adduct SiW10-diB10. For P2W17-APTES, also the two substitutions are favorable, 6.8 kcal·mol−1 for the first, and 7.9 kcal·mol−1 for the second.
The computed reaction free energies for the SiW10-APTES and P2W17-APTES derivatives are fully consistent with the experimental findings. For the keggin derivatives, the strong stabilization of the mono-adduct allows isolating both mono and di-adduct thanks to the formation of strong H-H contacts. In contrast, for the Dawson derivatives, the small difference of energies between mono- and bi-adducts (1.1 eV only) does not permit isolating the mono-adduct. Besides, an excess of [B10H9CO] (3 equivalents/POM instead of 2) is needed to get the pure di-adduct compound to avoid the formation of a mixture between mono and di-grafted adducts. A similar situation was previously obtained with Anderson-type derivatives since the mono and di-adduct of B10 with [MnMo6(Tris)2]3− are only separated by 6 eV and it was not possible to get mono-adduct [22]. This result highlights the role of the topology of the POM-APTES compounds and their faculty to stabilize species thanks to intramolecular interactions.
Finally, DFT studies provided the frontier orbitals for each compound (see Figure 9 and Figure S33 in Supplementary Materials). The results obtained for Keggin and Dawson derivatives exhibit the same feature. For POM-APTES the HOMO is located on one (for SiW10-APTES) or two (P2W17-APTES) amines of the APTES part, while the LUMO are localized on the W atoms of the POM part. By grafting the B10 clusters, the LUMO levels are slightly affected. LUMO remains localized on W atoms and only minor changes in energy are observed.
Conversely, the HOMO levels are drastically modified by the introduction of B10 clusters. Electrons of the HOMO orbitals are now mainly localized on one grafted B10 cluster. Interestingly, for SiW10-monoB10 the HOMO is delocalized between one B10 cluster and the amine of the second arm, which strongly interacts with the B10 through H-H contacts. The HOMO energy level increases in all cases within the range 0.78 to 0.92 eV. The HOMO-LUMO gaps, are thus significantly reduced upon the B10 grafting. Indeed, for SiW10-APTES, the gap decreases from 1.67 to 0.90 eV for the first substitution, and to 0.94 eV for the second. For P2W17-APTES, the gap evolves from 1.56 to 0.77 eV for the first substitution, and to 0.76 eV for the second.

2.6. Electrochemical Properties

The electronic spectra of compounds SiW10-monoB10, SiW10-diB10 and P2W17-diB10 recorded in CH3CN containing 0.1 mol.L−1 TBAClO4 (TBAP, tetrabutylammonium perchlorate) at room temperature and 2.10−4 mol.L−1 concentration are depicted in Figures S35 and S36 (Supplementary Materials).
The electronic spectra of precursors SiW10-APTES and P2W17-APTES display absorption bands in ultraviolet region corresponding to transition between p-orbitals of the oxo ligands and d-type orbitals centered on tungsten [33,34], while the cluster [B10H9CO] exhibits weak absorption band between 300 and 200 nm notably assigned to π−π* transitions [35]. Considering that the main contribution of the spectra comes from the LMCT band involving the W atoms and that the LUMO band centered on tungsten atoms are only slightly modified upon grafting of B10 cluster, no drastic changes are expected in the POM-B10 adducts. Indeed, the electronic spectra of SiW10-monoB10 and P2W17-diB10 match well with the sum of the spectra of SiW10-APTES or P2W17-APTES and one or two times that of [B10H9CO], respectively. The spectrum of SiW10-diB10 slightly differs from the sum of the component’s spectra probably due to a larger variation of LUMO energy level from SiW10-APTES to SiW10-diB10 and additional constraints due to the vicinity of the two boron clusters.
Since no evolution of the spectra were observed within 24 h in such a medium, the compounds appear chemically stable in these experimental conditions. The cyclic voltammograms (CVs) were thus recorded for all the SiW10 and P2W17 derivatives and are given in Figure 10 and in Figures S37 and S38 (Supplementary Materials), while the anodic and cathodic potentials are gathered in Table S3 (Supplementary Materials).
As depicted in Figure 10a and Figure S37, the CV of SiW10-APTES is poorly resolved and it is difficult to identify confidently all the reduction processes corresponding to the successive reduction in WVI centers into WV, well-known to be monoelectronic in non-aqueous solvents [36]. These waves seem nevertheless reversible with processes better resolved in oxidation. Besides, an irreversible process attributed to the oxidation of amine function of the APTES linker is also observed in oxidation around +0.452 V vs. Fc+/Fc (see Supplementary Materials). As for their parent precursor, CVs of SiW10-monoB10 and SiW10-diB10 display poorly resolved reversible electronic transfers, which appear shifted towards more negative potentials and one irreversible oxidation process around +0.452 V vs. Fc+/Fc assigned to the oxidation of the remaining amine function and/or of the B10 cluster (see Figure S39, Supplementary Materials). This behavior agrees with the increase in the charge from 4- in SiW10-APTES to 6- in SiW10-monoB10 and 8- in SiW10-diB10 and the electron donating character of the boron cluster [37] but does not evidence a strong electronic effect of the boron cluster on the POMs electronic properties.
The CVs of P2W17-APTES and P2W17-diB10 are given in Figure 10b and in Figure S38 (Supplementary Materials) and appears much more resolved than those of SiW10 derivatives. The CV of P2W17-APTES displays four reversible electronic transfers with cathodic peak potentials assigned to successive mono- or bi-electronic reductions of WVI centers into W(+V) [38] and two irreversible oxidation processes at Epa = +0.452 and +0.759 V, assigned to the oxidation of the terminal amine groups of the APTES linkers. Conversely to the di-adduct compound SiW10-diB10, the CVs of the Dawson derivative P2W17-diB10 give four reversible reduction processes significantly shifted towards the more positive potential compared to P2W17-APTES and one irreversible reduction process at Epc = −2.030 V vs. Fc+/Fc, which was not observed in the precursor. The opposite effect was expected. This effect probably results from a combination of a charge effect, the presence of protons (in DIPEAH+ cations) and of an electronic effect of boron cluster on P2W17 moiety but at this stage it is difficult to have a clear explanation of the contribution of all these effects which can be antagonist.
Although reduction waves in the Dawson derivatives are not very well resolved, it can be observed that the di-substituted species (green line in Figure 10) is reduced at lower potentials than P2W17-APTES, in agreement with the fact that the LUMO and LUMO+1 raise in energy upon B10 attachment. Also, the successive reduction waves seem just shifted left, which would conform with the almost constant difference in the LUMO and LUMO+1 energies along the series.

2.7. Electrocatalytic Properties for the Reduction in Protons into Hydrogen (HER)

Many POMs are known to catalyze protons reduction into hydrogen in aqueous or in non-aqueous conditions [7,39,40]. We verified by UV-Vis spectroscopy that [B10H9CO] and its adducts with POMs are stable in CH3CN in the presence of excess acetic acid (20 equivalents). In these conditions, it was interesting to study the reactivity of these compounds in regard to the electro-catalytic reduction in protons into hydrogen. The experiments were performed in CH3CN + 0.1 M TBAP by using acetic acid as a source of protons, and as a weak acid in such a medium (pKa = 22.3) [41].
Figure 11 and Figure S40 (Supplementary Materials) show the evolution of CVs upon stepwise addition of acetic acid up to 20 equivalents of acid/POM for all P2W17 and SiW10 derivatives, respectively. For all the compounds, the addition of acetic acid, gives a new irreversible reduction wave, which grows gradually with the amount of acid, expressed as γ = [acid]/[POM]. As shown in Figure 11 and Figure S40, at a given potential of −2.2 V vs Fc+/Fc, a linear dependence of the catalytic current versus γ is obtained, a behavior featuring the electro-catalytic reduction of protons. However, the effect of the addition of acetic acid in the solution appears stronger for P2W17 derivatives than for SiW10 based compounds.
To evidence the electrocatalytic process, linear voltammetry of P2W17-diB10 in the presence of 20 equivalents of acetic acid was performed and compared to similar experiments performed without catalyst or on platinum electrode (Figure 11d). We notice that in the presence of the catalyst P2W17-diB10, the current density is almost doubled and there is a 250-mV overvoltage decrease compared to the solution without catalyst. Indeed, the proton reduction with respect to platinum starts at −1.400 V vs. Fc+/Fc, while it starts at −1.750 V vs. Fc+/Fc with catalyst and at −2.000 V vs. Fc+/Fc without catalyst. Finally, the formation of hydrogen is unambiguously demonstrated by gas chromatography analysis during electrolysis performed at −2.200 V vs Fc+/Fc during 4.5 h (Figure S40, Supplementary Materials).
To compare the efficiency of all compounds, the catalytic efficiency (CAT) can be estimated using Equation (1):
C A T = 100 J P O M + 20   e q .   C H 3 C O O H J P O M   a l o n e J P O M   a l o n e  
Table 1 summarizes the CAT values measured for our products at −2.2 V vs Fc+/Fc. Interestingly, the two precursors SiW10-APTES and P2W17-APTES exhibit similar efficiency. Also, the efficiency of SiW10-monoB10 and SiW10-diB10 adducts are lower than that of SiW10-APTES, while it is the opposite for P2W17-diB10, which appears much more efficient than its parent precursor, in agreement with cyclic voltammetry experiments. Indeed, a less negative reduction potential of the POM part should facilitate the electro-catalytic reduction in protons.
In terms of mechanism, three key steps have to be considered: the protonation, the reduction in the catalyst and the transfer of electron towards the protons to give dihydrogen. For protonation step, since catalysis is observed in all compounds, it must occur on the most basic sites, either on the oxo groups of the POM moiety, on the free amine groups in SiW10-APTES and P2W17-APTES or on boron clusters for SiW10-diB10 and P2W17-diB10. DFT calculations evidence that the most nucleophilic sites are found on the oxo ligands of the POM parts which are consequently the preferential sites for protonation (see Figure 12 and Figure S34 in Supplementary Materials).
For the reduction step, as seen in Figure S41c,d in Supplementary Materials, during electrolysis, the P2W17 derivatives turned to blue as expected for the reduction in such species before returning back colorless when the current is stopped indicating that the reduced POM probably transfers electrons to protons to produce hydrogen. We understand well that if this reduction occurs at higher potential, it should favor the process. P2W17-diB10 is thus logically the most efficient compound.
To sum up, even if the decaborate cluster is probably not directly involved in the HER process, it plays two indirect roles: (1) the covalent grafting on POMs increases the electronic density on the POM which should facilitate the protonation step, and (2) the covalent grafting can modifies the reduction potential of the POM moieties in POM-borate adducts, which favors the reduction step of the POM species when shifted towards more positive potentials as observed in P2W17-diB10.

3. Conclusions

In this work, we succeeded in combining covalently the anionic [B10H9CO] cluster with anionic SiW10 and P2W17 derivatives using functionalized silyl derivatives (APTES) as a linker. The coupling between the two families of anionic parts appears much stronger than that with Anderson-type POMs we previously reported [23] and detailed NMR study allowed establishing the optimized conditions for the synthesis of target compounds. Hence, the selective isolation of mono- and di-adduct compounds of boron cluster with SiW10-APTES, namely [(SiW10O36)(B10H9CONHC3H6Si)(NH2C3H6Si)O]6− and [(SiW10O36)(B10H9CONHC3H6Si)2O]8− was successfully achieved, while only the di-adduct [(P2W17O61)(B10H9CONHC3H6Si)2O]10− was isolated with P2W17-APTES. To the best of our knowledge, it is the first time that a mono-adduct can be isolated directly from the synthesis by functionalization of the SiW10-APTES precursor. DFT studies supported by experimental NMR data evidenced that the formation of intramolecular H-H dihydrogen contact is the driving force for the preferred formation of the mono-adduct species and such a synthetic strategy could open the route toward the formation of hybrid POMs with two different functional groups.
All these compounds were fully characterized by multi-NMR techniques including 1H, 11B, 13C, 15N, 29Si, 183W and 31P as well as multi-dimensional correlations such as COSY, HMBC (1H-13C and 1H-15N) and ROESY NMR allowing focusing on each part of the adducts, i.e., POM, linker and boron cluster. These characterizations demonstrated unambiguously the formation of the targeted adducts and were also consistent with FT-IR and MALDI-TOF spectrometry data. DFT studies permitted to get optimized structures for all compounds consistent with the NMR data.
The electrochemical studies allowed studying the electronic effects of the grafting of the reducing boron cluster on some oxidized POMs with probable antagonist effect between charge effect and the variation of frontiers orbitals levels upon grafting of B10 cluster. Finally, electro-catalytic reduction in protons into hydrogen was evidenced for these systems, the best efficiency being obtained with P2W17-diB10. The process appears mainly effective on the POM part while the boron cluster participates only indirectly to the process.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27227663/s1. Experimental section including general methods, the 29Si, 31P and 1H NMR titration studies (Figures S1–S10) and the syntheses of compounds used in this study; FT-IR spectra (Figures S11 and S12); MALDI-TOF data (Table S1) and spectra (Figures S13–S16); a summary of NMR data (Table S2), 11B NMR studies (Figures S17–S20), 29Si NMR spectra (Figures S21 and S22), 31P NMR spectra (Figure S23), 1H NMR studies including COSY, ROESY and 1H-15N HMBC experiments (Figures S24–S29); 13C NMR studies (Figures S30–S32); a DFT part including computational details, frontiers orbitals of P2W17 derivatives (Figure S33) and electron density maps on SiW10 derivatives (Figure S34); electronic spectra (Figures S35 and S36), additional cyclovoltammograms (Figures S37–S39) and table of potentials (Table S3); Evidence of the production of hydrogen and picture of the reduced POM during the HER process (Figures S40 and S41).

Author Contributions

Synthesis and solution studies, M.D. and Z.E.H.; elemental analyses, N.L.; supervision of NMR studies, M.H.; MALDI-TOF experiments, V.G. and D.T.; DFT Calculation, A.M. and C.B.; electrochemical studies, J.E.C. and A.R.; supervision of the work and funding acquisition, E.C., D.N. and S.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a public grant overseen by the French National Research Agency as part of the «Investissements d’Avenir» program (Labex Charm3at, ANR-11-LABX-0039-grat), by the Paris Ile-de-France Region-DIM “Respore”, and by a mobility program PHC CEDRE (project POMBORON n°42237UG). M.D. thanks AZM association and IUF for financial support and Z.E.H. thanks CampusFrance for Eiffel doctoral grant. We also thank CERCA Program of the Generalitat de Catalunya, the ICIQ Foundation, and the Spanish Ministerio de Ciencia e Innovación through project PID2020-112806RB-I00 and through the Severo Ochoa Excellence Accreditation 2020–2023 (CEX2019-000925-S, MCI/AEI).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

A data set collection of computational results is available in the ioChem-BD repository and can be accessed via https://dx.doi.org/10.19061/iochem-bd-1-217 (accessed on 1 November 2022).

Acknowledgments

We acknowledge the Centre National de la Recherche Scientifique (CNRS) and the Ministère de l’Education Nationale de l’Enseignement Supérieur, de la Recherche et de l‘Innovation (MESRI) for their financial support. This study results from an international collaboration supported by IRN-CNRS 2019–2023.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lin, C.G.; Fura, G.D.; Long, Y.; Xuan, W.M.; Song, Y.F. Polyoxometalate-based supramolecular hydrogels constructed through host–guest interactions. Inorg. Chem. Front. 2017, 4, 789–794. [Google Scholar] [CrossRef]
  2. Izzet, G.; Abecassis, B.; Brouri, D.; Piot, M.; Matt, B.; Serapian, S.A.; Bo, C.; Proust, A. Hierarchical Self-Assembly of Polyoxometalate-Based Hybrids Driven by Metal Coordination and Electrostatic Interactions: From Discrete Supramolecular Species to Dense Monodisperse Nanoparticles. J. Am. Chem. Soc. 2016, 138, 5093–5099. [Google Scholar] [CrossRef] [PubMed]
  3. Zhu, Y.; Yin, P.C.; Xiao, F.P.; Li, D.; Bitterlich, E.; Xiao, Z.C.; Zhang, J.; Hao, J.T.; Liu, B.; Wang, Y.; et al. Bottom-Up Construction of POM-Based Macrostructures: Coordination Assembled Paddle-Wheel Macroclusters and Their Vesicle-like Supramolecular Aggregation in Solution. J. Am. Chem. Soc. 2013, 135, 17155–17160. [Google Scholar] [CrossRef] [PubMed]
  4. Vickers, J.W.; Lv, H.J.; Sumliner, J.M.; Zhu, G.B.; Luo, Z.; Musaev, D.G.; Geletii, Y.V.; Hill, C.L. Differentiating Homogeneous and Heterogeneous Water Oxidation Catalysis: Confirmation that [Co4(H2O)2(α-PW9O34)2]10– Is a Molecular Water Oxidation Catalyst. J. Am. Chem. Soc. 2013, 135, 14110–14118. [Google Scholar] [CrossRef] [PubMed]
  5. Sarma, B.B.; Neumann, R. Polyoxometalate-mediated electron transfer–oxygen transfer oxidation of cellulose and hemicellulose to synthesis gas. Nat. Commun. 2014, 5, 4621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Natali, M.; Orlandi, M.; Berardi, S.; Campagna, S.; Bonchio, M.; Sartorel, A.; Scandola, F. Photoinduced Water Oxidation by a Tetraruthenium Polyoxometalate Catalyst: Ion-pairing and Primary Processes with Ru(bpy)32+ Photosensitizer. Inorg. Chem. 2012, 51, 7324–7331. [Google Scholar] [CrossRef]
  7. Keita, B.; Nadjo, L. Polyoxometalate-based homogeneous catalysis of electrode reactions: Recent achievements. J. Mol. Catal. A Chem. 2007, 262, 190–215. [Google Scholar] [CrossRef]
  8. Freire, C.; Fernandes, D.M.; Nunes, M.; Abdelkader, V.K. POM & MOF-based Electrocatalysts for Energy-related Reactions. ChemCatChem 2018, 10, 1703–1730. [Google Scholar]
  9. Zhao, J.-W.; Li, Y.-Z.; Chen, L.-J.; Yang, G.-Y. Research progress on polyoxometalate-based transition-metal–rare-earth heterometallic derived materials: Synthetic strategies, structural overview and functional applications. Chem. Commun. 2016, 52, 4418–4445. [Google Scholar] [CrossRef]
  10. Hasenknopf, B. Polyoxometalates: Introduction to a class of inorganic compounds and their biomedical applications. Front. Biosci. 2005, 10, 275–287. [Google Scholar] [CrossRef] [Green Version]
  11. Yamase, T. Anti-tumor, -viral, and -bacterial activities of polyoxometalates for realizing an inorganic drug. J. Mater. Chem. 2005, 15, 4773–4782. [Google Scholar] [CrossRef]
  12. Bijelic, A.; Aureliano, M.; Rompel, A. The antibacterial activity of polyoxometalates: Structures, antibiotic effects and future perspectives. Chem. Commun. 2018, 54, 1153–1169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Bijelic, A.; Aureliano, M.; Rompel, A. Polyoxometalates as Potential Next-Generation Metallodrugs in the Combat Against Cancer. Angew. Chem. Int. Ed. 2019, 58, 2980–2999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Sivaev, I.B.; Bregadze, V.I.; Kuznetsov, N.T. Derivatives of the closo-dodecaborate anion and their application in medicine. Russ. Chem. Bull. 2002, 51, 1362–1374. [Google Scholar] [CrossRef]
  15. Zhu, Y.; Hosmane, N.S. Nanostructured boron compounds for cancer therapy. Pure Appl. Chem. 2018, 90, 653–663. [Google Scholar] [CrossRef] [Green Version]
  16. Hey-Hawkins, E.; Viñas-Teixidor, C. Boron-Based Compounds: Potential and Emerging Applications in Medicine; John Wiley&Sons Ltd.: Hoboken, NJ, USA, 2018. [Google Scholar]
  17. Hosmane, N.S. Boron Science: New Technologies and Applications; CRC Press: Boca Raton, FL, USA, 2012. [Google Scholar]
  18. Zhu, Y.; Hosmane, N.S. Applications and perspectives of boron-enriched nanocomposites in cancer therapy. Future Med. Chem. 2013, 5, 705–714. [Google Scholar]
  19. Sivaev, I.B.; Prikaznov, A.V.; Naoufal, D. Fifty years of the closo-decaborate anion chemistry. Collect. Czechoslov. Chem. Commun. 2010, 75, 1149–1199. [Google Scholar] [CrossRef]
  20. Mahfouz, N.; Abi Ghaida, F.; El Hajj, Z.; Diab, M.; Floquet, S.; Mehdi, A.; Naoufal, D. Recent Achievements on Functionalization within closo-Decahydrodecaborate [B10H10]2− Clusters. Chem. Sel. 2022, 7, e202200770. [Google Scholar]
  21. Abi-Ghaida, F.; Clement, S.; Safa, A.; Naoufal, D.; Mehdi, A. Multifunctional Silica Nanoparticles Modified via Silylated-Decaborate Precursors. J. Nanomater. 2015, 2015, 608432. [Google Scholar] [CrossRef] [Green Version]
  22. Abi-Ghaida, F.; Laila, Z.; Ibrahim, G.; Naoufal, D.; Mehdi, A. New triethoxysilylated 10-vertex closo-decaborate clusters. Synthesis and controlled immobilization into mesoporous silica. Dalton Trans. 2014, 43, 13087–13095. [Google Scholar] [CrossRef]
  23. Diab, M.; Mateo, A.; Al Cheikh, J.; Haouas, M.; Ranjbari, A.; Bourdreux, F.; Naoufal, D.; Cadot, E.; Bo, C.; Floquet, S. Unprecedented coupling reaction between two anionic species of a closo-decahydrodecaborate cluster and an Anderson-type polyoxometalate. Dalton Trans. 2020, 49, 4685–4689. [Google Scholar] [CrossRef] [PubMed]
  24. Izzet, G.; Volatron, F.; Proust, A. Tailor-made Covalent Organic-Inorganic Polyoxometalate Hybrids: Versatile Platforms for the Elaboration of Functional Molecular Architectures. Chem. Rec. 2017, 17, 250–266. [Google Scholar] [CrossRef] [PubMed]
  25. Proust, A.; Matt, B.; Villanneau, R.; Guillemot, G.; Gouzerh, P.; Izzet, G. Functionalization and post-functionalization: A step towards polyoxometalate-based materials. Chem. Soc. Rev. 2012, 41, 7605–7622. [Google Scholar] [CrossRef] [PubMed]
  26. Piot, M.; Hupin, S.; Lavanant, H.; Afonso, C.; Bouteiller, L.; Proust, A.; Izzet, G. Charge Effect on the Formation of Polyoxometalate-Based Supramolecular Polygons Driven by Metal Coordination. Inorg. Chem. 2017, 56, 8490–8496. [Google Scholar] [CrossRef] [Green Version]
  27. Shelly, K.; Knobler, C.B.; Hawthorne, M.F. Synthesis of monosubstituted derivatives of closo decahydrodecaborate(2-). X-ray crystal structures of [closo-2-B10H9CO] and [closo-2-B10H9NCO]2−. Inorg. Chem. 1992, 31, 2889–2892. [Google Scholar] [CrossRef]
  28. Mayer, C.R.; Fournier, I.; Thouvenot, R. Bis- and Tetrakis(organosilyl) Decatungstosilicate, [γ-SiW10O36(RSi)2O]4− and [γ-SiW10O36(RSiO)4]4−: Synthesis and Structural Determination by Multinuclear NMR Spectroscopy and Matrix-Assisted Laser Desorption/Desorption Time-of-Flight Mass Spectrometry. R. Chem. Eur. J. 2000, 6, 105–110. [Google Scholar] [CrossRef]
  29. Bonchio, M.; Carraro, M.; Scorrano, G.; Bagno, A. Photooxidation in Water by New Hybrid Molecular Photocatalysts Integrating an Organic Sensitizer with a Polyoxometalate Core. Adv. Synth. Catal. 2004, 346, 648–654. [Google Scholar] [CrossRef]
  30. Modugno, G.; Monney, A.; Bonchio, M.; Albrecht, M.; Carraro, M. Transfer Hydrogenation Catalysis by a N-Heterocyclic Carbene (NHC) Iridium Complex on a Polyoxometalate Platform. Eur. J. Inorg. Chem. 2014, 14, 2356–2360. [Google Scholar] [CrossRef]
  31. Carraro, M.; Modugno, G.; Fiorani, G.; Maccato, C.; Sartorel, A.; Bonchio, M. Organic-Inorganic Molecular Nano-Sensors: A Bis-Dansylated Tweezer-Like Fluoroionophore Integrating a Polyoxometalate Core. Eur. J. Org. Chem. 2012, 281–289. [Google Scholar] [CrossRef]
  32. Mayer, C.R.; Roch-Marchal, C.; Lavanant, H.; Thouvenot, R.; Sellier, N.; Blais, J.C.; Sécheresse, F. New Organosilyl Derivatives of the Dawson Polyoxometalate [α2-P2W17O61(RSi)2O]6−: Synthesis and Mass Spectrometric Investigation. Chem. Eur. J. 2004, 10, 5517–5523. [Google Scholar] [CrossRef]
  33. Maestre, J.M.; Lopez, X.; Bo, C.; Poblet, J.-M.; Daul, C. A DFT Study of the Electronic Spectrum of the α-Keggin Anion [CoIIW12O40]6−. Inorg. Chem. 2002, 41, 1883–1888. [Google Scholar] [CrossRef] [PubMed]
  34. Ravelli, D.; Dondi, D.; Fagnoni, M.; Albini, A.; Bagno, A. Predicting the UV spectrum of polyoxometalates by TD-DFT. J. Comput. Chem. 2011, 32, 2983–2987. [Google Scholar] [CrossRef] [PubMed]
  35. Pogula, L.P.; Ramakrishna, D.S. Charge Distribution and UV Absorption Spectra of Liquid Crystals with Structural Element [closo-B10H10] 2−: A Theoretical Approach. IARJSET 2015, 2, 92–95. [Google Scholar]
  36. Kemmegne-Mbouguen, J.C.; Floquet, S.; Zang, D.; Bonnefont, A.; Ruhlmann, L.; Simonnet-Jégat, C.; López, X.; Haouas, M.; Cadot, E. Electrochemical properties of the [SiW10O36(M2O2E2)]6− polyoxometalate series (M = Mo(v) or W(v); E = S or O). New J. Chem. 2019, 43, 1146–1155. [Google Scholar]
  37. Sivaev, I.B.; Prikaznov, A.V.; Anufriev, S.A. On relative electronic effects of polyhedral boron hydrides. J. Organomet. Chem. 2013, 747, 254–256. [Google Scholar] [CrossRef]
  38. Boujita, M.; Boixel, J.; Blart, E.; Mayer, C.R.; Odobel, F. Redox properties of hybrid Dawson type polyoxometalates disubstituted with organo-silyl or organo-phosphoryl moieties. Polyhedron 2008, 27, 688–692. [Google Scholar] [CrossRef]
  39. Liu, R.J.; Zhang, G.J.; Cao, H.B.; Zhang, S.J.; Xie, Y.B.; Haider, A.; Kortz, U.; Chen, B.H.; Dalal, N.S.; Zhao, Y.S.; et al. Enhanced proton and electron reservoir abilities of polyoxometalate grafted on graphene for high-performance hydrogen evolution. Energy Environ. Sci. 2016, 9, 1012–1023. [Google Scholar] [CrossRef]
  40. Keita, B.; Kortz, U.; Holzle, L.R.B.; Brown, S.; Nadjo, L. Efficient Hydrogen-Evolving Cathodes Based on Proton and Electron Reservoir Behaviors of the Phosphotungstate [H7P8W48O184]33− and the Co(II)-Containing Silicotungstates [Co6(H2O)30{Co9Cl2(OH)3(H2O)9(β-SiW8O31)3}]5− and [{Co3(B-β-SiW9O33(OH))(B-β-SiW8O29OH)2}2]22−. Langmuir 2007, 23, 9531–9953. [Google Scholar]
  41. Fourmond, V.; Jacques, P.A.; Fontecave, M.; Artero, V. H2 evolution and molecular electrocatalysts: Determination of overpotentials and effect of homoconjugation. Inorg. Chem. 2010, 49, 10338–10347. [Google Scholar] [CrossRef]
Figure 1. Molecular structures (DFT-optimized geometry) of (A) SiW10-APTES and (B) P2W17-APTES platforms highlighting the two different topologies of the APTES linker, and of (C) [B10H9CO] (X-ray diffraction structure from reference [27]). Legend: C in black, H in white, N in dark blue, Si in pink, O in red, B in blue, WO6 octahedra in orange and PO4 tetrahedra in green.
Figure 1. Molecular structures (DFT-optimized geometry) of (A) SiW10-APTES and (B) P2W17-APTES platforms highlighting the two different topologies of the APTES linker, and of (C) [B10H9CO] (X-ray diffraction structure from reference [27]). Legend: C in black, H in white, N in dark blue, Si in pink, O in red, B in blue, WO6 octahedra in orange and PO4 tetrahedra in green.
Molecules 27 07663 g001
Figure 2. Evolution of the proportions of the products in the system SiW10-APTES/B10H9CO/DIPEA as a function of B10H9CO/SiW10-APTES ratio at fixed DIPEA/B10H9CO ratio of 2. The proportion of each species are determined by integration of the 29Si NMR signals. Reproduced with permission from the doctoral thesis manuscript of Dr Manal Diab, University Paris Saclay/Lebanese University, May 2018.
Figure 2. Evolution of the proportions of the products in the system SiW10-APTES/B10H9CO/DIPEA as a function of B10H9CO/SiW10-APTES ratio at fixed DIPEA/B10H9CO ratio of 2. The proportion of each species are determined by integration of the 29Si NMR signals. Reproduced with permission from the doctoral thesis manuscript of Dr Manal Diab, University Paris Saclay/Lebanese University, May 2018.
Molecules 27 07663 g002
Figure 3. Scheme of syntheses of POM-borates adducts. The optimal quantities of reactants were determined by NMR titration studies. The reactions are performed in dry acetonitrile, at room temperature under inert atmosphere. Molecular structures are optimized geometry obtained by DFT. Legend: C in black, H in white, N in dark blue, Si in pink, O in red, B in blue, WO6 octahedra in orange and PO4 tetrahedra in green.
Figure 3. Scheme of syntheses of POM-borates adducts. The optimal quantities of reactants were determined by NMR titration studies. The reactions are performed in dry acetonitrile, at room temperature under inert atmosphere. Molecular structures are optimized geometry obtained by DFT. Legend: C in black, H in white, N in dark blue, Si in pink, O in red, B in blue, WO6 octahedra in orange and PO4 tetrahedra in green.
Molecules 27 07663 g003
Figure 4. Reflector positive ion MALDI-TOF spectra of (a) SiW10-APTES, (b) SiW10-monoB10, and (c) SiW10-diB10. Zooms of major peaks in the 3000–5000 m/z range are displayed with their respective simulated spectra.
Figure 4. Reflector positive ion MALDI-TOF spectra of (a) SiW10-APTES, (b) SiW10-monoB10, and (c) SiW10-diB10. Zooms of major peaks in the 3000–5000 m/z range are displayed with their respective simulated spectra.
Molecules 27 07663 g004
Figure 5. (a) 11B{1H} NMR spectra of SiW10-monoB10, SiW10-diB10 and TBA[B10H9CO] in CD3CN. (b) 1H NMR spectra of SiW10-monoB10, SiW10-diB10 and SiW10-APTES in CD3CN. * indicates the signal of the protonated amine DIPEAH+ present as a counter-cation. Reproduced with permission from the doctoral thesis manuscript of Dr Manal Diab, University Paris Saclay/Lebanese University, May 2018.
Figure 5. (a) 11B{1H} NMR spectra of SiW10-monoB10, SiW10-diB10 and TBA[B10H9CO] in CD3CN. (b) 1H NMR spectra of SiW10-monoB10, SiW10-diB10 and SiW10-APTES in CD3CN. * indicates the signal of the protonated amine DIPEAH+ present as a counter-cation. Reproduced with permission from the doctoral thesis manuscript of Dr Manal Diab, University Paris Saclay/Lebanese University, May 2018.
Molecules 27 07663 g005
Figure 6. (a) 183W NMR spectra of SiW10-monoB10, SiW10-diB10 and SiW10-APTES in CD3CN. (b) 183W NMR spectra of P2W17-diB10 and P2W17-APTES in CD3CN.
Figure 6. (a) 183W NMR spectra of SiW10-monoB10, SiW10-diB10 and SiW10-APTES in CD3CN. (b) 183W NMR spectra of P2W17-diB10 and P2W17-APTES in CD3CN.
Molecules 27 07663 g006
Figure 7. Optimized molecular structures of the POM-borates derivatives. SiW10-monoB10 (a) in «closed» form and (b) in «open» form; (c,d) two views of the most stable configuration of SiW10-diB10; (e,f) two views of the most stable configuration of P2W17-diB10. Dashed lines are given for shortest H-H contacts. Legend: C in black, H in white, B in blue, N in dark blue, Si in pink, WO6 octahedra in orange and PO4 tetrahedra in green.
Figure 7. Optimized molecular structures of the POM-borates derivatives. SiW10-monoB10 (a) in «closed» form and (b) in «open» form; (c,d) two views of the most stable configuration of SiW10-diB10; (e,f) two views of the most stable configuration of P2W17-diB10. Dashed lines are given for shortest H-H contacts. Legend: C in black, H in white, B in blue, N in dark blue, Si in pink, WO6 octahedra in orange and PO4 tetrahedra in green.
Molecules 27 07663 g007
Figure 8. Energetic profiles of the formation of mono- and di-adduct from the starting precursors in CD3CN. (a) SiW10-APTES, SiW10-monoB10 (open and closed isomers), and SiW10-diB10; (b) P2W17-APTES, P2W17-monoB10, and P2W17-diB10..
Figure 8. Energetic profiles of the formation of mono- and di-adduct from the starting precursors in CD3CN. (a) SiW10-APTES, SiW10-monoB10 (open and closed isomers), and SiW10-diB10; (b) P2W17-APTES, P2W17-monoB10, and P2W17-diB10..
Molecules 27 07663 g008
Figure 9. Frontier orbitals energies and energy gaps (eV) for the SiW10-APTES, SiW10-monoB10 and SiW10-diB10 species. Color code: W green, O red, Si light brown, B dark brown, C grey, N blue, H white; HOMO: red/blue; LUMO: orange/cyan. MO surfaces plotted at a 0.03 isovalue.
Figure 9. Frontier orbitals energies and energy gaps (eV) for the SiW10-APTES, SiW10-monoB10 and SiW10-diB10 species. Color code: W green, O red, Si light brown, B dark brown, C grey, N blue, H white; HOMO: red/blue; LUMO: orange/cyan. MO surfaces plotted at a 0.03 isovalue.
Molecules 27 07663 g009
Figure 10. Comparison of cyclic voltammograms (a) for the three compounds SiW10-APTES, SiW10-monoB10 and SiW10-diB10, and (b) for P2W17-APTES and P2W17-diB10 in the reduction part. The electrolyte was CH3CN + 0.1 M TBAClO4. Dashed lines are only guide for eyes.
Figure 10. Comparison of cyclic voltammograms (a) for the three compounds SiW10-APTES, SiW10-monoB10 and SiW10-diB10, and (b) for P2W17-APTES and P2W17-diB10 in the reduction part. The electrolyte was CH3CN + 0.1 M TBAClO4. Dashed lines are only guide for eyes.
Molecules 27 07663 g010
Figure 11. Cyclic voltammograms of (a) P2W17-APTES and (b) P2W17-diB10 after addition of variable amounts of acetic acid. (c) Plots of the cathodic currents measured at −2.2 V vs Fc+/Fc as a function of the ratio [acid]/[POM] for P2W17-APTES and P2W17-diB10. (d) Comparison of HER with and without catalyst P2W17-diB10, and with platinum after addition of an excess of acetic acid corresponding to the quantity added for a ratio [acid]/[POM] = 20. In all cases, the electrolyte was CH3CN + 0.1 M TBAClO4. The reference electrode was a saturated calomel electrode (SCE). Reproduced with permission from the doctoral thesis manuscript of Dr Manal Diab, University Paris Saclay/Lebanese University, May 2018.
Figure 11. Cyclic voltammograms of (a) P2W17-APTES and (b) P2W17-diB10 after addition of variable amounts of acetic acid. (c) Plots of the cathodic currents measured at −2.2 V vs Fc+/Fc as a function of the ratio [acid]/[POM] for P2W17-APTES and P2W17-diB10. (d) Comparison of HER with and without catalyst P2W17-diB10, and with platinum after addition of an excess of acetic acid corresponding to the quantity added for a ratio [acid]/[POM] = 20. In all cases, the electrolyte was CH3CN + 0.1 M TBAClO4. The reference electrode was a saturated calomel electrode (SCE). Reproduced with permission from the doctoral thesis manuscript of Dr Manal Diab, University Paris Saclay/Lebanese University, May 2018.
Molecules 27 07663 g011
Figure 12. Two views of the molecular electrostatic potential in atomic units (a.u.) projected onto an electron density isosurface (0.03 e·au−3) for P2W17-APTES, P2W17-monoB10 and P2W17-diB10 species.
Figure 12. Two views of the molecular electrostatic potential in atomic units (a.u.) projected onto an electron density isosurface (0.03 e·au−3) for P2W17-APTES, P2W17-monoB10 and P2W17-diB10 species.
Molecules 27 07663 g012
Table 1. Electrocatalytic efficiency for the reduction of protons into hydrogen at E = −2.2 V vs. Fc+/Fc for 20 equivalents of CH3COOH added in CH3CN.
Table 1. Electrocatalytic efficiency for the reduction of protons into hydrogen at E = −2.2 V vs. Fc+/Fc for 20 equivalents of CH3COOH added in CH3CN.
CompoundCatalytic Efficiency (%)
SiW10-APTES827
SiW10-monoB10524
SiW10-diB10548
P2W17-APTES854
P2W17-diB101340
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Diab, M.; Mateo, A.; El Cheikh, J.; El Hajj, Z.; Haouas, M.; Ranjbari, A.; Guérineau, V.; Touboul, D.; Leclerc, N.; Cadot, E.; et al. Grafting of Anionic Decahydro-Closo-Decaborate Clusters on Keggin and Dawson-Type Polyoxometalates: Syntheses, Studies in Solution, DFT Calculations and Electrochemical Properties. Molecules 2022, 27, 7663. https://doi.org/10.3390/molecules27227663

AMA Style

Diab M, Mateo A, El Cheikh J, El Hajj Z, Haouas M, Ranjbari A, Guérineau V, Touboul D, Leclerc N, Cadot E, et al. Grafting of Anionic Decahydro-Closo-Decaborate Clusters on Keggin and Dawson-Type Polyoxometalates: Syntheses, Studies in Solution, DFT Calculations and Electrochemical Properties. Molecules. 2022; 27(22):7663. https://doi.org/10.3390/molecules27227663

Chicago/Turabian Style

Diab, Manal, Ana Mateo, Joumada El Cheikh, Zeinab El Hajj, Mohamed Haouas, Alireza Ranjbari, Vincent Guérineau, David Touboul, Nathalie Leclerc, Emmanuel Cadot, and et al. 2022. "Grafting of Anionic Decahydro-Closo-Decaborate Clusters on Keggin and Dawson-Type Polyoxometalates: Syntheses, Studies in Solution, DFT Calculations and Electrochemical Properties" Molecules 27, no. 22: 7663. https://doi.org/10.3390/molecules27227663

Article Metrics

Back to TopTop