Next Article in Journal
β,β-Isomer of Open-Wells–Dawson Polyoxometalate Containing a Tetra-Iron(III) Hydroxide Cluster: [{Fe4(H2O)(OH)5}(β,β-Si2W18O66)]9−
Next Article in Special Issue
Zirconium-Catalyzed Alkene Hydrophosphination and Dehydrocoupling with an Air-Stable, Fluorescent Primary Phosphine
Previous Article in Journal
Synthesis, Characterization and Applications in Catalysis of Polyoxometalate/Zeolite Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Naphthyl-Containing Organophosphonate Derivatives of Keggin-Type Polyoxotungstates

1
Departamento de Química Inorgánica, Facultad de Ciencia y Tecnología, Universidad del País Vasco UPV/EHU, Bilbao 48080, Spain
2
BCMaterials, Parque Científico y Tecnológico de Bizkaia, Derio 48160, Spain
3
Servicios Generales de Investigación SGIker, Universidad del País Vasco UPV/EHU, Bilbao 48080, Spain
4
Departamento de Química Orgánica II, Facultad de Ciencia y Tecnología, Universidad del País Vasco UPV/EHU, Bilbao 48080, Spain
5
Departamento de Química Física, Facultad de Ciencia y Tecnología, Universidad del País Vasco UPV/EHU, Bilbao 48080, Spain
*
Authors to whom correspondence should be addressed.
Inorganics 2016, 4(2), 14; https://doi.org/10.3390/inorganics4020014
Submission received: 31 March 2016 / Revised: 27 April 2016 / Accepted: 3 May 2016 / Published: 12 May 2016
(This article belongs to the Special Issue Organophosphorus Chemistry 2016)

Abstract

:
New organophosphonate derivatives of monovacant Keggin-type polyoxotungstates that contain naphthyl groups have been synthesized and characterized in both solid state and solution. Single-crystal structural analysis shows that two phosphonate groups occupy the vacant position of the lacunary cluster unit in the isostructural compounds [N(C4H9)4]3[H(POC11H9)2(α-HBW11O39)] (TBA-1) and [N(C4H9)4]3[H(POC11H9)2(α-SiW11O39)] (TBA-2). Liquid-solution UV–Vis transmittance and solid-state diffuse reflectance spectroscopy studies reveal the presence of a new absorption band in the visible region, the charge transfer character of which has been further confirmed by time-dependent density functional theory (TD-DFT) calculations. The latter evidence that the charge transfer process is dominated by transitions from the highest occupied molecular orbital (HOMO), localized in the aromatic ring of the organic group, to the lowest unoccupied molecular orbital (LUMO), localized in the Keggin anion. Photoluminescence studies show that the fluorescent properties of the 1-naphthylmethylphosphonate group are quenched upon its incorporation into the inorganic oxo-tungstate skeleton. The solution stability of the hybrid clusters has been evaluated by a combination of 1H-, 13C- and 31P-Nuclear Magnetic Resonance spectroscopy and Electrospray Ionization-Mass Spectrometry. The hybrid polyanion [H(POC11H9)2(α-HBW11O39)]3− (1) herein constitutes the first structurally characterized organo-p-block containing borotungstate, and hence it confirms that this strategy for the organic functionalization of polyoxometalate clusters can be applied to new platforms belonging to the family of group-13 heteropolyoxotungstates.

Graphical Abstract

1. Introduction

The organic derivatization of polyoxometalates (POMs) represents one of the most relevant topics within the chemistry of this well-known family of anionic metal-oxo clusters with applications in current fields of interest such as catalysis, medicine and materials science [1,2,3,4,5]. The resulting organic–inorganic hybrids avoid the processing drawbacks that unmodified POMs can display for being incorporated into functional architectures or devices [6,7,8]. The covalent attachment of organic molecules to POM skeletons can confer tuned properties on the resulting hybrid species and can also allow for the immobilization of the clusters on diverse surfaces or matrixes. Nevertheless, highly elaborate functionalities for applications in, e.g., optics, molecular electronics or sensing often require the postfunctionalization of preformed hybrid POM platforms via multistep synthetic work [9].
Three main synthetic methods have been developed to prepare well-defined, solution-stable molecular hybrid POMs [10]. These include strategies involving the replacement of shell O atoms with O- or N-donor ligands as exemplified by tris(alkoxo)-capped Anderson–Evans clusters [11], hexavanadates [12] or [H4P2V3W15O62]5− anions [13], as well as by organoimido/diazenido derivatives of Lindqvist-type molybdates [14]. Another approach is based on the reaction between 3d- and/or 4f-metal substituted POMs with exposed centers toward multidentante organic ligands [15,16,17,18]. However, the most explored route consists in combining lacunary polyoxotungstates with p-block organoderivatives. The nucleophilic character of the oxygen atoms delimiting the lacunae is significantly increased compared with those of the plenary anion, and this greatly favors the reaction with electrophilic groups. The use of organotin entities has resulted in some interesting compounds [19,20,21] and a few examples of organogermyl, -arsenyl and -stibyl functionalized POMs can also be found in the literature [22,23]. Nevertheless, organosilyl and -phosphoryl groups constitute the derivatizing agents that have attracted the greatest interest since the first report by Knoth in the early 1980s [24].
In the case of organosilyl derivatives, a large variety of functionalities have been incorporated into monolacunary Keggin or Wells–Dawson anions containing up to two organogroups or one dimeric μ-oxo-bridged (RSi)2O unit per vacant site [25,26]. In regard to di- and trilacunary anions, two to four organosilyl entities have been selectively anchored to the POM vacant sites [27,28]. The functionalization of mono-, di- and trilacunary Keggin-type phospho- or silicotungstates with organophosphonate moieties has been thoroughly explored as well [29,30,31]. Some reports on the reactivity of Wells–Dawson [P2W17O61]10− and Keggin-type [XW11O39]n (X = B, Ga, Ge) monovacant frameworks can be found in the literature, but they do not include any X-ray crystal structure analysis [32,33,34]. The immobilization of such type of hybrid POM platforms in silica matrixes [35], metallic [36,37] or polymeric [38] nanoparticles, and silicon or metallic surfaces has resulted in advanced functional materials targeting the heterogenization of homogeneous catalysts [39], molecular memories [40] and selective biomolecule-adhesion properties [41].
In order to incorporate custom-designed organic moieties to the clusters, two strategies are usually followed. The first approach consists in preparing the desired p-block organo-group, followed by its reaction with lacunary POMs. In the second method, the p-block organoderivative with a reactive pendant group is first inserted in the vacant site of the cluster and the resulting hybrid POM species is then used as platform for further postfunctionalization. Following these two synthetic routes and making use of click chemistry or Pd catalyzed coupling reactions, photoactive electron transfer complexes that mimic the bio-inspired artificial photosynthesis have been prepared among others [42,43,44,45]. Magnetically and catalytically active d-metal centers have also been included in the structure by using organic functionalities suitable for metal coordination [46]. Alternatively, oxophilic 3d- or 4f-metals have been incorporated by taking advantage of the free and accessible O atoms in the phosphonate groups of Keggin-type POM derivatives [47,48]. Organophosphonate-functionalized POM species have shown interesting catalytic properties for oxidation reactions by themselves [49], and furthermore, compounds bearing coordinated metal centers represent original molecular models for isolated single-site catalytic species dispersed onto matrixes like silica [50].
Herein we report the synthesis and full characterization in both the solid state and liquid solution of 1-naphthylmethylphosphonate-containing derivatives of monovacant Keggin anions, namely [N(C4H9)4]3[H(POC11H9)2(HBW11O39)] (TBA-1) and [N(C4H9)4]3[H(POC11H9)2(α-SiW11O39)] (TBA-2). These hybrid POM platforms have been prepared with the aim of being functionalized with 3d- or 4f-metal centers as models for supported site-isolated catalysts with an aromatic environment. The aromatic 1-naphthylmethylphosphonate subunit should not only confer photoactive properties on the system, but could also enhance the selectivity of the catalytic process by interacting with the substrate and providing steric hindrance. We selected the [HBW11O39]8− anion for our studies to determine whether borotungstates could be organically derivatized following this route in spite of its peculiar reactivity when compared with analogues that contain heteroatoms from groups 14 and 15, and we employed the isomorphic and topologically equivalent [SiW11O39]8− as pattern in the reaction with 1-naphthylmethylphosphonic acid.

2. Results and Discussion

2.1. Synthesis, Infrared Spectroscopy and Thermal Analyses

Organic functionalization of Keggin-type anions with organophosphonate groups was carried out using the well-established method for the incorporation of organosilyl and -phosphoryl moieties into lacunary POM anions developed by Mayer and Thouvenot [30]. Stoichiometric reactions of the monolacunary Keggin-type anions [HnXW11O39]8− (X = B, n = 1; X = Si, n = 0) and 1-naphthylmethylphosphonic acid in acetonitrile were carried out with tetrabutylammonium (TBA) bromide as phase transfer agent to dissolve the POM cluster in the organic media. After an overnight reflux and filtering KBr and traces of unreacted POM precursor off, the crude products were obtained upon solvent removal under vacuum. According to 31P-NMR experiments, these crude products were not pure as multiple signals were observed in the spectra. Recrystallization in acetonitrile allowed us not only to obtain single crystals suitable for X-ray diffraction studies, but also to isolate the title compounds as pure, homogeneous phases. All the further characterization was performed on these pure crystalline batches.
The original synthesis was carried out employing the [HBW11O39]8− anion as POM precursor to evaluate whether lacunary heteropolytunsgtates with heteroatoms of the group 13 can act as inorganic platforms for their derivatization with p-block organo-groups. This question is yet to be experimentally confirmed as straightforward analogies with other monovacant Keggin-type species are usually unfeasible due to the peculiar behavior of borotungstates in aqueous solution when compared with isoelectronic species like [XW11O39]8− (X = Si, Ge). For example, the reaction of tungstate ions with [HBW11O39]8− does not directly lead to the plenary α-[BW12O40]5− Keggin ion. Instead, one ditungstate fragment places over the vacant site to lead to the [BW13O46H3]8− species, which further condenses into oligomeric anions at lower pH values [51]. To our knowledge, there are only three examples of structurally characterized molecular hybrid borotungstates: the [{Ru(L)}2(HBW11O39)2WO2]10− (L = p-cymene, benzene) clusters based on the {HBW11} unit [52] and the dimethyltin-containing [{(CH3)2Sn}6(OH)2O2(H2BW13O46)2]12− anion showing {H2BW13} subunits [53]. Although the insertion of organophosphonate groups in monolacunay borotungstates has been previously suggested [32], compound TBA-1 reported herein represents the first example in the literature of a structurally characterized group-13 heteropolytungstate that is derivatized with organo-p-block moieties. The use of the topologically equivalent [SiW11O39]8− precursor in combination with 1-naphthylmethylphosphonic acid led to the isolation of the isostructural compound TBA-2.
The functionalization of the POM precursor was firstly identified by infrared spectroscopy (FT-IR). FT-IR spectra of compounds TBA-1 and TBA-2 are depicted in Figure 1 together with those of their corresponding K8[HBW11O39]·13H2O ({HBW11}) and TBA8[SiW11O39] ({SiW11}) precursors. Both compounds exhibit characteristic bands of strong intensity in the region below 1000 cm−1 that are associated with monolacunary α-Keggin-type anions. The antisymmetric stretching vibrational bands νas(X–Oc), νas(W–Ot) and νas(W–Ob–W) at 1220, 971 and 734 cm−1 for TBA-1 are shifted in comparison with those of {HBW11}, which appear at 1236, 954 and 752 cm−1, respectively. In the case of TBA-2, these signals are observed at 1028, 978, and 802 cm−1 and undergo blue shifts by ca. 10 cm−1 when compared with those of the parent {SiW11} anion. The presence of organophosphonate groups in both compounds is confirmed by the signals associated with the stretching of the P–O bonds that appear in the 1060–1080 cm−1 range. The organic region above 1100 cm−1 is dominated by signals of medium to strong intensity that are observed in the 1176–1300 cm−1 range and associate with the C–N and C–H bonds from the TBA cations and the stretching vibrations of C=C bonds in the naphthalenic system. It is worth highlighting that the protonation of the central O atom in the borotungstate subunit of TBA-1 can be unequivocally inferred by IR spectroscopy on the basis of the νas(B–O–H) vibrational band observed at 1220 cm−1, which is red shifted by 20 cm−1 in comparison with that of the precursor [54].
Thermogravimetric analyses (TGA) show curves with similar profiles for both compounds (Figure S1). The wide thermal stability range, which extends up to ca. 280 °C for TBA-1 and to ca. 290 °C for TBA-2, originates from the absence of any solvent molecules in the solid samples upon being left to dry overnight in air. The first weight loss is completed at 415 °C and comprises the loss of ca. 19% of the total mass, which accounts for the 3 tetrabutylammonium molecules (%mass, calcd. (found) for 3 × C16H36N: 19.33 (19.55) for TBA-1 and 19.25 (19.08) for TBA-2). The final step, which corresponds to the organic ligand combustion and POM breakdown, extends continuously up to temperatures in the 800–850 °C range (%mass, calcd. (found) for two naphthylmethyl groups, 2 × C11H9: 8.33 (8.40) for TBA-1 and 8.31 (8.25) for TBA-1). The final residues are 72.1% of the initial mass for TBA-1 (calcd. for HBP2W11O40: 72.7%) and 72.7% for TBA-2 (calcd. for P2SiW11O40: 72.8%).

2.2. Crystal Structure

Compounds TBA-1 and TBA-2 are isostructural and crystallize in the triclinic space group P−1 with two [H(C11H9PO)2(α-HnXW11O39)]3− (X = B, n = 1 (1); X = Si, n = 0 (2)) hybrid POMs and six tetrabutylammonium (TBA) cations in the asymmetric unit. Unfortunately, only five out of the six TBA cations could be satisfactorily modeled in the crystal structure of TBA-2 due to severe crystallographic disorder, whereas the poorer quality of crystals of TBA-1 only allowed us to locate one TBA ion. The molecular structures of the hybrid POMs 1 and 2 consist of a monolacunary α-Keggin-type anion that incorporates two 1-naphthylmethylphosphonate moieties in the vacant position as antenna ligands via corner-sharing of the phosphonate group with one {W2O10} unit and one {W3O13} trimer (Figure 2). The cluster shows idealized Cs symmetry with a mirror plane that contains the heteroatom and the bridging O atoms of the {W2O10} unit (Figure 2B). This ideal symmetry is broken by the relative arrangement of the naphthyl residues (Figure 2C). The planes containing the aromatic systems form angles in the 15°–25° range with the ideal mirror plane. However, both naphthyl residues from each of the POM units are arranged in a staggered conformation. The aromatic systems are rotated about 20° with respect to each other as defined by the <centroid···P···P···centroid> torsion angle involving equivalent ring centroids in both residues. As shown in Figure 2D, this orientation allows for establishing bifurcated intramolecular C–H···OPOM hydrogen bonds that involves the methylenic groups pointing towards the polyanionic surface and bridging O atoms from the POM unit (Table 1).
The P atoms are bonded to the methylenic C atom from the 1-naphthylmethyl moiety, two O atoms from the POM and one terminal O atom (Table S1), and this bonding scheme results in distorted tetrahedral geometries. Bond Valence Sum Calculations (BVS) [55] indicate partial protonation of the terminal O atoms in the phosphonate moieties (BVS values: 1.26–1.45) in good agreement with the number of cations determined by elemental and thermal analyses. One proton is disordered over the P=O bonds of the two organophosphonate groups for each of the hybrid POMs in the asymmetric unit. BVS calculations also indicate protonation of the central OPOM atom pointing toward the vacant position in both crystallographically independent polyanions of compound TBA-1 (BVS values: 1.15–1.17) and this fact is fully consistent with the presence of the B–O–H stretching band in the IR spectrum.
The relative arrangement of the naphthyl residues allows the insertion of a bulky TBA cation (N1 and N4), in such a way that this results embraced between the two aromatic systems via nonconventional cation···π [56] and C–H···π type interactions [57] (Figure 2E). The distances of the C atoms from the TBA cations to the ring centroids (3.683(4)–4.188(6)Å) compare well with those calculated in the literature for such type of C–H···π contacts [58]. The crystal packing displays a pronounced two-dimensional character with hybrid anions linked through C–H···OPOM hydrogen bonds in bilayers that stack along the [001] direction. These bilayers comprise POM clusters located at z = 0.25 and z = 0.75, and arrange in such a way that each POM is connected to two neighbors of the same z level along the crystallographic y-axis and to another two in the contiguous hemilayer through a hydrogen bonding network involving the naphthyl groups (Figure 3, Table S2). The N1 and N4 TBA cations placed between naphthyl groups reinforce the connectivity within the bilayer. Moreover, the N3 and N5 cations connect POMs of the same z level in a zigzag mode along the [010] direction through an additional set of massive C–H···OPOM hydrogen bonds, whereas the remaining N2 species (and most likely the sixth undetermined TBA cation) occupy interlamellar spaces.

2.3. UV–Vis Spectroscopy and Photoluminescent Properties

The electronic properties of the title compounds have been explored by both solution and diffuse reflectance UV–Vis spectroscopy. Figure 4 displays the UV–Vis spectra of acetonitrile solutions of TBA-2 with different concentrations (c1 = 10−4 M; c2 ≈ 10−6 M; c3 ≈ 10−8 M), together with those obtained for the 1-naphthylmethylphosphonic acid and the TBA8[SiW11O39] POM precursor, which has been prepared for comparative purposes following reported procedures [59]. The rapid transformation of {BW11} to {BW12} in slightly acidic aqueous medium precluded us from preparing the corresponding TBA8[HBW11O39] analogue to perform similar comparative studies for the TBA-1 derivative. Methathesis with TBABr led to the isolation of the very stable plenary Keggin anion α-[BW12O40]5− as a TBA salt according to IR spectroscopy. Samples with different concentrations of the title compounds have been used for our studies due to the great differences in the relative intensity of the absorption bands.
The electronic spectrum of the parent {SiW11} precursor displays an absorption maximum at 265 nm that is associated with the O→W ligand-to-metal charge transfer (LMCT) transition of the inorganic POM framework. In turn, the 1-naphthylmethylphosphonic acid shows two absorption bands at 224 and 289 nm, which correspond to the well-known π−π* transitions of the naphthalenic system [60]. The former band is also present in the spectrum of TBA-2 (see the spectrum of the 10−8 M solution) and is accompanied by a second broad absorption centered at 275 nm that originates from a small bathochromic shift of about 10 nm of the LMCT band upon functionalization of the monolacunary cluster with organophosphonate groups. Such type of bathochromic shift is indicative of an electronic transfer between the conjugated organic fragment and the metal-oxo cluster [61]. The spectrum recorded for the 10−6 M solution affords a better resolution in this spectral region. The absorption band is considerably wider than that observed for the LMCT transition in the parent {SiW11} precursor and it shows a shoulder at 283 nm that is in good correspondence with the naphthyl π−π* transitions observed for the phosphonic acid in this region. The combination of these two features strongly indicates that the absorption of TBA-2 centered at 275 nm has a strong contribution of the naphthalenic system.
The spectrum of the concentrated solution of TBA-2 shows and additional band centered at ca. 400 nm that could be ascribed to a charge transfer between the aromatic system and the inorganic POM cluster in which the former acts as electron donor and the latter as acceptor. This phenomenon has been previously observed for similar compounds in the literature [61,62]. The charge-transfer process can be noticed visually because the incorporation of the colorless 1-naphthylmethylphosphonate moiety into the inorganic skeleton of the colorless {SiW11} leads to the formation of the orange-yellow compound TBA-2. A similar behavior is observed in the UV–Vis spectra of TBA-1 solutions (Figure S2). The bands at ca. 222 and 282 nm can be associated with the well-known π−π* transitions of the naphthyl residue, whereas that at 274 nm corresponds to the LMCT process. The maximum for the charge transfer band between the naphthyl residue and the inorganic POM cluster is blue-shifted by 15 nm upon comparison with that of TBA-2 (ca. 385 nm in 1 vs ca. 400 nm in 2). Accordingly, crystals of compound TBA-1 display a more yellowish color than those of TBA-2 (Figure 5).
In order to analyze the nature of the band centered at ca. 400 nm, theoretical calculations for the hybrid POM 2 at the TDA-B3LYP/def2-SVP//BP86-D3/def2-SVP level of theory have been performed. They show qualitative agreement with the experimental results, with several low intensity transitions in the visible region (408, 380 and 350 nm) that fall under the very wide experimental band centered at 400 nm, as well as a cluster of higher intensity transitions centered at 290 nm (Figure 4). The difference density plots of the molecular orbitals involved in the low intensity bands clearly indicate their organic to inorganic charge transfer character, with the naphthalenic system acting as electron donor and the POM cluster as acceptor (W dxy with a component of O p orbitals). On the other hand, the difference densities calculated for the cluster of transitions centered at 290 nm clearly support their previous assignment as O→W LMCT transition in the inorganic POM framework (Figure 6D) combined with π−π* transitions in the naphthyl groups (Figure 6E).
The diffuse reflectance spectra registered for powdered crystalline samples of TBA-1, TBA-2 and their precursors K8[HBW11O39]·13H2O and K8[SiW11O39]·13H2O (Figure 7) confirm the features stated from the solution studies. The bands associated with the monolacunary Keggin anions undergo red shift upon organic functionalization from 313 and 318 nm for the {HBW11} and {SiW11} POM precursors to 336 and 341 nm for the hybrid compounds TBA-1 and TBA-2. Strong charge transfer bands originating from the covalent connection between the naphthyl moieties and the inorganic oxo-cluster can also be observed in the visible region.
Naphthalene and its derivatives present photoluminescent properties when they are excited with UV light [60]. To determine whether these emission properties are modified upon the incorporation of the 1-naphthylmethylphosphonate groups into monovacant Keggin anions, we measured the fluorescence spectra of 10−6 M acetonitrile solutions of the title compounds TBA-1 and TBA-2 and compared the results with that recorded for commercial 1-naphthylmethylphosphonic acid (Figure 8). The spectrum of the free phosphonic acid (λexc = 283 nm) shows a broad emission band with the typical naphthalenic vibrational structure in the ca. 300−400 nm spectral range. However, our hybrid polyanions 1 and 2 do not display any emission when being excited at any wavelength corresponding to their absorption maxima (225, 275 or 280 nm), nor even at the charge transfer band (385–400 nm). Similar results were obtained when the experiments were carried out on solid crystalline samples (Figure S3). Nevertheless, a weak fluorescence reminiscent of that originating from the naphthyl group was obtained when 1 and 2 were excited at 283 nm, but only when the spectrofluorimeter was set to work in a High Sensitivity mode. The quenching of the naphthalenic emission in the hybrid POMs 1 and 2 could result from two different processes: (i) the strong absorption of the Keggin-type cluster at the excitation wavelength (ca. 280 nm) prevents the excitation of the fluorophore; or (ii) the quenching of the fluorescence proceeds via naphthyl-to-cluster charge transfer. The former could in principle be disregarded according to the UV–Vis spectrum of TBA-2 (Figure 4), which unequivocally shows that the naphthyl groups undergo strong absorption at 283 nm. Therefore, the quenching of the naphthyl fluorescence upon incorporation of the organophosphonate group into the Keggin-type skeleton is more likely to take place via naphthyl-to-POM charge transfer, which allows the excited state to deactivate through non-radiative processes. TD-DFT calculations support the latter interpretation as they clearly show that the naphthyl groups in the hybrid POMs 1 and 2 are excited in the 280–290 nm range (Figure 6D,E). Analogous quenching of the photoluminescent properties of aromatic systems such as pyrene upon incorporation into POM cluster frameworks has been previously reported in the literature for hybrid species similar to polyanions 1 and 2 [62].

2.4. Solution Behavior

In order to investigate the solution stability of the hybrid POMs 1 and 2, we performed 31P-, 1H- and 13C-nuclear magnetic resonance (NMR) studies in deuterated solvents and compared the resulting spectra with that of the commercial 1-naphthylmethylphosphonic acid.
The 31P-NMR spectra of compounds TBA-1 and TBA-2 in acetone (Figure 9) exhibit one singlet at 27.9 and 24.5 ppm, respectively. This fact indicates: (i) the absence of any phosphorous containing impurities in the sample; and (ii) that the attachment mode of the two organic groups to the lacunary anion is equivalent for both hybrid anions in good agreement with the molecular structures determined from single-crystal X-ray diffraction studies. All the chemical shifts lay in the expected range for organophosphonate derivatized monolacunary Keggin type POMs [63]. Depending on the heteroatom of the Keggin-type fragment, the signal shifts considerably upfield (2, Δδ = 2.9 ppm) or slightly downfield (1, Δδ = 0.5 ppm) upon comparison with that originated from the parent commercial phosphonic acid (δ = 27.4 ppm).
The title compounds TBA-1 and TBA-2 display similar 1H-NMR spectra that are depicted in Figure S4 together with that recorded for the naphthylmethylphosphonic acid. The grafting of the organophosphonate moiety could be confirmed by the absence of any signal corresponding to hydroxyl groups in the spectra of TBA-1 and TBA-2, which for the phosphonic acid appears as a broad singlet at δ = 10.59 ppm that integrates for two protons. The characteristic resonances for the n-butyl groups of the TBA cations observed at ca. 1 (–CH3), 1.5 (–CH2–CH3), 1.8 (–CH2–CH2–) and 3.4 (–N–CH2–) ppm have been integrated for 36, 24, 24 and 24 protons, and this fact confirms the presence of three cations per POM unit. Protons in the aliphatic n-butyl chain are less shielded as they place closer to the N atom and show the expected spin coupling constant of J = 7.4 Hz [64]. The signals originating from the methylenic group of the organophosphonate moiety undergo drastic changes upon incorporation into the POM vacant position. The spin system, determined as two singlets at 3.47 and 3.51 ppm in the phosphonic acid, becomes diastereotopic in the hybrid POM structure in such a way that four different protons can be differentiated: Ha and Hb from one of the P(O)CH2C10H7 moieties and Hc and Hd from the other one (Figure 10). These four resonances appear as doublets with a geminal coupling constant of about J = 15.2 Hz and arrange into two well-separated groups of signals because those that originate from the two H atoms pointing at the POM surface undergo significantly larger shift downfield due to the intramolecular C–H···OPOM interactions, which promote the deshielding of the protons [15] (Ha and Hc: Δδ = 0.69–0.76 ppm for TBA-1 and Δδ = 0.62–0.66 ppm for TBA-2; Hb and Hd: Δδ = 0.42–0.47 ppm for TBA-1 and Δδ = 0.34–0.38 ppm for TBA-2). It is also worth noting the roof effect of the signals, which is characteristic of almost overlapping resonances showing similar intensity. In regard to the naphthylic H atoms, the spectra of TBA-1 and TBA-2 show different multiplets in the δ = 7.35–8.29 ppm range that are analogous to those observed for the phosphonic acid in the δ = 7.41–8.12 ppm region. On the other hand, the singlet located at 4.80 ppm in the spectrum of TBA-1 can only be ascribed to the protonation of the central OPOM atom (B–O–H) in the vacant position of the cluster, which has also been observed in the IR spectrum and inferred from BVS calculations on the molecular structure of 1. These results also suggest that the hybrid POMs 1 and 2 show solution stability in organic media like acetone.
The 13C-NMR experiments were first carried out for deuterated acetone solutions (Figure S5). The TBA cations in compounds TBA-1 and TBA-2 could be easily assigned to singlets appearing at 14 (–CH3), 20 (–CH2–CH3), 25 (–CH2–CH2–) and 59 (–N–CH2–) ppm. The aromatic signals are observed in the 126–135 ppm region. Some of them appear as doublets due to the spin coupling with the close NMR active P nucleus (J = 3–11 Hz). The spectrum of the 1-naphthylmethylphosphonic acid is very similar to that observed for the hybrid POMs with an additional doublet at 32.2 ppm that is ascribed to the methylenic group, which shows a large 1J = 137.5 Hz spin-coupling constant with the vicinal P atom. Unfortunately, the resonances corresponding to this group in compounds TBA-1 and TBA-2 were partially overlapped with the strong signal originating from the solvent and this fact prevented us from performing a full analysis. Therefore, the final 13C-NMR spectra for compounds TBA-1 and TBA-2 were registered using solutions in deuterated acetonitrile (Figure 11). The TBA cations appear as three singlets at 13.9 (–CH3), 20.4 (–CH2–CH3) and 24.4 (–CH2–CH2–) ppm and a resonance at 59.5 ppm for TBA-1 and 59.3 ppm for TBA-2 that corresponds to the C atoms covalently bonded to the N atoms. Close inspection of this resonance reveals an apparent triplet, which is most likely the result of three singlets arising from the three virtually identical but inequivalent TBA cations. The phosphonate groups are observed as follows: one doublet corresponding to the methylene group at ca. 59.4 ppm with 1J = 134.1 Hz for TBA-1 and 1J = 138.3 Hz for TBA-2, together with a set of resonances spanning in the 126–135 ppm region. The positions for singlets and doublets of the aromatic region are virtually identical in both cases: 126.5 (J = 4.5 Hz), 126.6, 126.9, 128.5 (J = 4.5 Hz), 129.2, 129.5 (J = 11.0 Hz), 130.0 (J = 8.0 Hz), 133 (J = 5.5 Hz) and 135 (J = 2.5 Hz) ppm. The spin coupling constants of 11.0, 8.0, and <5.5 Hz have been tentatively assigned as 2J, 3J and 4J, respectively.
The Electrospray-Ionization Mass Spectrometry (ESI-MS) studies confirm the solution stability of the hybrid polyanions 1 and 2. The spectrum of TBA-1 recorded in negative-ion mode displays two main groups of signals spanning from m/z 1000 to 1700. These signals have been attributed to the intact anionic species 1 with 3− and 2− charge states and different extents of associated counterions according to isotopic pattern inspection and m/z spacing. The main species detected corresponds to the pristine anion {1}3− (m/z 1011.8), where {1} = [H(C11H9PO)2(HBW11O39)], whereas the second maxima originates from the intact species with an associated tetrabutylammonium cation, {1 + TBA}2− (m/z 1638.3). Figure 12 shows the ESI mass spectrum together with a comparison of the experimental and calculated isotopic patterns for these two specific ions. Inspection of the crystal structure of 1 revealed two protonation sites accounting for one disordered proton and the presence of one TBA cation embraced between the two naphthyl residues through multiple C–H··· π type interactions. These two features could still be retained in solution according to the ESI-MS results. Analogous experiments carried out for the TBA-2 derivative reveals a similar behavior. The most abundant species detected corresponds also to the intact hybrid anion {2}3− (1017.4 m/z) and to the species with an associated TBA cation {2 + TBA}2− (m/z 1646.8) (Figure S6). To summarize, the combination of multinuclear (1H, 13C and 31P) NMR studies and ESI-MS experiments confirms that the hybrid POMs 1 and 2 are stable in polar organic media like acetone or acetonitrile and do not dissociate into any of their constituent building blocks.

3. Experimental Section

3.1. Materials and Methods

The precursors K8[α-SiW11O39]·13H2O and K8[α-HBW11O39]·13H2O were synthesized according to literature procedures [51,65] and identified by IR spectroscopy. All other reagents were purchased from commercial sources and used without further purification. Carbon, nitrogen, and hydrogen were determined on a EuroVector EA 3000 CHNSO analyzer (EuroVector, Milan, Italy). Fourier transform infrared (FT-IR) spectra were recorded on KBr pellets using a Shimadzu FTIR-8400S spectrophotometer (Shimadzu, Kyoto, Japan). Diffuse Reflectance studies were performed on a UV–Vis-NIR Varian Cary 500 spectrophotometer (Varian, Palo Alto, CA, USA). UV–Vis and fluorescence spectra were, respectively, registered on a Shimadzu UV-1240 spectrophotometer and a Shimadzu RF-5300 spectrofluorimeter equipped with a 150 W Xenon arc lamp (Shimadzu, Kyoto, Japan) using standard quartz cuvettes. Thermogravimetric analyses were carried out from room temperature to 850 °C at a rate of 5 °C·min−1 on a Mettler-Toledo TGA/SDTA851e thermobalance (Mettler Toledo, Greifensee, Switzerland) under a 50 cm3·min−1 flow of synthetic air. 1H- and 13C-NMR spectra were acquired on a Bruker AC-500 spectrometer (500 MHz for 1H and 125.7 MHz for 13C) (Bruker, Karlsruhe, Germany). Spectra were referenced to external tetramethylsilane via the residual protonated solvent (1H) or the solvent itself (13C). All chemical shifts (δ) are reported in parts per million (ppm). For (CD3)2CO (acetone-d6) the shifts are referenced to 2.05 ppm for 1H-NMR spectroscopy and 29.84 ppm for 13C-NMR spectroscopy. For CD3CN the shifts are referenced to 118.26 ppm for 13C-NMR spectroscopy [66]. Abbreviations used in the description of resonances are: s (singlet), d (doublet), t (triplet), q (quartet), m (multiplet). Coupling constants (J) are quoted to the nearest 0.1 Hz. Proton-decoupled 31P-NMR spectra were recorded on the same spectrometer working at 212.4 MHz using 85% phosphoric acid as internal standard (0 ppm). The ESI mass spectra for 10−5 M acetonitrile solutions were recorded in a Waters QTOF Premier instrument (Walters Corporation, Milford, MA, USA) with orthogonal Z-spray electrospray interface. The solution was introduced at a flow rate of 10 mL·min−1 and N2 was used as the desolvation and cone gas (flow rates of 300 and 30 L·h−1, respectively). A capillary voltage of 3.3 kV in the negative scan mode (V-mode) and relatively low cone voltages (Uc = 15 V) were used to control the extent of fragmentation of the gas-phase detected species.

3.2. Synthesis of [N(C4H9)4]3[H(C11H9PO)2(HBW11O39)] (TBA-1)

To a suspension of K8[HBW11O39]·13H2O (3.22 g, 1 mmol) in CH3CN (50 mL) stirred at room temperature solid [N(C4H9)4]Br (1.605 g, 5 mmol) and C11H9PO3H2 (0.444 g, 2 mmol) were added. After 10 min, HCl 4 M (1 mL, 4 mmol) was added dropwise and the mixture was left overnight at reflux. Then, the insoluble white solid was removed by filtration and an orange powder was obtained by evaporation of the resulting solution in a rotary evaporator. The crude compound was washed with water and recrystallized from CH3CN. Prismatic orange crystals of TBA-1 suitable for X-ray diffraction were obtained upon slow evaporation of the final solution in an open vial for ca. 3 days. Yield: 2.33 g, 62% based on W. Elemental Analyses (%): calcd. (found) for C70H128BN3O41P2W11: C, 22.3 (23.0); H, 3.48 (3.71); N, 1.14 (1.15). IR (cm−1): 3404 s, 2968 m, 2871 w, 1482 s, 1465 m, 1381 m, 1219 m,1176 w, 1155 w, 1058 s, 1029 s, 973 vs, 918 vs, 896 vs, 828 vs, 738 s, 560 w, 533 m, 425 w. 1H-NMR (500 MHz, acetone-d6): δ = 0.99 (36H, t, J = 7.4 Hz, –CH3), 1.46 (24H, tq, J =7.4 Hz, 7.4 Hz, –CH2–CH3), 1.76–1.84 (24H, m, –CH2–CH2–), 3.36–3.43 (24H, m, –N–CH2–), 3.89 (1H, d, J = 15.3 Hz, P(O)CHaHb), 3.94 (1H, d, J = 15.3 Hz, P(O)CHcHd), 4.22 (1H, d, J = 15.3 Hz, P(O)CHaHb), 4.27 (1H, d, J = 15.3 Hz, P(O)CHcHd), 4.80 (1H, s, B–O–H), 7.36 (2H, ddd, J = 8.3, 7.0, 1.1 Hz, CaromH), 7.40–7.50 (4H, m, CaromH), 7.73 (2H, dd, J = 8.3, 2.8 Hz, CaromH), 7.78–7.84 (4H, m, CaromH), 8.29 (2H, d, J = 8.3Hz, CaromH) ppm. 13C-NMR (CD3CN): δ = 13.9 (s, –CH3), 20.4 (s, –CH2–CH3), 24.4 (s, –CH2–CH2–), 29.9 (d, 1J = 134.1 Hz, P(O)CH2), 59.5 (s, –N–CH2–), 125.9 (s, Carom), 126.5 (d, J = 4.5 Hz, Carom), 126.7 (s, Carom), 127(s, Carom), 128.5 (d, J = 4.5 Hz, Carom), 129.1 (d, J = 11.4 Hz, Carom), 129.3 (s, Carom), 130.0 (d, J = 7.7 Hz, Carom), 133.2 (d, J = 5.5 Hz, Carom), 134.8 (d, J = 2.7 Hz, Carom) ppm. 31P-NMR (acetone-d6): δ = 27.9 ppm.

3.3. Synthesis of [N(C4H9)4]3[H(C11H9PO)2(SiW11O39)] (TBA-2)

The synthetic procedure above was followed but for using K8[SiW11O39]·13H2O (3.22 g, 1 mmol) instead of K8[HBW11O39]·13H2O. Prismatic orange crystals of TBA-2 suitable for X-ray diffraction were obtained upon slow evaporation of the final solution in an open vial for ca. 3 days. Yield: 2.12 g, 56% based on W. Elemental Analyses (%): calcd. (found) for C70H127N3O41P2SiW11: C, 22.3 (22.7); H, 3.49 (3.61); N, 1.14 (1.22). IR (cm−1): 3430 s, 2965 m, 2873 w, 1642 m, 1479 w, 1376 w, 1176 w, 1155 w, 1078 s, 978 vs, 936 s, 918 vs, 889 s, 853 m, 731 s, 550 w, 531 w, 423 w. 1H-NMR (500 MHz, acetone-d6): δ = 0.98 (36H, t, J = 7.4 Hz, –CH3), 1.45 (24H, tq, J = 7.4 Hz, 7.4 Hz, –CH2–CH3), 1.73–1.84 (24H, m, –CH2–CH2–), 3.30–3.44 (24H, m, –N–CH2–), 3.81 (1H, d, J = 15.2 Hz, P(O)CHaHb), 3.85 (1H, d, J = 15.2 Hz, P(O)CHcHd), 4.08 (1H, d, J = 15.2 Hz, P(O)CHaHb), 4.13 (1H, d, J = 15.2 Hz, P(O)CHcHd), 7.35 (2H, ddd, J = 8.3, 7.0, 1.1 Hz, CaromH), 7.38–7.48 (4H, m, CaromH), 7.72 (2H, dd, J = 8.3, 2.8 Hz, CaromH), 7.82–7.89 (4H, m, CaromH), 8.28 (2H, d, J = 8.3 Hz, CaromH) ppm. 13C-NMR (CD3CN): δ = 13.9 (s, –CH3), 20.4 (s, –CH2–CH3), 24.4 (s, –CH2–CH2–), 30.3 (d, 1J = 138.3 Hz, P(O)CH2), 59.3 (s, –N–CH2–), 126.0 (s, Carom), 126.5 (d, J = 4.5 Hz, Carom), 126.6 (s, Carom), 128.3 (d, J = 4.5 Hz, Carom), 129.2 (s, Carom), 129.5 (d, J = 10.9 Hz, Carom), 130.0 (d, J = 8.0 Hz, Carom), 133.2 (d, J = 5.8 Hz, Carom), 134.8 (d, J = 2.5 Hz, Carom) ppm. 31P-NMR (acetone-d6): δ = 24.5 ppm.

3.4. X-ray Crystallography

Crystallographic data for compounds TBA-1 and TBA-2 are summarized in Table 2. Intensity data for compound TBA-2 were collected at 100(2) K on an Agilent Technologies SuperNova diffractometer (Agilent Technologies Inc., Santa Clara, CA, USA) equipped with an Eos CCD detector and mirror-monochromated Mo radiation (λ = 0.71073 Å). The data acquisition for compound TBA-1 was carried out at 150(2) K using mirror monochromated Cu radiation (λ = 1.54184 Å) and an Atlas CCDC detector. Data collections, unit cell determinations, intensity data integrations, routine corrections for Lorentz and polarization effects, and analytical absorption corrections with face indexing were performed using the CrysAlis Pro software package (Agilent Technologies UK Ltd., Oxford, UK) [67]. The structures were solved using OLEX2 (OlexSys Ltd in Durham University, Durham, UK) [68] and refined by full-matrix least-squares with SHELXL-97 (University of Göttingen, Göttingen, Germany) [69]. Final geometrical calculations were carried out with PLATON (Utrecht University, Utrecht, The Netherlands) [70] as integrated in WinGX (University of Glasgow, Glasgow, UK) [71]. Bond valence sum (BVS) calculations [55] were performed using the BVSumCalc program (courtesy of Michael H. Dickman). Thermal vibrations were treated anisotropically for atoms contained in the hybrid-POM framework (W, Si, O). All the bond lengths and isotropic thermal ellipsoids for aromatic and alkyl groups were normalized using SADI and SIMU-type restraints from SHELXL due to the considerable structural disorder. For the aromatic systems, some of the 1,3-distances (i.e., distances between two atoms that are both bonded to the same atom, or “angle distances”) were restrained to 2.42(2) Å. Hydrogen atoms of the 1-napththylmethyl groups and the methylenic carbons of the n-butyl chains were placed in calculated positions using standard SHELXL parameters. The H atoms of the –CH3 carbons in the n-butyl chains were added as idealized methyl groups with staggered geometry (AFIX33) for the refinement to converge.
Only five out of the six TBA counterions per two POM clusters determined by TGA, NMR and elemental analyses were located in the Fourier map of TBA-2. A sixth cation was found to be disordered over two positions, but it could not be successfully modeled. The much poorer quality of the crystals of TBA-1 only allowed us to model one of the six TBA cations in the asymmetric unit. Therefore, the final structural models contained large solvent accessible voids accounting for ca. 55% (TBA-1) and 20% (TBA-2) of the unit cell that have been analyzed using the SQUEEZE tool of PLATON. The void of 5898 Å3 in TBA-1 is located at (x, y, z) = (0, 0, 0) and corresponds to 3301 electrons, whereas that of 2209 Å3 in TBA-2 is located at (x, y, z) = (1/2, 1/2, 0) and accounts for 1299 electrons. The calculated void could easily host the undetermined 2 TBA cations in the unit cell of TBA-2, which occupy ca. 1800 Å3 if they are considered as spherical units with a radius of 6 Å. The remaining void space could be occupied by solvent molecules that are rapidly lost upon the exposure of the crystals at ambient conditions. In the case of TBA-1, the rough approximation used for estimating the volume occupied by the undetermined cations is inaccurate for such a large void accounting for more than the 50% of the unit cell. The hkl reflection file generated with SQUEEZE contained 23,423 (TBA-1) and 21,696 (TBA-2) observed reflections out of the 25,376 and 22,761 reflections in the original files. Refinement of the structural model using the squeeze-generated reflection file led to a significant improvement of the final R and wR2 agreement factors in comparison to the original ones [TBA-1: R, 0.138 and wR2 (all), 0.396; TBA-2: R, 0.107 and wR2 (all), 0.247]. CCDC-1471565 (TBA-1) and 1471564 (TBA-2) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.

3.5. Computational Details

All TD-DFT calculations were carried out with ORCA 3.0.3 software [72]. The geometry of the [H(C11H9PO)2(α-SiW11O39)]3− anion, taken from the crystal structure of TBA-2, was optimized using the BP86 [73,74] generalized-gradient-approximation density functional, modified with Grimme’s DFT-D3 atom-pairwise dispersion correction [75] with Becke–Johnson damping [76] to take into account London-dispersion effects. The Ahlrichs’ double-ζ AO basis set with one set of polarization functions (def2-SVP) [77] was used for all atoms. The Stuttgart–Dresden quasi-relativistic effective core potential SD(60,MWB) [78] was applied to W atoms. The UV–Vis absorption spectrum was calculated (at the optimized geometry) using linear-response time-dependent density functional calculations (TD-DFT) within the Tamm–Dancoff approximation [79], with the same basis sets and ECPs as during geometry optimization, and using the B3LYP hybrid density functional [80]. In order to speed up the calculation, the essentially linear RIJCOSX approximation [81,82] has been used, which has been shown to lead to virtually no loss of accuracy [83].

4. Conclusions

The study presented herein represents a solid confirmation of the fact that the classical approach to functionalize lacunary POM clusters with organo-derivatives of the p-block can be applied to new platforms belonging to heteropolyoxotungstates with heteroatoms of the group 13. In this work, an organophosphonate derivative of a Keggin-type borotungstate (TBA-1) has been structurally characterized for the first time. Experiments carried out using the monolacunary Keggin-type silicotungstate have resulted in the hybrid TBA-2 analogue and demonstrate that the functionalization with organophosphonate precursors such as 1-naphthylmethylphosphonic acid can be extended to isoelectronic POM species regardless of the different reactivity conferred by the heteroatoms. The UV–Vis spectroscopy reveals the presence of a charge transfer process promoted by transitions from the naphthyl aromatic system to the Keggin anion as confirmed by TD-DFT calculations. This electronic feature could be at the origin of the quenching that the fluorescent properties of the 1-naphthylmethylphosphonate group undergo upon its incorporation into the inorganic oxo-tungstate skeleton. A combination of multinuclear NMR studies and ESI-MS spectrometry confirms the solution stability of the hybrid frameworks in the organic media. We plan to react the obtained hybrid POM platforms with catalytically active 3d- and 4f-metals centers in the near future with the aim of preparing models for supported site-isolated catalysts with an aromatic environment that will confer photoactive properties on the system and might enhance the selectivity of the catalytic process by interacting with the substrates and providing steric hindrance.

Supplementary Materials

The following supplementary materials are available online at www.mdpi.com/2304-6740/4/2/14/s1, Figure S1: TGA curves for TBA-1 and TBA-2, Table S1: Bond lengths and angles for the P atoms in TBA-1 and TBA-2, Table S2: Geometrical parameters for the intermolecular C–H···O hydrogen bonds involving 1-naphthylmethylphosphonate groups and OPOM atoms in TBA-1 and TBA-2, Figure S2: UV–Vis spectra of acetonitrile solutions of TBA-1 with different concentrations, Figure S3: Fluorescence emission spectra of solid samples of TBA-1, TBA-2 and 1-naphthylmethylphosphonic acid in the 300–450 nm region (λexc = 283 nm), Figure S4: 1H-NMR spectra of TBA-1 and TBA-2 in acetone-d6 compared with that of the commercial 1-naphthylmethylphosphonic acid, Figure S5: 13C-NMR spectra of TBA-1 and TBA-2 in acetone-d6 compared with that of the commercial 1-naphthylmethylphosphonic acid, Figure S6: Negative ESI mass spectrum of a CH3CN solution of TBA-2 (Uc = 15 V) with details of the signals corresponding to the species {2}3− = [H(C11H9PO)2(SiW11O39)]3− and {2 + TBA}2− = {(C16H36N)[H(C11H9PO)2(SiW11O39)]}2− compared with the simulated isotopic patterns.

Acknowledgments

This work was funded by the Spanish Ministerio de Economía y Competitividad (grant MAT2013-48366-C2-2P). The authors thank Cristian Vicent (Serveis Centrals d’Instrumentació Científica, Universitat Jaume I) for the ESI-MS experiments. Technical and human support provided by SGIker (UPV/EHU) is gratefully acknowledged.

Author Contributions

Nerea Andino prepared the title compounds and performed its physicochemical characterization in close collaboration with Beñat Artetxe and Santiago Reinoso; Leire San Felices collected the single-crystal X-ray diffraction data and solved the structures; Pablo Vitoria carried out the theoretical calculations; Santiago Reinoso and Beñat Artetxe prepared the manuscript and acted as scientific coordinators; Jose I. Martínez analyzed the NMR spectra; Fernando López Arbeloa and Beñat Artetxe were in charge of conducting the solution UV–Vis and fluorescence spectra; and Juan M. Gutiérrez-Zorrilla conceived the work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pope, M.T. Heteropoly and Isopoly Oxometalates; Springer: Berlin, Germany, 1983. [Google Scholar]
  2. Polyoxometalates: From Platonic Solids to Anti-Retroviral Activity; Pope, M.T.; Müller, A. (Eds.) Kluwer: Dordrecht, The Netherlands, 1994.
  3. Polyoxometalate Chemistry: Some Recent Trends; Sécheresse, F. (Ed.) World Scientific: Singapore, 2013.
  4. Gouzerh, P.; Proust, A. Main-group element, organic, and organometallic derivatives of polyoxometalates. Chem. Rev. 1998, 98, 77–111. [Google Scholar] [CrossRef] [PubMed]
  5. Long, D.-L.; Tsunashima, R.; Cronin, L. Polyoxometalates: Building blocks for functional nanoscale systems. Angew. Chem. Int. Ed. 2010, 49, 1736–1758. [Google Scholar] [CrossRef] [PubMed]
  6. Proust, A.; Thouvenot, R.; Gouzerh, P. Functionalization of polyoxometalates: Towards advanced applications in catalysis and materials science. Chem. Commun. 2008, 1837–1852. [Google Scholar] [CrossRef] [PubMed]
  7. Song, Y.-F.; Tsunashima, R. Recent advances on polyoxometalate-based molecular and composite materials. Chem. Soc. Rev. 2010, 41, 7384–7402. [Google Scholar] [CrossRef] [PubMed]
  8. Santoni, J.-P.; Hanan, G.S.; Hasenknopf, B. Covalent multi-component systems of polyoxometalates and metal complexes: Toward multi-functional organic–inorganic hybrids in molecular and material sciences. Coord. Chem. Rev. 2014, 281, 64–85. [Google Scholar] [CrossRef]
  9. 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]
  10. Dolbecq, A.; Dumas, E.; Mayer, C.R.; Mialane, P. Hybrid organic–inorganic polyoxometalate compounds: From structural diversity to applications. Chem. Rev. 2010, 110, 6009–6048. [Google Scholar] [CrossRef] [PubMed]
  11. Blazevic, A.; Rompel, A. The Anderson–Evans polyoxometalate: From inorganic building blocks via hybrid organic–inorganic structures to tomorrows “Bio-POM”. Coord. Chem. Rev. 2016, 307, 42–64. [Google Scholar] [CrossRef]
  12. Yin, P.; Wu, P.; Xiao, Z.; Li, D.; Bitterlich, E.; Zhang, J.; Cheng, P.; Vezenov, D.V.; Liu, T.; Wei, Y. A double-tailed fluorescent surfactant with a hexavanadate cluster as the head group. Angew. Chem. Int. Ed. 2011, 50, 2521–2522. [Google Scholar] [CrossRef] [PubMed]
  13. Pradeep, C.P.; Long, D.-L.; Newton, G.N.; Song, Y.-F.; Cronin, L. Supramolecular metal oxides: Programmed hierarchical assembly of a protein-sized 21 kDa [(C16H36N)19{H2NC(CH2O)3P2V3W15O59}4]5− polyoxometalate assembly. Angew. Chem. Int. Ed. 2008, 47, 4388–4391. [Google Scholar] [CrossRef] [PubMed]
  14. Peng, Z. Rational synthesis of covalently bonded organic–inorganic hybrids. Angew. Chem. Int. Ed. 2004, 43, 930–935. [Google Scholar] [CrossRef] [PubMed]
  15. Artetxe, B.; Reinoso, S.; San Felices, L.; Vitoria, P.; Pache, A.; Martín-Caballero, J.; Gutiérrez-Zorrilla, J.M. Functionalization of Krebs-Type polyoxometalates with N,O-chelating ligands: A systematic study. Inorg. Chem. 2015, 54, 241–252. [Google Scholar] [CrossRef] [PubMed]
  16. Rousseau, G.; Oms, O.; Dolbecq, A.; Marrot, J.; Mialane, P. Route for the elaboration of functionalized hybrid 3d-substituted trivacant Keggin anions. Inorg. Chem. 2011, 50, 7376–7378. [Google Scholar] [CrossRef] [PubMed]
  17. Fang, X.; Anderson, T.M.; Hill, C.L. Enantiomerically pure polytungstates: Chirality transfer through zirconium coordination centers to nanosized inorganic clusters. Angew. Chem. Int. Ed. 2005, 44, 3540–3544. [Google Scholar] [CrossRef] [PubMed]
  18. Zheng, S.-T.; Zhang, J.; Yang, G.-Y. Designed synthesis of POM–organic frameworks from {Ni6PW9} building blocks under hydrothermal conditions. Angew. Chem. Int. Ed. 2008, 47, 3909–3913. [Google Scholar] [CrossRef] [PubMed]
  19. Zonnevijlle, F.; Pope, M.T. Attachment of organic groups to heteropoly oxometalate anions. J. Am. Chem. Soc. 1979, 101, 2731–2732. [Google Scholar] [CrossRef]
  20. Kortz, U.; Hussain, F.; Reicke, M. The ball-shaped heteropolytungstates [{Sn(CH3)2(H2O)}24{Sn(CH3)2}12(A-XW9O34)12]36−. Angew. Chem. Int. Ed. 2005, 44, 3773–3777. [Google Scholar] [CrossRef] [PubMed]
  21. Boglio, C.; Micoine, K.; Derat, E.; Thouvenot, R.; Hasenknopf, B.; Thorimbert, S.; Lacôte, E.; Malacria, M. Regioselective activation of oxo ligands in functionalized Dawson polyoxotungstates. J. Am. Chem. Soc. 2008, 130, 4553–4561. [Google Scholar] [CrossRef] [PubMed]
  22. Nomiya, K.; Togashi, Y.; Kasahara, Y.; Aoki, S.; Seki, H.; Noguchi, M.; Yoshida, S. Synthesis and structure of Dawson polyoxometalate-based, multifunctional, inorganic–organic hybrid compounds: Organogermyl complexes with one terminal functional group and organosilyl analogues with two terminal functional groups. Inorg. Chem. 2011, 50, 9606–9619. [Google Scholar] [CrossRef] [PubMed]
  23. Piedra-Garza, L.F.; Dickman, M.H.; Moldovan, O.; Breuning, H.J.; Kortz, U. Organoantimony-containing polyoxometalate: [{PhSbOH}3(A-α-PW9O34)2]9−. Inorg. Chem. 2009, 48, 411–413. [Google Scholar] [CrossRef] [PubMed]
  24. Knoth, W.H. Derivatives of heteropolyanions. 1. Organic derivatives of W12SiO404−, W12PO403−, and Mo12SiO404−. J. Am. Chem. Soc. 1979, 101, 759–760. [Google Scholar] [CrossRef]
  25. Judeinstein, P.; Deprunb, C.; Nadjo, L. Synthesis and multispectroscopic characterization of organically modified polyoxometalates. Dalton Trans. 1991, 1991–1997. [Google Scholar] [CrossRef]
  26. 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] [PubMed]
  27. Mazeaud, A.; Ammari, N.; Robert, F.; Thouvenot, R. Coordination chemistry of polyoxometalates: Rational synthesis of the mixed organosilyl derivatives of trivacant polyoxotungstates α-A-[PW9O34(tBuSiO)3(RSi)]3− and α-B-[AsW9O33(tBuSiO)3(HSi)]3−. Angew. Chem. Int. Ed. Engl. 1996, 35, 1961–1964. [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/Ionization Time-of-Flight Mass Spectrometry. Chem. Eur. J. 2000, 6, 105–110. [Google Scholar] [CrossRef]
  29. Kim, G.-S.; Hagen, K.S.; Hill, C.L. Synthesis, structure, spectroscopic properties, and hydrolytic chemistry of organophosphonoyl polyoxotungstates of formula [C6H5P(O)]2Xn+WllO39](8−n)− (X = P5+, Si4+)]. Inorg. Chem. 1992, 31, 5316–5324. [Google Scholar] [CrossRef]
  30. Mayer, C.R.; Thouvenot, R. Organophosphoryl derivatives of trivacant tungstophosphates of general formula α-A-[PW9O34(RPO)2]5−: Synthesis and structure determination by Multinuclear Magnetic Resonance Spectroscopy.(31P, 183W). Dalton Trans. 1998, 7–13. [Google Scholar] [CrossRef]
  31. Mayer, C.R.; Herson, P.; Thouvenot, R. Organic–inorganic hybrids based on polyoxometalates. Synthesis and structural characterization of bis(organophosphoryl) decatungstosilicates [γ-SiW10O36(RPO)2]4−. Inorg. Chem. 1999, 38, 6152–6158. [Google Scholar] [CrossRef] [PubMed]
  32. Sun, Z.-G.; Liu, Q.; Liu, J.-F. Synthesis and spectroscopic characterization of organophosphoryl polyoxotungstates [C6H11P(O)]2Xn+ W11O39(8−n)− (Xn+ = P5+, Si4+, B3+, Ga3+). Polyhedron 2000, 19, 125–128. [Google Scholar] [CrossRef]
  33. Boujtita, 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]
  34. Sun, Z.-G.; Zhu, Z.-M.; You, W.-S. Synthesis and characterization of organophosphoryl polyoxotungstates [RP(O)]2Xn+W11O39](8−n)− (Xn+ = P5+, Si4+, Ge4+, Ga3+). Transition Met. Chem. 2003, 28, 849–851. [Google Scholar] [CrossRef]
  35. Schroden, R.C.; Blanford, C.F.; Melde, B.J.; Johnson, B.J.S.; Stein, A. Direct synthesis of ordered macroporous silica materials functionalized with polyoxometalate clusters. Chem. Mater. 2001, 13, 1074–1081. [Google Scholar] [CrossRef]
  36. Mayer, C.R.; Cabuil, V.; Lalot, T.; Thouvenot, R. Incorporation of magnetic nanoparticles in new hybrid networks based on heteropolyanions and polyacrylamide. Angew. Chem. Int. Ed. 1999, 38, 3672–3675. [Google Scholar] [CrossRef]
  37. Mayer, C.R.; Neveu, S.; Cabuil, V. A Nanoscale hybrid system based on gold nanoparticles and heteropolyanions. Angew. Chem. Int. Ed. 2002, 41, 501–503. [Google Scholar] [CrossRef]
  38. Cannizzo, C.; Mayer, C.R.; Sécheresse, F.; Larpent, C. Covalent hybrid materials based on nanolatex particles and Dawson polyoxometalates. Adv. Mater. 2005, 17, 2888–2892. [Google Scholar] [CrossRef]
  39. Villanneau, R.; Marzouk, A.; Wang, Y.; Ben Djamaa, A.; Laugel, G.; Proust, A.; Launay, F. Covalent grafting of organic–inorganic polyoxometalates hybrids onto mesoporous SBA-15: A key step for new anchored homogeneous catalysts. Inorg. Chem. 2013, 52, 2958–2965. [Google Scholar] [CrossRef] [PubMed]
  40. Joo, N.; Renaudineau, S.; Delapierre, G.; Bidan, G.; Chamoreau, L.-S.; Thouvenot, R.; Gouzerh, P.; Proust, A. Organosilyl/-germyl polyoxotungstate hybrids for covalent grafting onto silicon surfaces: Towards molecular memories. Chem. Eur. J. 2010, 16, 5043–5051. [Google Scholar] [CrossRef] [PubMed]
  41. Mercier, D.; Boujday, S.; Annabi, C.; Villanneau, R.; Pradier, C.-M.; Proust, A. Bifunctional polyoxometalates for planar gold surface nanostructuration and protein immobilization. J. Phys. Chem. C 2012, 116, 13217–13224. [Google Scholar] [CrossRef]
  42. Duffort, V.; Thouvenot, R.; Afonso, C.; Izzet, G.; Proust, A. Straightforward synthesis of new polyoxometalate-based hybrids exemplified by the covalent bonding of a polypyridyl ligand. Chem. Commun. 2009, 6062–6064. [Google Scholar] [CrossRef] [PubMed]
  43. Elliott, K.J.; Harriman, A.; le Pleux, L.; Pellegrin, Y.; Blart, E.; Mayer, C.R.; Odobel, F.A. Porphyrin–polyoxometallate bio-inspired mimic for artificial photosynthesis. Phys. Chem. Chem. Phys. 2009, 11, 8767–8773. [Google Scholar] [CrossRef] [PubMed]
  44. Matt, B.; Renaudineau, S.; Chamoreau, L.-M.; Afonso, C.; Izzet, G.; Proust, A. Hybrid polyoxometalates: Keggin and Dawson silyl derivatives as versatile platforms. J. Org. Chem. 2011, 76, 3107–3111. [Google Scholar] [CrossRef] [PubMed]
  45. Odobel, F.; Severac, M.; Pellegrin, Y.; Blart, E.; Fosse, C.; Cannizzo, C.; Mayer, C.R.; Eliott, K.J.; Harriman, A. Coupled sensitizer–catalyst dyads: Electron-transfer reactions in a perylene–polyoxometalate conjugate. Chem. Eur. J. 2009, 15, 3130–3138. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, L.; Hu, L.; Liu, Q.; Du, Z.-L.; Li, F.-B.; Li, G.-H.; Zhu, X.-J.; Wong, W.-Y.; Wang, L.; Li, H. Synthesis and photoelectric properties of new Dawson-type polyoxometalate-based dimeric and oligomeric Pt(II)-acetylide inorganic–organic hybrids. Dalton Trans. 2015, 44, 306–315. [Google Scholar] [CrossRef] [PubMed]
  47. Villanneau, R.; Ben Djamâa, A.; Chamoreau, L.-M.; Gontard, G.; Proust, A. Bisorganophosphonyl and -organoarsenyl derivatives of heteropolytungstates as hard ligands for early-transition-metal and lanthanide cations. Eur. J. Inorg. Chem. 2013, 1815–1820. [Google Scholar] [CrossRef]
  48. Villanneau, R.; Racimor, D.; Messner-Henning, E.; Rousselière, H.; Picart, S.; Thouvenot, R.; Proust, A. Insights into the coordination chemistry of phosphonate derivatives of heteropolyoxotungstates. Inorg. Chem. 2011, 50, 1164–1166. [Google Scholar] [CrossRef] [PubMed]
  49. Carraro, M.; Sandei, L.; Sartorel, A.; Scorrano, G.; Bonchio, M. Hybrid polyoxotungstates as second-generation POM-based catalysts for microwave-assisted H2O2 activation. Org. Lett. 2006, 8, 3671–3674. [Google Scholar] [CrossRef] [PubMed]
  50. Guillemot, G.; Matricardi, E.; Chamoreau, L.-M.; Thouvenot, R.; Proust, A. Oxidovanadium(V) anchored to silanol-functionalized polyoxotungstates: Molecular models for single-site silica-supported vanadium catalysts. ACS Catal. 2015, 5, 7415–7423. [Google Scholar] [CrossRef]
  51. Tézé, A.; Michelon, M.; Hervé, G. Syntheses and structures of the tungstoborate anions. Inorg. Chem. 1997, 36, 505–509. [Google Scholar] [CrossRef]
  52. Nsouli, N.H.; Chubarova, E.V.; Al-Oweini, R.; Bassil, B.S.; Sadakane, M.; Kortz, U. Organoruthenium-containing heteropoly-23-tungstate family [{Ru(L)}2(α-XW11O39)2WO2]m (L = benzene, p-cymene; X = GeIV, SiIV, m = 10; BIII, m = 12). Eur. J. Inorg. Chem. 2013, 1742–1747. [Google Scholar] [CrossRef]
  53. Reinoso, S.; Dickman, M.H.; Matei, M.F.; Kortz, U. 13-Tungstoborate stabilized by an organostannoxane hexamer. Inorg. Chem. 2007, 46, 4383–4385. [Google Scholar] [CrossRef] [PubMed]
  54. Reinoso, S.; Vitoria, P.; San Felices, L.; Gutiérrez-Zorrilla, J.M. Diiron(III)-containing 23-tungsto-2-borate: First evidence of 3d metal substitution in the {BW13} framework. Eur. J. Inorg. Chem. 2013, 1644–1648. [Google Scholar] [CrossRef]
  55. Brown, I.D.; Alternatt, D. Bond-valence parameters obtained from a systematic analysis of the inorganic crystal structure database. Acta Crystallogr. 1985, B41, 244–247. [Google Scholar] [CrossRef]
  56. Dougherty, D.A. The cation-π interaction. Acc. Chem. Res. 2013, 46, 885–893. [Google Scholar] [CrossRef] [PubMed]
  57. Nishio, M.; Umezawa, Y.; Fantini, J.; Weiss, M.S.; Chakrabarti, P. CH−π hydrogen bonds in biological macromolecules. Phys. Chem. Chem. Phys. 2014, 16, 12648–12683. [Google Scholar] [CrossRef] [PubMed]
  58. Gallivan, J.P.; Dougherty, D.A. A computational study of cation–π interactions vs. salt bridges in aqueous media: Implications for protein engineering. J. Am. Chem. Soc. 2000, 122, 870–874. [Google Scholar] [CrossRef]
  59. Goto, Y.; Kamata, K.; Yamaguchi, K.; Uehara, K.; Hikichi, S.; Mizuno, N. Synthesis, structural characterization, and catalytic performance of dititanium-substituted γ-Keggin silicotungstate. Inorg. Chem. 2006, 45, 2347–2356. [Google Scholar] [CrossRef] [PubMed]
  60. Maeda, H.; Maeda, T.; Mizuno, K. Absorption and fluorescence spectroscopic properties of 1- and 1,4-silyl-substituted naphthalene derivatives. Molecules 2012, 17, 5108–5125. [Google Scholar] [CrossRef] [PubMed]
  61. Xie, Y.; Zhang, Q.; Zhao, Z.; Wu, X.; Chen, S.; Lu, C. New optical supramolecular compound constructed from a polyoxometalate cluster and an organic substrate. Inorg. Chem. 2008, 47, 8086–8090. [Google Scholar] [CrossRef] [PubMed]
  62. Matt, B.; Coudret, C.; Viala, C.; Jouvenot, D.; Loiseau, F.; Izzet, G.; Proust, A. Elaboration of covalently linked polyoxometalates with ruthenium and pyrene chromophores and characterization of their photophysical properties. Inorg. Chem. 2011, 50, 7761–7768. [Google Scholar] [CrossRef] [PubMed]
  63. Sun, Z.-G.; You, W.-S.; Li, J.; Liu, J.-F. Synthesis and spectroscopic characterization of organophosphoryl polyoxotungstates α-[{RP(O)}2SiW11O39]4−. Inorg. Chem. Commun. 2003, 6, 238–240. [Google Scholar] [CrossRef]
  64. Pretsch, E.; Clerc, T.; Seibl, J.; Simon, W. Tables of Spectral Data for Structure Determination of Organic Compounds; Springer-Verlag: Berlin, Germany, 1989. [Google Scholar]
  65. Hervé, G.; Tézé, A. Study of α and β-enneatungstosilicates and germanates. Inorg. Chem. 1977, 16, 2115–2117. [Google Scholar] [CrossRef]
  66. Fulmer, G.R.; Miller, A.J. M.; Sherden, N.H.; Gottlieb, H.E.; Nudelman, A.; Stoltz, B.M.; Bercaw, J.E.; Goldberg, K.I. NMR chemical shifts of trace impurities: Common laboratory solvents, organics, and gases in deuterated solvents relevant to the organometallic chemist. Organometallics 2010, 29, 2176–2179. [Google Scholar] [CrossRef]
  67. CrysAlisPro Software System; Version 171.36.24; Agilent Technologies UK Ltd.: Oxford, UK, 2012.
  68. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  69. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. 2008, A64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  70. Spek, A.L. Structure validation in chemical crystallography. Acta Crystallogr. 2009, D65, 148–155. [Google Scholar] [CrossRef] [PubMed]
  71. Farrugia, L.J. WinGX suite for small-molecule single-crystal crystallography. J. Appl. Crystallogr. 1999, 32, 837–836. [Google Scholar] [CrossRef]
  72. Neese, F. The ORCA program system. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2012, 73–78. [Google Scholar] [CrossRef]
  73. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic-behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  74. Perdew, J.P. Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B 1986, 33, 8822–8824. [Google Scholar] [CrossRef]
  75. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 elements H–Pu. J. Chem. Phys. 2010, 132, 154014. [Google Scholar] [CrossRef] [PubMed]
  76. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef] [PubMed]
  77. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design an assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef] [PubMed]
  78. Andrae, D.; Haeussermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Energy-adjusted ab initio pseudopotentials for the second and third row transition elements. Theor. Chim. Acta 1990, 77, 123–141. [Google Scholar] [CrossRef]
  79. Hirata, S.; Head-Gordon, M. Time-dependent density functional theory within the Tamm-Dancoff approximation. Chem. Phys. Lett. 1999, 314, 291–299. [Google Scholar] [CrossRef]
  80. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  81. Neese, F.; Wennmohs, F.; Hansen, A.; Becker, U. Efficient, approximate and parallel Hartree–Fock and hybrid DFT calculations. A ‘Chain-Of-Spheres’ algorithm for the Hartree–Fock exchange. Chem. Phys. 2009, 356, 98–109. [Google Scholar] [CrossRef]
  82. Izsak, R.; Neese, F. An overlap fitted chain of spheres exchange method. J. Chem. Phys. 2011, 135, 144105. [Google Scholar] [CrossRef] [PubMed]
  83. Petrenko, T.; Kossmann, S.; Neese, F. Efficient time-dependent density functional theory approximations for hybrid density functionals: Analytical gradients and parallelization. J. Chem. Phys. 2011, 134, 054116. [Google Scholar] [CrossRef] [PubMed]
Figure 1. FT-IR spectra of TBA-1 (A) and TBA-2 (B) compared with those of their corresponding precursors K8[HBW11O39]·13H2O and TBA8[SiW11O39].
Figure 1. FT-IR spectra of TBA-1 (A) and TBA-2 (B) compared with those of their corresponding precursors K8[HBW11O39]·13H2O and TBA8[SiW11O39].
Inorganics 04 00014 g001
Figure 2. Polyhedral representation of the molecular structure of POMs 1 and 2 (A); The connectivity of the phosphonate groups and the relative arrangement of the naphthyl residues are highlighted in (B) and (C), respectively; Ball-and-stick representation of the hybrid POMs with intramolecular C–H···OPOM hydrogen bonds depicted in red dashed lines (D); Detail of the TBA cation embraced between the two aromatic systems via C–H···π interactions depicted as green dashed lines (E). Color code: WO6, gray octahedra (W, gray); XO4, green tetrahedra (X, green); P, violet; O, red; C, black; H, pink; N, cyan. Aromatic and n-butyl H atoms are omitted for clarity.
Figure 2. Polyhedral representation of the molecular structure of POMs 1 and 2 (A); The connectivity of the phosphonate groups and the relative arrangement of the naphthyl residues are highlighted in (B) and (C), respectively; Ball-and-stick representation of the hybrid POMs with intramolecular C–H···OPOM hydrogen bonds depicted in red dashed lines (D); Detail of the TBA cation embraced between the two aromatic systems via C–H···π interactions depicted as green dashed lines (E). Color code: WO6, gray octahedra (W, gray); XO4, green tetrahedra (X, green); P, violet; O, red; C, black; H, pink; N, cyan. Aromatic and n-butyl H atoms are omitted for clarity.
Inorganics 04 00014 g002
Figure 3. View of the crystal packing of compounds TBA-1 and TBA-2 along the [100] direction (A); Connectivity between POMs through intermolecular C–H···OPOM hydrogen bonds (depicted as red lines) within bilayers (B); Projection of a layer along the [001] direction where the N3 and N5 TBA cations connect POMs of the same z level in a zigzag mode (C). Color code: same as in Figure 2, crystallographically independent POM clusters are depicted in blue and gray. Butyl groups from TBA cations and aromatic H atoms are omitted for clarity.
Figure 3. View of the crystal packing of compounds TBA-1 and TBA-2 along the [100] direction (A); Connectivity between POMs through intermolecular C–H···OPOM hydrogen bonds (depicted as red lines) within bilayers (B); Projection of a layer along the [001] direction where the N3 and N5 TBA cations connect POMs of the same z level in a zigzag mode (C). Color code: same as in Figure 2, crystallographically independent POM clusters are depicted in blue and gray. Butyl groups from TBA cations and aromatic H atoms are omitted for clarity.
Inorganics 04 00014 g003
Figure 4. UV–Vis spectra of acetonitrile solutions of TBA-2 with different concentrations compared with those of the commercial 1-naphthylmethylphosphonic acid and the TBA8[SiW11O39] precursor. The simulated spectrum from theoretical TD-DFT calculations and the first 100 transitions with calculated relative intensity are also depicted.
Figure 4. UV–Vis spectra of acetonitrile solutions of TBA-2 with different concentrations compared with those of the commercial 1-naphthylmethylphosphonic acid and the TBA8[SiW11O39] precursor. The simulated spectrum from theoretical TD-DFT calculations and the first 100 transitions with calculated relative intensity are also depicted.
Inorganics 04 00014 g004
Figure 5. Photographs of the crystalline batches of TBA-1 and TBA-2 upon being left to dry in air overnight.
Figure 5. Photographs of the crystalline batches of TBA-1 and TBA-2 upon being left to dry in air overnight.
Inorganics 04 00014 g005
Figure 6. Difference densities for several CT transitions calculated for the hybrid POM 2 at the TDA-B3LYP/def2-SVP//BP86-D3/def2-SVP level of theory. The electronic transfer is from the red to the blue surfaces: (A) 408 nm; (B) 380 nm; (C) 350 nm; (D) 290 nm and (E) 290 nm.
Figure 6. Difference densities for several CT transitions calculated for the hybrid POM 2 at the TDA-B3LYP/def2-SVP//BP86-D3/def2-SVP level of theory. The electronic transfer is from the red to the blue surfaces: (A) 408 nm; (B) 380 nm; (C) 350 nm; (D) 290 nm and (E) 290 nm.
Inorganics 04 00014 g006
Figure 7. Diffuse Reflectance UV–Vis spectra of powdered crystalline samples of TBA-1 and TBA-2 compared with those of the K8[HBW11O39]·13H2O and K8[SiW11O39]·13H2O precursors.
Figure 7. Diffuse Reflectance UV–Vis spectra of powdered crystalline samples of TBA-1 and TBA-2 compared with those of the K8[HBW11O39]·13H2O and K8[SiW11O39]·13H2O precursors.
Inorganics 04 00014 g007
Figure 8. Fluorescence emission spectra of 10−6 M acetonitrile solutions of TBA-1, TBA-2 and 1-naphthylmethylphosphonic acid in the 300–450 nm region (λexc = 283 nm).
Figure 8. Fluorescence emission spectra of 10−6 M acetonitrile solutions of TBA-1, TBA-2 and 1-naphthylmethylphosphonic acid in the 300–450 nm region (λexc = 283 nm).
Inorganics 04 00014 g008
Figure 9. 31P-NMR spectra of TBA-1 and TBA-2 in acetone-d6 compared with that of the commercial 1-naphthylmethylphosphonic acid.
Figure 9. 31P-NMR spectra of TBA-1 and TBA-2 in acetone-d6 compared with that of the commercial 1-naphthylmethylphosphonic acid.
Inorganics 04 00014 g009
Figure 10. Expanded regions of the naphthyl (A) and methylenic (B) protons in the 1H-NMR spectra of TBA-1 and TBA-2 recorded in acetone-d6 compared with those of the commercial 1-naphthylmethylphosphonic acid. The diastereotopic spin system for the methylenic group in the hybrid POMs 1 and 2 is schematically depicted. The signal assigned to the B–O–H proton is highlighted in red.
Figure 10. Expanded regions of the naphthyl (A) and methylenic (B) protons in the 1H-NMR spectra of TBA-1 and TBA-2 recorded in acetone-d6 compared with those of the commercial 1-naphthylmethylphosphonic acid. The diastereotopic spin system for the methylenic group in the hybrid POMs 1 and 2 is schematically depicted. The signal assigned to the B–O–H proton is highlighted in red.
Inorganics 04 00014 g010
Figure 11. 13C-NMR spectra of TBA-1 and TBA-2 in CD3CN. The signals labeled as * correspond to the solvent.
Figure 11. 13C-NMR spectra of TBA-1 and TBA-2 in CD3CN. The signals labeled as * correspond to the solvent.
Inorganics 04 00014 g011
Figure 12. Negative ESI mass spectrum of a CH3CN solution of TBA-1 (Uc = 15 V) and comparison of the signals corresponding to the species {1}3− = [H(C11H9PO)2(HBW11O39)]3− and {1 + TBA}2− = {(C16H36N)[H(C11H9PO)2(HBW11O39)]}2− with the simulated isotopic patterns.
Figure 12. Negative ESI mass spectrum of a CH3CN solution of TBA-1 (Uc = 15 V) and comparison of the signals corresponding to the species {1}3− = [H(C11H9PO)2(HBW11O39)]3− and {1 + TBA}2− = {(C16H36N)[H(C11H9PO)2(HBW11O39)]}2− with the simulated isotopic patterns.
Inorganics 04 00014 g012
Table 1. Geometrical parameters (Å, °) for the intramolecular C–H···O hydrogen bonds involving the methylenic C atoms of the 1-naphthylmethylphosphonate groups and the OPOM atoms in compounds TBA-1 and TBA-2.
Table 1. Geometrical parameters (Å, °) for the intramolecular C–H···O hydrogen bonds involving the methylenic C atoms of the 1-naphthylmethylphosphonate groups and the OPOM atoms in compounds TBA-1 and TBA-2.
D–H···AH···AD···A<D–H···A>
TBA-1
C1–H1B···O212.853.47(3)121
C1–H1B···O282.713.32(3)120
C21–H21A···O252.593.09(3)111
C21–H21A···O312.633.29(3)124
C201–H20B···O2212.863.40(3)115
C201–H20B···O2282.813.47(3)125
C221–H22A···O2252.613.27(3)123
C221–H22A···O2312.633.11(3)111
TBA-2
C1–H1B···O212.463.05(5)111
C1–H1B···O282.553.15(5)127
C21–H21A···O252.773.46(4)127
C21–H21A···O312.893.44(4)115
C201–H20B···O2212.863.52(4)124
C201–H20B···O2282.943.47(4)115
C221–H22A···O2252.523.04(5)113
C221–H22A···O2312.483.18(5)127
Table 2. Crystallographic data for TBA-1 and TBA-2.
Table 2. Crystallographic data for TBA-1 and TBA-2.
ParametersTBA-1TBA-2
FormulaC70H128BN3O41P2W11C70H127N3O41P2SiW11
FW (g·mol−1)3762.83779.1
Crystal SystemTriclinicTriclinic
Space GroupP−1P−1
a (Å)17.4020(5)16.9676(5)
b (Å)24.2390(3)24.3022(3)
c (Å)28.6175(5)28.5333(7)
α (˚)70.095(2)70.030(2)
β (˚)72.336(2)72.703(2)
γ (˚)89.976(2)89.957(2)
V3)10743.3(4)10491.3(4)
Z44
ρcalcd (g·cm−3)2.3262.393
μ (mm−1)22.02212.124
T (K)150(2)100(2)
λ (Å)1.54184 (Cu )0.71073 (Mo )
Collected Reflections8033465644
Unique Reflections (Rint)38193 (0.126)36457 (0.048)
Observed Reflections [I > 2σ(I)]2342321696
Parameters/Restraints767/2501459/546
R(F)a [I > 2σ(I)]0.0950.083
wR(F2)a [all data]0.2700.216
GoF0.9641.078
a R(F) = Σ||FoFc||/Σ|Fo|; wR(F2) = {Σ[w(Fo2Fc2)2]/Σ[w(Fo2)2]}1/2.

Share and Cite

MDPI and ACS Style

Andino, N.; Artetxe, B.; Reinoso, S.; Vitoria, P.; San Felices, L.; Martínez, J.I.; López Arbeloa, F.; Gutiérrez-Zorrilla, J.M. Naphthyl-Containing Organophosphonate Derivatives of Keggin-Type Polyoxotungstates. Inorganics 2016, 4, 14. https://doi.org/10.3390/inorganics4020014

AMA Style

Andino N, Artetxe B, Reinoso S, Vitoria P, San Felices L, Martínez JI, López Arbeloa F, Gutiérrez-Zorrilla JM. Naphthyl-Containing Organophosphonate Derivatives of Keggin-Type Polyoxotungstates. Inorganics. 2016; 4(2):14. https://doi.org/10.3390/inorganics4020014

Chicago/Turabian Style

Andino, Nerea, Beñat Artetxe, Santiago Reinoso, Pablo Vitoria, Leire San Felices, Jose I. Martínez, Fernando López Arbeloa, and Juan M. Gutiérrez-Zorrilla. 2016. "Naphthyl-Containing Organophosphonate Derivatives of Keggin-Type Polyoxotungstates" Inorganics 4, no. 2: 14. https://doi.org/10.3390/inorganics4020014

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop