Next Article in Journal
Nonlocal Gravity: Fundamental Tetrads and Constitutive Relations
Next Article in Special Issue
Molecular Blueprinting by Word Processing
Previous Article in Journal
A Chaotic Genetic Algorithm with Variable Neighborhood Search for Solving Time-Dependent Green VRPTW with Fuzzy Demand
Previous Article in Special Issue
Conformer Selection by Electrostatic Hexapoles: A Theoretical Study on 1-Chloroethanol and 2-Chloroethanol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Neutral Heteroleptic Molybdenum Cluster trans-[{Mo6I8}(py)2I4]

by
Margarita V. Marchuk
1,
Yuri A. Vorotnikov
1,
Anton A. Ivanov
1,
Ilia V. Eltsov
2,
Natalia V. Kuratieva
1 and
Michael A. Shestopalov
1,*
1
Nikolaev Institute of Inorganic Chemistry of Siberian Branch of Russian Academy of Sciences, 3 Acad, Lavrentiev Ave., 630090 Novosibirsk, Russia
2
Department of Natural Sciences, Novosibirsk State University, 1 Pirogova St., 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Symmetry 2022, 14(10), 2117; https://doi.org/10.3390/sym14102117
Submission received: 15 September 2022 / Revised: 1 October 2022 / Accepted: 10 October 2022 / Published: 12 October 2022

Abstract

:
Despite that the chemistry of octahedral cluster complexes has been actively developed recently, there are still a lot of unexplored areas. For example, to date, only a few halide M6-clusters with N-heterocycles are known. Here, we obtained an apically heteroleptic octahedral iodide molybdenum cluster complex with pyridine ligands—trans-[{Mo6I8}(py)2I4] by the direct substitution of iodide apical ligands of [{Mo6I8}I6]2– in a pyridine solution. The compound co-crystalized with a monosubstituted form [{Mo6I8}(py)I5] in the ratio of 1:4, and thus, can be described by the formula (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py. The composition was studied using XRPD, elemental analyses, and 1H-NMR and IR spectroscopies. According to the absorption and luminescence data, the partial substitution of apical ligands weakly affects optical properties.

Graphical Abstract

1. Introduction

Octahedral halide M6-clusters of the general formula [{M6X8}L6]n (M = Mo, W, X = halogen, L—ligand of any nature) are unique compounds possessing a number of valuable practically oriented properties—this is the main point to consider when regarding compounds as promising components of different multifunctional materials. Among these properties are (i) bright luminescence in red/near infrared spectral regions (with microsecond lifetimes and high quantum yields) under the excitation of ultraviolet/visible light (up to 550 nm) or X-rays [1,2,3,4,5,6,7,8]; (ii) the efficient photosensitization of the reactive oxygen species generation process (mainly singlet oxygen, 1O2) [4,8,9,10]; (iii) high radiopacity due to a high local concentration of heavy atoms in a cluster core [7,11]. Thus, different groups have demonstrated the utilization of cluster complexes individually or as active components of the materials in numerous areas, such as catalytic systems [12,13,14,15,16], liquid crystals [17,18], oxygen sensors [19,20,21,22], solar concentrators [16,23,24], an active layer of solar cells [25,26], bioimaging, photodynamic therapy (PDT) [27,28,29,30,31,32,33], as components of self-sterilizing materials [34,35,36,37], or as radiocontrast agents [11]. The abundance of publications on the practical application of the clusters to some extent confirms their high potential. However, terra incognita areas are still presented in cluster chemistry, especially in the case of molybdenum and tungsten halide clusters. One of the almost unknown parts correspond to Mo6 and W6 clusters with nitrogen heterocycles. N-heterocycles are a huge class of organic compounds that have great structural diversity, functionalization potential, and possess various biological activities [38,39]. To date, only a few examples of such clusters with N-heterocyclic ligands are reported and only three of them are structurally characterized [40,41]. On the contrary, a fairly large number of complexes with coordinated heterocycles have been obtained in the case of rhenium clusters [42,43], indicating that N-heterocyclic ligands can preserve their useful properties upon coordination to metal centers. For example, clusters [{Re6Q8}(btrz)6]4– (Q = S, Se; btrz is benzotriazolate) similar to pure btrzH are able to bind directly to DNA [42].
Herein, we present the preparation of the new iodide Mo6-cluster with N-heterocyclic ligand, namely, pyridine—trans-[{Mo6I8}(py)2I4]. The compound composition was confirmed using SCXRD (single crystal X-ray diffraction), XRPD (X-ray powder diffraction), elemental analyses, and 1H-NMR and IR spectroscopies. Optical properties, such as absorption and luminescence, were studied in a solid state and in a solution.

2. Materials and Methods

All reagents and solvents employed were commercially available and used as received without further purification. (Bu4N)2[{Mo6I8}I6] was prepared by a metathesis reaction of Cs2[{Mo6I8}I6] and Bu4NI in acetone [44]. Cs2[{Mo6I8}I6] was synthesized from Mo, I2, and CsI by a high-temperature reaction according to the procedure described in [44].
Elemental analyses were obtained using a EuroVector EA3000 Elemental Analyser (S.p.A., Milan, Italy). FTIR spectra in the range of 4000–400 cm−1 for solid (Bu4N)2[{Mo6I8}I6] and (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py as the KBr disk or for liquid pyridine placed between the two aperture plates (KBr) were recorded on a Scimitar FTS 2000 (Digilab LLC, Canton, MA, USA). An energy-dispersive X-ray spectroscopy (EDS) was performed on a Hitachi TM3000 TableTop SEM (Hitachi High-Technologies Corporation, Tokyo, Japan) with Bruker QUANTAX 70 EDS equipment. X-ray powder diffraction (XRPD) patterns were recorded on a Shimadzu XRD 7000S diffractometer (Shimadzu, Kyoto, Japan) (Cu Kα radiation, graphite monochromator, and silicon plate as an external standard). The 1H-NMR spectra were recorded from a saturated DMSO-d6 solution at room temperature on a Bruker Avance III 500 FT-spectrometer (Bruker BioSpin AG, Faellanden, Switzerland) with working frequencies of 499.93 MHz. The chemical shifts were reported in ppm of the δ scale and referred to the signal of residual protons of solvent: δ = 2.5 ppm. Thermal properties were studied on a Thermo Microbalance TG 209 F3 Tarsus (NETZSCH, Selb, Germany) from 30 to 650 °C, at a rate of 10 °C min−1 in He flow (30 mL min−1). Mass spectrometric (MS) detection was performed with a direct injection of liquid samples via an automatic syringe pump KDS 100 (KD scientific Inc., Holliston, MA, USA) at a 180 μL h–1 rate by an electrospray ionization quadrupole time-of-flight (ESI-q-TOF) high resolution mass spectrometer Maxis 4G (Bruker Daltonics, Bremen, Germany). Mass spectra were recorded in the negative mode within the 700–4500 m/z range. The MS calibration was performed externally using an ESI-L calibration mix (Agilent Technologies, Santa Clara, CA, USA); the typical resolution was ca. 50,000, and the accuracy was <1 ppm. Diffuse reflectance spectra were recorded using a UV-Vis-NIR 3101 PC spectrophotometer (Shimadzu Corporation, Kyoto, Japan). The excitation (monitored at λem = 700 nm) and emission (excitation wavelength was 350 nm) spectra were recorded by Cary Eclipse Fluorescence Spectrophotometer (Agilent Technologies, Santa Clara, CA, USA) for compounds in the solid state and in pyridine solutions. The absorbance of pyridine solutions was set at <0.1 at 355 nm.

2.1. Synthesis

(Bu4N)2[{Mo6I8}I6] (300 mg, 0.105 mmol) was dissolved in 3 mL of pyridine (99.5%, 0.037 mol). The resulting mixture was heated in a sealed glass tube at 110 °C for 48 h. During the reaction, a crystalline powder was formed on the walls of the tube. The powder was then separated from the solution, washed 3 times with acetone and air dried. Yield (calcd for [{Mo6I8}(py)2I4]): 56 mg (23%). EDS: Mo/I = 6:13.2. FTIR (KBr, cm−1): νs(CH) = 2953, 2920, 2849, 1217, δs(CC) = 1599, 1483, 1439, 1356, δ(CH) = 1150, 1065, 1007, δs(CH) = 752, 696, 631. 1H NMR ([{Mo6I8}(py)I5], 500 MHz, DMSO) δ 7.54 (t, 3,5-Py), 8.12 (t, 4-Py), 9.38 (d, J = 5.4 Hz, 2,6-Py), 1H NMR (trans-[{Mo6I8}(py)2I4], 500 MHz, DMSO) δ 7.53 (t, 3,5-Py), 8.12 (t, 4-Py), 9.38 (d, J = 5.3 Hz, 2,6-Py). A single crystal suitable for X-ray structural analysis was grown by the slow cooling of the tube during 24 h.

2.2. Crystallography

Single-crystal X-ray diffraction data were collected using a Bruker Nonius X8 Apex 4K CCD (Bruker Corporation, Billerica, MA, USA) diffractometer with graphite monochromatized MoKα radiation (λ = 0.71073 Å). Absorption corrections were made empirically using the SADABS program [45]. The structures were solved by the direct method and further refined by the full-matrix least-squares method using the SHELXTL program package [45,46]. All non-hydrogen atoms were refined anisotropically. Table S1 summarizes crystallographic data, while CCDC 2207425 contains the supplementary crystallographic data for this paper.

3. Results and Discussion

3.1. Crystal Structure and Literture Analysis

The polycrystalline product formed after the slow cooling of the reaction mixture which contained single crystals suitable for SCXRD. The compound crystalizes as solvate with two pyridines in monoclinic space group P 21/c (Z = 2). The crystallographic data are summarized in Table S1. The unit cell contains three independent Mo atoms and three inner Ii atoms belonging to the same cluster unit. All atoms belonging to cluster core {Mo6Ii8} are located in general positions while the center of the cluster coincides with the special centrosymmetrical position (0, 0, ½). Two of molybdenum atoms are coordinated to apical iodine ligand (Ia) and other one by pyridine (pya) or iodine (Ia) with occupancies 0.9 and 0.1, respectively. Thus, the cluster unit represents superposition of at least two cluster complexes, namely anionic [{Mo6I8}(py)I5] and neutral [{Mo6I8}(py)2I4] (Figure 1B,C). According to the symmetry of the cluster unit (Ci site symmetry) the neutral cluster complex [{Mo6I8}(py)2I4] is trans-isomer. In addition, such position symmetry assumes the presence of third possible cluster unit—the initial [{Mo6I8}I6]2– (Figure 1A). Based on the atoms occupancies and Ci symmetry, the possible ratio of cluster units [{Mo6I8}I6]2–:[{Mo6I8}(py)I5]:[{Mo6I8}(py)2I4] is 0.01:0.18:0.81. According to probability calculation, the content of [{Mo6I8}I6]2– is low, and thus, the formula of the compound can be designated as (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py.
Additionally, solvate pyridine molecule/pyridinium cation is located in the unit cell and interacts with terminal ligands (pyridine or Ia, respectively). The N···H–C hydrogen contacts (pysolv···pycoord) are about 3.31 Å and pyH···Ia are about 4.22 Å. The cluster unit has a typical for this type of clusters geometry, and Mo–Mo, Mo–Ii and Mo–Ia distances are in good agreement with the literature data (Table 1) [40,41].
To the best of our knowledge, only four articles on preparation of halide Mo6-clusters with N-heterocyclic ligands are reported. J. Kraft and H. Schäfer in 1985 described preparation of heteroleptic clusters [{Mo6X8}(py)2X4] (X = Cl, Br) and [{Mo6Cl8}(N-base)2X4] (N-base = 2,2′-bipyridine and 2-phenylpyridine; X = Cl, Br, I) [47]. These compounds were obtained by direct reaction of [{Mo6Cl8}X2X4/2] with corresponding N-heterocycle in ethanol or chloroform under reflux conditions. All compounds obtained were characterized only by elemental and thermogravimetric (TG) analyses, while no single-crystal data were obtained. Nevertheless, the authors managed to determine the crystal lattice parameters of compounds with pyridine and 2,2′-bipyridine ligands (Table S2). The crystal lattice parameters for [{Mo6X8}(py)2X4] are different from reported here, presumably due to the presence of solvate pyridine molecules in (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py.
Another article by C. Perrin with co-authors dating from 2006 devoted to the preparation of heteroleptic (Bu4N)[{Mo6X8}(R-py)Br5] and homoleptic [{Mo6X8}(R-py)6](CF3SO3)4 clusters (X = Br, I, R = H, 4-tert-Bu, 4-vinyl; in some cases R-py—dendrons of complicate composition) from (Bu4N)2[{Mo6Br8}(CF3SO3)Br5] and (Bu4N)2[Mo6Br8}(CF3SO3)6], respectively [48]. These compounds were well characterized by NMR and mass-spectrometry, however, single crystals data even for complexes with simple pyridine ligands were not obtained.
The first crystallographic data of Mo6-halide clusters were obtained only in 2020 by M. Sokolov et al. for homoleptic {Mo6I8}-cluster with triazole ligands, namely, (Bu4N)2[{Mo6I8(N3C2H(COOCH3))6] [40]. This compound was obtained by [2+3] cycloaddition of CH3O(O)C≡C(O)OCH3 to hexaazide cluster (Bu4N)2[{Mo6I8}(N3)6]. Later, this scientific group obtained two other clusters with tetrazole ligands, [{Mo6I8}(N4C(C6H5))6]2– and [{Mo6I8}(N4C(C6F5))6]2–, using the same method [41]. Mo–N bond lengths in [{Mo6I8}(py)2I4] are similar to those for triazole- and tetrazole-coordinated Mo6-clusters (Table 1).

3.2. Synthetic Features and Characterization

Heating a dark red solution of (Bu4N)2[{Mo6I8}I6] in pyridine at 110 °C for 48 h leads to the precipitation of a dark red crystalline product. Single-crystal analysis as well as elemental analysis (CHN, EDS) revealed the formation of a new cluster compound with the formula (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py, which can be considered as a co-crystallization of two cluster complexes in the ratio of 1:4—anionic [{Mo6I8}(py)I5] (Figure 1B) and neutral trans-[{Mo6I8}(py)2I4] (Figure 1C), i.e., {(pyH)[{Mo6I8}(py)I5]·py}0.2{trans-[{Mo6I8}(py)2I4]·2py}0.8. Thus, during the reaction, the substitution of one or two terminal I-ligands occurs. A low reaction yield (about 23%) is probably associated with the equilibrium in the synthesis process presented in the following equations:
[{Mo6I8}I6]2− + py ⇌ [{Mo6I8}(py)I5] + I
[{Mo6I8}(py)I5] + py ⇌ [{Mo6I8}(py)2I4] + I
Iodine anion, released into the solution during the substitution reaction, shifts the equilibrium to the reagent’s side. Moreover, the formation of soluble products also shifts the equilibrium to the same side. Neutral trans-[{Mo6I8}(py)2I4] is partially soluble in hot pyridine, while [{Mo6I8}(py)I5], as an ionic compound, should be relatively soluble in pyridine even at room temperature. Indeed, HR-MS of mother liquor showed the presence of monosubstituted form besides unreacted [{Mo6I8}I6]2– (Figure S2). Thus, when the concentration of I and [{Mo6I8}(py)I5] exceeds some limit, the reaction completely stops, which results in the low yield of the desired neutral cluster. This suggestion was supported by additional experiments. Subsequent heating of the mother liquor after the removal of the solid (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py, as well as heating the (Bu4N)2[{Mo6I8}I6] solution in pyridine in the presence of an excess of I (in the form of Bu4NI), did not lead to the formation of any precipitate of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py. At the same time, heating (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py in pure pyridine in the presence of Bu4NI resulted in the complete dissolution of the solid, confirming the reversibility of the process.
Phase purity and composition of the compound obtained was confirmed by XRPD, FTIR, and 1H-NMR. Diffractogram of the powder product almost fully coincides with calculated data, except for several low intensity reflexes at low angles (Figure 2A). We believe that these signals could correspond to the admixture of pure (pyH)[{Mo6I8}(py)I5]. According to FTIR, all ligand-related bands are preserved in (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py (Figure 2B). The coordination of the pyridine ligands was also confirmed using NMR spectroscopy. Due to the low solubility of products (less than 1 mg in 0.6 mL of solvent), only the 1H NMR and 2D ¹H-¹H COSY spectra were obtained. According to the NMR data, three series of the signal in the aromatic area with a ratio of 1:0.4:0.2 were observed (Figure 2C and Figure S3). The main components are the free-pyridine molecules with chemical shifts at 7.39, 7.78, and 8.58 ppm for meta-, para-, and ortho-protons, respectively. The high content of free pyridine can be explained by its leaching from the powdered complex. The coordination of pyridine to molybdenum atoms in the cluster core causes downfield shifts of the signals due to the electron-withdrawing nature of the {Mo6I8}4+ motif. As expected, the greatest shift in comparison with pure pyridine was observed for 2,6-atoms of hydrogen: 0.65 and 0.81 ppm. It is highly likely that the most downfield signal (9.38 ppm) refers to monosubstituted form [{Mo6I8}(py)I5], whereas the second signal (9.22 ppm) refers to the main product, trans-[{Mo6I8}(py)2I4], since the electron-withdrawing effect of the {Mo6I8}4+ core is probably decreased with the addition of pyridine ligands. Chemical shifts for other hydrogen atoms are close, and the signals from mono- and di-substituted forms overlap (Figure S3) at 8.12 ppm for para-positions (downshift relative to the free pyridine by 0.24 ppm) and 7.53 for meta-positions (downshift by 0.15 ppm). In an attempt to increase the cluster concentration in DMSO-d6, we heated the solution at 80 °C for 30 min. This, indeed, led to a more intense color of the solution, while 1H NMR spectra contained only the forms that referred to free-pyridine molecules (Figure S4), thus, indicating the complete substitution of py ligands with DMSO.
According to the literature, known [{Mo6X8}(N-base)2X4] compounds lose two pyridine molecules at about 500 °C with the formation of Mo6X12 [47]. The compound obtained here, (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py, demonstrates three stages of temperature decomposition: a gradual mass loss (≈3%) up to 360 °C, a sharp mass loss in temperature interval of 360–460 °C (≈7%), and a gradual mass loss up to 550 °C (≈2%) (Figure 2D). In total, the compound loss of approximately 12% of mass corresponds to 3.8 molecules of pyridine (calcd. 12.4%) and correlates well with crystal data and elemental analyses. A slight increase in the decomposition temperature in comparison with [{Mo6X8}(N-base)2X4] (X = Cl, Br) may indicate stronger Mo-N bonds in the case of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py, however, the lack of crystalline structure data do not allow us to analyze bond lengths.

3.3. Optical Properties

The absorption properties of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py were studied both in the solution and solid state and were compared to the initial cluster (Bu4N)2[{Mo6I8}I6] (Figure 3). According to NMR data, DMSO can substitute pyridine ligands in the cluster (Figure S4), thus, pyridine was used as a solvent in order to prevent even a partial substitution. One can see that in the pyridine solution, both clusters have almost similar absorption profiles extended up to ~550 nm, though (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py demonstrates more pronounced peaks at ~351 and ~500 nm, and a redshifted peak at ~426 nm in comparison to (Bu4N)2[{Mo6I8}I6] (~420 nm) (Figure 3A). Overall, these changes are caused by changes in the ligand’s environment, yet this effect is weak due to preservation of the majority of apical iodide ligands. Even less differences were observed in the case of the absorption in the solid state (Figure 3B). One can see that the profile shapes of both compounds are almost similar. The optical energy gaps (Eg), referred to as spin-allowed transitions (singlet state to singlet state), were calculated via the Tauc plot (see the inset in Figure 3). The Eg values of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py and (Bu4N)2[{Mo6I8}I6] are close and equal to 1.75 and 1.6 eV, respectively.
Powdered (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py exhibits very low emissions, but its maximum coincides with that of (Bu4N)2[{Mo6I8}I6] (λem ~ 700 nm) (Figure 4A). Interestingly, the reverse situation is observed in the solution in pyridine, and the complex obtained here exhibits an emission intensity comparable to that of the initial complex (Figure 4B). This behavior is explained by a decrease in inter-cluster interactions in the solution, thus, reducing the efficiency of non-radiative energy transfer [2,49]. Indeed, according to crystal data, the distances between apical and inner (Ia-Ii and Ii-Ii) iodine ligands belonging to neighboring cluster units are 3.800–3.875 Å (Figures S5 and S6), which is lower than the sum of van der Waals radii (1.98 Å), thus, indicating the formation of weak halogen (I-I) bonds [50]. Obviously, these interactions do not exist in the solution, resulting in the appearance of emissions after dissolution. Due to the presence of a large cation in (Bu4N)2[{Mo6I8}I6], there are no noticeable inter-cluster interactions, and thus, the differences in its emission in the solution and solid state are not so significant. It should be noted that the excitation and emission spectra of both compounds have an identical shape, which is consistent with the weak effect of the partial substitution on the optical properties.

4. Conclusions

To conclude, here, by simple heating (Bu4N)2[{Mo6I8}I6] in pyridine, we obtained a rare example of apically heteroleptic neutral octahedral iodide molybdenum cluster complexes with N-base ligand—trans-[{Mo6I8}(py)2I4]. Due to the reversibility of the reaction process, the iodine anions released during the synthesis shift the equilibrium to the reagent’s side, stopping the reaction, and thus, resulting in a low yield of the desired product. According to SCXRD, the entitled compound co-crystallizes with anionic monosubstituted form [{Mo6I8}(py)I5] in the ratio of 1:4. TGA data revealed higher temperatures of pyridine ligand loss in comparison with the literature, thus, indicating stronger Mo-N bonds in the case of Mo6 iodide clusters. This can be an additional driving force of the reaction, despite the natural competition of iodine to be bonded to Mo6I8 units. The formation of [{Mo6I8}(py)I5] was also confirmed by HR-MS and 1H-NMR. The absorption of the entitled compound is almost similar to the initial cluster both in the solution and solid state, with optical bandgap values of 1.75 vs. 1.6 eV for (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py and (Bu4N)2[{Mo6I8}I6], respectively. Similar behavior was observed in the case of luminescent properties in the pyridine solution, while in the solid state, the pyridine-substituted cluster showed almost no emissions. Overall, this indicates a weak effect of the partial ligand substitution on optical properties.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/sym14102117/s1, Figure S1: Illustration of pysolv···pycoord and pyH···Ia interactions.; Figure S2: Fragment of ESI-MS of mother liquor illustrating the presence of [{Mo6I8}(py)I5].; Figure S3: 1H-COSY NMR spectrum of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py; Figure S4: 1H-NMR spectrum of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py after heating in DMSO-d6; Figure S5: Ii-Ia and Ii-Ii interactions in (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py; Figure S6: Layers in (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py; Table S1: Selected crystallographic parameters of the single-crystal X-ray diffraction structural analysis for (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py; Table S2: Crystal lattice parameters for (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py and [{Mo6X8}(py)2X4] (X = Cl, Br).

Author Contributions

Conceptualization, Y.A.V. and M.A.S.; methodology, Y.A.V.; validation, Y.A.V., A.A.I. and M.A.S.; formal analysis, Y.A.V., I.V.E., N.V.K. and M.A.S.; investigation, M.V.M.; resources, I.V.E. and N.V.K.; data curation, Y.A.V. and M.A.S.; writing—original draft preparation, Y.A.V. and M.A.S.; writing—review and editing, Y.A.V. and M.A.S.; visualization, Y.A.V.; supervision, Y.A.V. and M.A.S.; project administration, M.A.S.; funding acquisition, A.A.I. and M.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by The Council for Grants of the President of the Russian Federation, grant number MK-87.2022.1.3. YAV thanks The Council for Grants of the President of the Russian Federation (scholarship № SP-3328.2021.1). NIIC team thanks the Ministry of Science and Higher Education of the Russian Federation (№ 121031700321-3 and 121031700313-8).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mikhaylov, M.A.; Sokolov, M.N. Molybdenum Iodides—From Obscurity to Bright Luminescence. Eur. J. Inorg. Chem. 2019, 2019, 4181–4197. [Google Scholar] [CrossRef]
  2. Marchuk, M.V.; Vorotnikova, N.A.; Vorotnikov, Y.A.; Kuratieva, N.V.; Stass, D.V.; Shestopalov, M.A. Optical Property Trends in a Family of {Mo6I8} Aquahydroxo Complexes. Dalton Trans. 2021, 50, 8794–8802. [Google Scholar] [CrossRef]
  3. Kirakci, K.; Kubat, P.; Fejfarova, K.; Martincik, J.; Nikl, M.; Lang, K. X-ray Inducible Luminescence and Singlet Oxygen Sensitization by an Octahedral Molybdenum Cluster Compound: A New Class of Nanoscintillators. Inorg. Chem. 2016, 55, 803–809. [Google Scholar] [CrossRef]
  4. Efremova, O.A.; Vorotnikov, Y.A.; Brylev, K.A.; Vorotnikova, N.A.; Novozhilov, I.N.; Kuratieva, N.V.; Edeleva, M.V.; Benoit, D.M.; Kitamura, N.; Mironov, Y.V.; et al. Octahedral Molybdenum Cluster Complexes with Aromatic Sulfonate Ligands. Dalton Trans. 2016, 45, 15427–15435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Mikhailov, M.A.; Brylev, K.A.; Abramov, P.A.; Sakuda, E.; Akagi, S.; Ito, A.; Kitamura, N.; Sokolov, M.N. Synthetic Tuning of Redox, Spectroscopic, and Photophysical Properties of {Mo6I8}4+ Core Cluster Complexes by Terminal Carboxylate Ligands. Inorg. Chem. 2016, 55, 8437–8445. [Google Scholar] [CrossRef] [PubMed]
  6. Akagi, S.; Fujii, S.; Kitamura, N. A Study on the Redox, Spectroscopic, and Photophysical Characteristics of a Series of Octahedral Hexamolybdenum(II) Clusters: [{Mo6X8}Y6]2− (X, Y = Cl, Br, or I). Dalton Trans. 2018, 47, 1131–1139. [Google Scholar] [CrossRef]
  7. Evtushok, D.V.; Melnikov, A.R.; Vorotnikova, N.A.; Vorotnikov, Y.A.; Ryadun, A.A.; Kuratieva, N.V.; Kozyr, K.V.; Obedinskaya, N.R.; Kretov, E.I.; Novozhilov, I.N.; et al. A Comparative Study of Optical Properties and X-ray Induced Luminescence of Octahedral Molybdenum and Tungsten Cluster Complexes. Dalton Trans. 2017, 46, 11738–11747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Vorotnikov, Y.A.; Efremova, O.A.; Vorotnikova, N.A.; Brylev, K.A.; Edeleva, M.V.; Tsygankova, A.R.; Smolentsev, A.I.; Kitamura, N.; Mironov, Y.V.; Shestopalov, M.A. On the Synthesis and Characterisation of Luminescent Hybrid Particles: Mo6 Metal Cluster Complex/SiO2. RSC Adv. 2016, 6, 43367–43375. [Google Scholar] [CrossRef] [Green Version]
  9. Jackson, J.A.; Turro, C.; Newsham, M.D.; Nocera, D.G. Oxygen Quenching of Electronically Excited Hexanuclear Molybdenum and Tungsten Halide Clusters. J. Phys. Chem. 1990, 94, 4500–4507. [Google Scholar] [CrossRef]
  10. Kirakci, K.; Kubát, P.; Langmaier, J.; Polívka, T.; Fuciman, M.; Fejfarová, K.; Lang, K. A Comparative Study of the Redox and Excited State Properties of (nBu4N)2[Mo6X14] and (nBu4N)2[Mo6X8(CF3COO)6] (X = Cl, Br, or I). Dalton Trans. 2013, 42, 7224–7232. [Google Scholar] [CrossRef] [PubMed]
  11. Krasilnikova, A.A.; Shestopalov, M.A.; Brylev, K.A.; Kirilova, I.A.; Khripko, O.P.; Zubareva, K.E.; Khripko, Y.I.; Podorognaya, V.T.; Shestopalova, L.V.; Fedorov, V.E.; et al. Prospects of Molybdenum and Rhenium Octahedral Cluster Complexes as X-ray Contrast Agents. J. Inorg. Biochem. 2015, 144, 13–17. [Google Scholar] [CrossRef]
  12. Ivanova, M.N.; Vorotnikov, Y.A.; Plotnikova, E.E.; Marchuk, M.V.; Ivanov, A.A.; Asanov, I.P.; Tsygankova, A.R.; Grayfer, E.D.; Fedorov, V.E.; Shestopalov, M.A. Hexamolybdenum Clusters Supported on Exfoliated h-BN Nanosheets for Photocatalytic Water Purification. Inorg. Chem. 2020, 59, 6439–6448. [Google Scholar] [CrossRef] [PubMed]
  13. Kumar, S.; Khatri, O.P.; Cordier, S.; Boukherroub, R.; Jain, S.L. Graphene Oxide Supported Molybdenum Cluster: First Heterogenized Homogeneous Catalyst for the Synthesis of Dimethylcarbonate from CO2 and Methanol. Chem. Eur. J. 2015, 21, 3488–3494. [Google Scholar] [CrossRef]
  14. Barras, A.; Cordier, S.; Boukherroub, R. Fast Photocatalytic Degradation of Rhodamine B over [Mo6Br8(N3)6]2− Cluster Units under Sun Light Irradiation. Appl. Catal. B Environ. 2012, 123–124, 1–8. [Google Scholar] [CrossRef]
  15. Feliz, M.; Atienzar, P.; Amela-Cortés, M.; Dumait, N.; Lemoine, P.; Molard, Y.; Cordier, S. Supramolecular Anchoring of Octahedral Molybdenum Clusters onto Graphene and Their Synergies in Photocatalytic Water Reduction. Inorg. Chem. 2019, 58, 15443–15454. [Google Scholar] [CrossRef] [Green Version]
  16. Nguyen, N.T.K.; Lebastard, C.; Wilmet, M.; Dumait, N.; Renaud, A.; Cordier, S.; Ohashi, N.; Uchikoshi, T.; Grasset, F. A Review on Functional Nanoarchitectonics Nanocomposites Based on Octahedral Metal Atom Clusters (Nb6, Mo6, Ta6, W6, Re6): Inorganic 0D and 2D Powders and Films. Sci. Technol. Adv. Mater. 2022, 23, 547–578. [Google Scholar] [CrossRef] [PubMed]
  17. Molard, Y. Clustomesogens: Liquid Crystalline Hybrid Nanomaterials Containing Functional Metal Nanoclusters. Acc. Chem. Res. 2016, 49, 1514–1523. [Google Scholar] [CrossRef] [PubMed]
  18. Guy, K.; Ehni, P.; Paofai, S.; Forschner, R.; Roiland, C.; Amela-Cortes, M.; Cordier, S.; Laschat, S.; Molard, Y. Lord of The Crowns: A New Precious in the Kingdom of Clustomesogens. Angew. Chem. Int. Ed. 2018, 57, 11692–11696. [Google Scholar] [CrossRef] [PubMed]
  19. Riehl, L.; Ströbele, M.; Enseling, D.; Jüstel, T.; Meyer, H.-J. Molecular Oxygen Modulated Luminescence of an Octahedro-hexamolybdenum Iodide Cluster having Six Apical Thiocyanate Ligands. Z. Anorg. Allg. Chem. 2016, 642, 403–408. [Google Scholar] [CrossRef]
  20. Fuhrmann, A.-D.; Seyboldt, A.; Schank, A.; Zitzer, G.; Speiser, B.; Enseling, D.; Jüstel, T.; Meyer, H.-J. Luminescence Quenching of Ligand-Substituted Molybdenum and Tungsten Halide Clusters by Oxygen and Their Oxidation Electrochemistry. Eur. J. Inorg. Chem. 2017, 2017, 4259–4266. [Google Scholar] [CrossRef]
  21. Amela-Cortes, M.; Paofai, S.; Cordier, S.; Folliot, H.; Molard, Y. Tuned Red NIR Phosphorescence of Polyurethane Hybrid Composites Embedding Metallic Nanoclusters for Oxygen Sensing. Chem. Commun. 2015, 51, 8177–8180. [Google Scholar] [CrossRef] [PubMed]
  22. Riehl, L.; Seyboldt, A.; Ströbele, M.; Enseling, D.; Jüstel, T.; Westberg, M.; Ogilby, P.R.; Meyer, H.J. A Ligand Substituted Tungsten Iodide Cluster: Luminescence vs. Singlet Oxygen Production. Dalton Trans. 2016, 45, 15500–15506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Khlifi, S.; Bigeon, J.; Amela-Cortes, M.; Dumait, N.; Loas, G.h.; Cordier, S.; Molard, Y. Switchable Two-Dimensional Waveguiding Abilities of Luminescent Hybrid Nanocomposites for Active Solar Concentrators. ACS Appl. Mater. Interfaces 2020, 12, 14400–14407. [Google Scholar] [CrossRef] [PubMed]
  24. Zhao, Y.; Lunt, R.R. Transparent Luminescent Solar Concentrators for Large-Area Solar Windows Enabled by Massive Stokes-Shift Nanocluster Phosphors. Adv. Energy Mater. 2013, 3, 1143–1148. [Google Scholar] [CrossRef]
  25. Renaud, A.; Jouan, P.-Y.; Dumait, N.; Ababou-Girard, S.; Barreau, N.; Uchikoshi, T.; Grasset, F.; Jobic, S.; Cordier, S. Evidence of the Ambipolar Behavior of Mo6 Cluster Iodides in All-Inorganic Solar Cells: A New Example of Nanoarchitectonic Concept. ACS Appl. Mater. Interfaces 2022, 14, 1347–1354. [Google Scholar] [CrossRef]
  26. Renaud, A.; Nguyen, T.K.N.; Grasset, F.; Raissi, M.; Guillon, V.; Delabrouille, F.; Dumait, N.; Jouan, P.Y.; Cario, L.; Jobic, S.; et al. Preparation by Electrophoretic Deposition of Molybdenum Iodide Cluster-Based Functional Nanostructured Photoelectrodes for Solar Cells. Electrochim. Acta 2019, 317, 737–745. [Google Scholar] [CrossRef]
  27. Kirakci, K.; Pozmogova, T.N.; Protasevich, A.Y.; Vavilov, G.D.; Stass, D.V.; Shestopalov, M.A.; Lang, K. A Water-Soluble Octahedral Molybdenum Cluster Complex as a Potential Agent for X-ray Induced Photodynamic Therapy. Biomater. Sci. 2021, 9, 2893–2902. [Google Scholar] [CrossRef]
  28. Kirakci, K.; Kubáňová, M.; Přibyl, T.; Rumlová, M.; Zelenka, J.; Ruml, T.; Lang, K. A Cell Membrane Targeting Molybdenum-Iodine Nanocluster: Rational Ligand Design toward Enhanced Photodynamic Activity. Inorg. Chem. 2022, 61, 5076–5083. [Google Scholar] [CrossRef]
  29. Koncošová, M.; Rumlová, M.; Mikyšková, R.; Reiniš, M.; Zelenka, J.; Ruml, T.; Kirakci, K.; Lang, K. Avenue to X-ray-Induced Photodynamic Therapy of Prostatic Carcinoma with Octahedral Molybdenum Cluster Nanoparticles. J. Mater. Chem. B 2022, 10, 3303–3310. [Google Scholar] [CrossRef]
  30. Vorotnikov, Y.A.; Novikova, E.D.; Solovieva, A.O.; Shanshin, D.V.; Tsygankova, A.R.; Shcherbakov, D.N.; Efremova, O.A.; Shestopalov, M.A. Single-Domain Antibody C7b for Address Delivery of Nanoparticles to HER2-Positive Cancers. Nanoscale 2020, 12, 21885–21894. [Google Scholar] [CrossRef]
  31. Svezhentseva, E.V.; Vorotnikov, Y.A.; Solovieva, A.O.; Pozmogova, T.N.; Eltsov, I.V.; Ivanov, A.A.; Evtushok, D.V.; Miroshnichenko, S.M.; Yanshole, V.V.; Eling, C.J.; et al. From Photoinduced to Dark Cytotoxicity through an Octahedral Cluster Hydrolysis. Chem. Eur. J. 2018, 24, 17915–17920. [Google Scholar] [CrossRef] [PubMed]
  32. Solovieva, A.O.; Vorotnikov, Y.A.; Trifonova, K.E.; Efremova, O.A.; Krasilnikova, A.A.; Brylev, K.A.; Vorontsova, E.V.; Avrorov, P.A.; Shestopalova, L.V.; Poveshchenko, A.F.; et al. Cellular Internalisation, Bioimaging and Dark and Photodynamic Cytotoxicity of Silica Nanoparticles Doped by {Mo6I8}4+ Metal Clusters. J. Mater. Chem. B 2016, 4, 4839–4846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Vorotnikov, Y.A.; Pozmogova, T.N.; Solovieva, A.O.; Miroshnichenko, S.M.; Vorontsova, E.V.; Shestopalova, L.V.; Mironov, Y.V.; Shestopalov, M.A.; Efremova, O.A. Luminescent Silica Mesoparticles for Protein Transduction. Mat. Sci. Eng. C Mater. 2019, 96, 530–538. [Google Scholar] [CrossRef]
  34. Beltran, A.; Mikhailov, M.; Sokolov, M.N.; Perez-Laguna, V.; Rezusta, A.; Revillo, M.J.; Galindo, F. A Photobleaching Resistant Polymer Supported Hexanuclear Molybdenum Iodide Cluster for Photocatalytic Oxygenations and Photodynamic Inactivation of Staphylococcus Aureus. J. Mater. Chem. B 2016, 4, 5975–5979. [Google Scholar] [CrossRef] [Green Version]
  35. Vorotnikova, N.A.; Alekseev, A.Y.; Vorotnikov, Y.A.; Evtushok, D.V.; Molard, Y.; Amela-Cortes, M.; Cordier, S.; Smolentsev, A.I.; Burton, C.G.; Kozhin, P.M.; et al. Octahedral Molybdenum Cluster as a Photoactive Antimicrobial Additive to a Fluoroplastic. Mat. Sci. Eng. C Mater. 2019, 105, 110150. [Google Scholar] [CrossRef]
  36. Kirakci, K.; Nguyen, T.K.N.; Grasset, F.; Uchikoshi, T.; Zelenka, J.; Kubát, P.; Ruml, T.; Lang, K. Electrophoretically Deposited Layers of Octahedral Molybdenum Cluster Complexes: A Promising Coating for Mitigation of Pathogenic Bacterial Biofilms under Blue Light. ACS Appl. Mater. Interfaces 2020, 12, 52492–52499. [Google Scholar] [CrossRef]
  37. Vorotnikova, N.A.; Bardin, V.A.; Vorotnikov, Y.A.; Kirakci, K.; Adamenko, L.S.; Alekseev, A.Y.; Meyer, H.J.; Kubat, P.; Mironov, Y.V.; Lang, K.; et al. Heterogeneous Photoactive Antimicrobial Coatings Based on a Fluoroplastic Doped with an Octahedral Molybdenum Cluster Compound. Dalton Trans. 2021, 50, 8467–8475. [Google Scholar] [CrossRef]
  38. Kerru, N.; Gummidi, L.; Maddila, S.; Gangu, K.K.; Jonnalagadda, S.B. A Review on Recent Advances in Nitrogen-Containing Molecules and Their Biological Applications. Molecules 2020, 25, 1909. [Google Scholar] [CrossRef] [Green Version]
  39. Mermer, A.; Keles, T.; Sirin, Y. Recent studies of nitrogen containing heterocyclic compounds as novel antiviral agents: A review. Bioorg. Chem. 2021, 114, 105076. [Google Scholar] [CrossRef]
  40. Mironova, A.D.; Mikhailov, M.A.; Brylev, K.A.; Gushchin, A.L.; Sukhikh, T.S.; Sokolov, M.N. Phosphorescent Complexes of {Mo6I8}4+ with Triazolates: [2+3] Cycloaddition of Alkynes to [Mo6I8(N3)6]2−. New J. Chem. 2020, 44, 20620–20625. [Google Scholar] [CrossRef]
  41. Mironova, A.; Gushchin, A.; Abramov, P.; Eltsov, I.; Ryadun, A.; Sokolov, M. [Mo6I8]4+ Complexes with Tetrazolate Ligands: [3+2] Cycloaddition of Aromatic Nitriles to [Mo6I8(N3)6]2−. Polyhedron 2021, 205, 115282. [Google Scholar] [CrossRef]
  42. Konovalov, D.I.; Ivanov, A.A.; Vorotnikov, Y.A.; Kuratieva, N.V.; Eltsov, I.V.; Kovalenko, K.A.; Shestopalov, M.A. Self-Assembled Microporous M-HOFs Based on an Octahedral Rhenium Cluster with Benzimidazole. Inorg. Chem. 2021, 60, 14687–14696. [Google Scholar] [CrossRef] [PubMed]
  43. Konovalov, D.I.; Ivanov, A.A.; Frolova, T.S.; Eltsov, I.V.; Gayfulin, Y.M.; Plunkett, L.; Bazzar, M.; Adawi, A.M.; Bouillard, J.-S.G.; Baiborodin, S.I.; et al. Water-Soluble Rhenium Clusters with Triazoles: The Effect of Chemical Structure on Cellular Internalization and the DNA Binding of the Complexes. Chem. Eur. J. 2020, 26, 13904–13914. [Google Scholar] [CrossRef] [PubMed]
  44. Kirakci, K.; Cordier, S.; Perrin, C. Synthesis and Characterization of Cs2Mo6X14 (X = Br or I) Hexamolybdenum Cluster Halides: Efficient Mo6 Cluster Precursors for Solution Chemistry Syntheses. Z. Anorg. Allg. Chem. 2005, 631, 411–416. [Google Scholar] [CrossRef]
  45. Bruker. APEX2, Version 1.08, SAINT, Version 07.03, SADABS, Version 02.11, SHELXTL, Version 06.12; Bruker AXS Inc.: Madison, WI, USA, 2004. [Google Scholar]
  46. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Kraft, J.; Schäfer, H. Die Reaktion von [Mo6Cl8]X4 mit N-Basen. Z. Anorg. Allg. Chem. 1985, 524, 137–143. [Google Scholar] [CrossRef]
  48. Méry, D.; Plault, L.; Ornelas, C.; Ruiz, J.; Nlate, S.; Astruc, D.; Blais, J.-C.; Rodrigues, J.; Cordier, S.; Kirakci, K.; et al. From Simple Monopyridine Clusters [Mo6Br13(Py-R)][n-Bu4N] and Hexapyridine Clusters [Mo6X8(Py-R)6][OSO2CF3]4 (X = Br or I) to Cluster-Cored Organometallic Stars, Dendrons, and Dendrimers. Inorg. Chem. 2006, 45, 1156–1167. [Google Scholar] [CrossRef]
  49. Costuas, K.; Garreau, A.; Bulou, A.; Fontaine, B.; Cuny, J.; Gautier, R.; Mortier, M.; Molard, Y.; Duvail, J.L.; Faulques, E.; et al. Combined Theoretical and Time-Resolved Photoluminescence Investigations of [Mo6Bri8Bra6]2− Metal Cluster Units: Evidence of Dual Emission. Phys. Chem. Chem. Phys. 2015, 17, 28574–28585. [Google Scholar] [CrossRef]
  50. Xin, D.; Matti, T.; Matti, H. Halogen Bonding in Crystal Engineering. In Recent Advances in Crystallography; Benedict, J.B., Ed.; IntechOpen: London, UK, 2012; Volume 262, pp. 143–168. [Google Scholar]
Figure 1. Crystal structures of [{Mo6I8}I6]2– (A) [{Mo6I8}(py)I5] (B) and [{Mo6I8}(py)2I4] (C) Color code: light blue octahedron—Mo6, purple—iodine, blue—nitrogen, black—carbon, gray—hydrogen.
Figure 1. Crystal structures of [{Mo6I8}I6]2– (A) [{Mo6I8}(py)I5] (B) and [{Mo6I8}(py)2I4] (C) Color code: light blue octahedron—Mo6, purple—iodine, blue—nitrogen, black—carbon, gray—hydrogen.
Symmetry 14 02117 g001
Figure 2. (A) Experimental and calculated diffractograms of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py. Asterisks demonstrate the presence of the unknown phase attributed to pure monosubstituted form (pyH)[{Mo6I8}(py)I5]. (B) FTIR spectra of (Bu4N)2[{Mo6I8}I6], py, and (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py. (C) 1H-NMR spectrum of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py. (D) TG curve of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py.
Figure 2. (A) Experimental and calculated diffractograms of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py. Asterisks demonstrate the presence of the unknown phase attributed to pure monosubstituted form (pyH)[{Mo6I8}(py)I5]. (B) FTIR spectra of (Bu4N)2[{Mo6I8}I6], py, and (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py. (C) 1H-NMR spectrum of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py. (D) TG curve of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py.
Symmetry 14 02117 g002
Figure 3. (A) UV-vis spectra of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py and (Bu4N)2[{Mo6I8}I6] in pyridine. (B) Diffuse reflectance spectra of of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py and (Bu4N)2[{Mo6I8}I6] converted to absorption spectra using the Kubelka-Munk function. Incert: Tauc plots used for determination of Eg.
Figure 3. (A) UV-vis spectra of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py and (Bu4N)2[{Mo6I8}I6] in pyridine. (B) Diffuse reflectance spectra of of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py and (Bu4N)2[{Mo6I8}I6] converted to absorption spectra using the Kubelka-Munk function. Incert: Tauc plots used for determination of Eg.
Symmetry 14 02117 g003
Figure 4. Excitation and emission spectra of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py and (Bu4N)2[{Mo6I8}I6] in the solid state (A) and in the pyridine solution (B).
Figure 4. Excitation and emission spectra of (pyH)0.2[{Mo6I8}(py)1.8I4.2]·1.8py and (Bu4N)2[{Mo6I8}I6] in the solid state (A) and in the pyridine solution (B).
Symmetry 14 02117 g004
Table 1. Interatomic distances in trans-[{Mo6I8}(py)2I4] and related compounds.
Table 1. Interatomic distances in trans-[{Mo6I8}(py)2I4] and related compounds.
Cluster ComplexInteratomic Distances, ÅReferences
Mo–MoMo–IiMo–IaMo–N
trans-[{Mo6I8}(py)2I4]2.675(2)–
2.6900(17)
2.7656(18)–
2.7891(16)
2.75(2)–
2.8432(19)
2.22(3)this work
[{Mo6I8}(N3C2H(COOCH3))6]2−2.6721(7)–
2.6794(6)
2.7590(6)–
2.7815(6)
-2.199(4)[40]
[{Mo6I8}(N4C(C6H5))6]2−2.6728(10)–
2.6856(9)
2.7588(8)–
2.7737(9)
2.844(12)–
2.995(8) #
2.230(8)–
2.238(9)
[41]
[{Mo6I8}(N4C(C6F5))6]2−2.6703(12)–
2.6771(12)
2.7557(9)–
2.7760(11)
-2.220(9)–
2.259(9)
# Partially substituted forms.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Marchuk, M.V.; Vorotnikov, Y.A.; Ivanov, A.A.; Eltsov, I.V.; Kuratieva, N.V.; Shestopalov, M.A. A Neutral Heteroleptic Molybdenum Cluster trans-[{Mo6I8}(py)2I4]. Symmetry 2022, 14, 2117. https://doi.org/10.3390/sym14102117

AMA Style

Marchuk MV, Vorotnikov YA, Ivanov AA, Eltsov IV, Kuratieva NV, Shestopalov MA. A Neutral Heteroleptic Molybdenum Cluster trans-[{Mo6I8}(py)2I4]. Symmetry. 2022; 14(10):2117. https://doi.org/10.3390/sym14102117

Chicago/Turabian Style

Marchuk, Margarita V., Yuri A. Vorotnikov, Anton A. Ivanov, Ilia V. Eltsov, Natalia V. Kuratieva, and Michael A. Shestopalov. 2022. "A Neutral Heteroleptic Molybdenum Cluster trans-[{Mo6I8}(py)2I4]" Symmetry 14, no. 10: 2117. https://doi.org/10.3390/sym14102117

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