Skip to content
Publicly Available Published by De Gruyter January 5, 2022

Synthesis and structural characterization of a new heterometallicmolybdate coordination polymer based on a µ3-bridging amino alcohol

  • Savita A. Kundaikar , Sudesh M. Morajkar , Wolfgang Bensch and Bikshandarkoil R. Srinivasan EMAIL logo

Abstract

The reaction of Na2MoO4·2H2O with 2-amino-2-(hydroxymethyl)propane-1,3-diol (LH) in water at room temperature results in the formation of the heterometallic coordination polymer [Mo2O6L2(Na2(H2O)4)]·2H2O 1 (L = 2-amino-3-hydroxy-2-(hydroxymethyl)propan-1-olato). The structure of 1 consists of a neutral (Mo2O6) unit located on an inversion center. The Mo atoms exhibit hexa-coordination and are bonded to two terminal and two bridging oxido ligands, an alkoxide oxygen and the amine N atoms of an anionic ligand L resulting in the formation of an edge-sharing {Mo2O8N2} bioctahedron. The Na+ cations of a centrosymmetric bis(μ2-aqua)-bridged (Na2(H2O)4)2+ unit are penta-coordinated and bonded to two symmetry related L ligands via the oxygen atoms of their OH groups. The µ3-bridging tetradentate binding mode of L results in the formation of a two-dimensional heterometallic coordination polymer. The constituents of 1 viz. (Mo2O6), (L), (Na2(H2O)4)2+ and lattice water molecules are interlinked with the aid of three varieties of hydrogen bonding interactions. The corresponding tungstate reported recently has been obtained through a similar synthetic protocol and is isostructural.

1 Introduction

The chemistry of molybdenum oxido compounds mainly termed as polyoxomolybdates (POMo) is described in all standard chemistry text books [1, 2] and is a frontier area of research in materials science. The interest in these compounds is due to the structural diversity and wide-ranging applications in several areas of science [3], [4], [5], [6], [7]. POMos include a large variety of clusters with varying nuclearity, and compounds containing as many as 368 Mo atoms were reported [8]. The study of oxoanions of Mo dates back to the days of Berzelius when a POMo containing 12 Mo atoms having the formula (NH4)3PMo12O40 was first prepared in acidic medium [9]. However, enormous developments in the area of POMo chemistry have been achieved in the last few decades, especially the structural characterization of high nuclearity POMos [3], [4], [5], [6], [7], [8, 10], [11], [12], [13], [14], [15], [16], [17], [18]. The synthesis of POMos is generally accomplished by acidification of an aqueous tetraoxidomolybdate [MoO4]2– solution in the presence of appropriate counter cations.

Aqueous reaction mixtures of acidified molybdate solutions containing organic amines are known to result in the formation of hepta- or octamolybdates as the favored product [19], [20], [21], [22], [23], [24], [25], [26], [27], [28], [29], [30], [31], [32]. In our research we have shown that the reaction of molybdic acid or its anhydrate MoO3 with organic amines or inorganic bases is a convenient method for the facile synthesis of heptamolybdates [31, 33], [34], [35], [36], [37], [38]. For example, the reaction of MoO3 with n-butylamine afforded a product containing both heptamolybdate and monomolybdate clusters [31]. However, weaker bases behave differently. The non-toxic compound 2-amino-2-(hydroxymethyl)propane-1,3-diol (LH) also referred to as tris(hydroxymethyl)aminomethane or Tris is used extensively in biological studies as a low cost component in buffer solutions and has a pKa of ≈8.5 and a pH range of 7–9 in buffer solutions which fit in the physiological pH range of most living organisms [39], [40], [41], [42]. LH has been used in polyoxometalate (POM) chemistry to obtain a variety of POM products with interesting crystal structures [43], [44], [45], [46], [47], [48]. In an earlier study we have shown that attempts to exchange the ammonium cations of (NH4)6(Mo7O24)·4H2O by (LH2)+ by reaction with LH result in the formation of a mixed cationic monomolybdate (NH4)(LH2)(MoO4) [49]. In a recent paper Gumerova et al. [50] have reported that a condensation reaction of the orthotungstate anion (WO4)2– acidified with HCl and buffered at pH = 7.5 with (LH2Cl) namely 1,3-dihydroxy-2-(hydroxymethyl)propan-2-aminium chloride results in the formation of a discrete polytungstate anion (POT) containing the neutral dinuclear [W2O6] unit. In the present study we report that a hitherto unknown compound [Mo2O6L2Na2(H2O)4]·2H2O 1 as well as its corresponding W analog 2 reported recently [50] can be synthesized in good yields by the reaction of sodium molybdate dihydrate (or sodium tungstate dihydrate) with 2-amino-2-(hydroxymethyl)propane-1,3-diol in water. The synthesis, spectral characterization and crystal structures of the isostructural compounds 1 and 2 are described in this report.

2 Experimental

2.1 Materials and methods

All chemicals were used as received from commercial sources without any further purification. Infrared (IR) spectra of the solid samples diluted with KBr were recorded on a Shimadzu (IR Prestige-21) FT-IR spectrometer in the range 4000–400 cm−1 at a resolution of 4 cm−1. Raman spectra of the solids and their aqueous solutions were measured by using an Agiltron PeakSeeker Pro Raman instrument with 785 nm laser irradiation and a laser power of 100 mW. Isothermal mass loss studies of 1 and 2 were performed in a temperature controlled electric furnace. Simultaneous thermogravimetry (TG) and differential thermal analyses (DTA) of powdered samples of 1 and 2 were performed in alumina crucibles in the temperature range of 30–600 °C, using a Netzsch STA-409 PC thermal analyzer. X-ray intensity data for 1 and 2 were collected on a STOE- Image Plate diffraction System (IPDS-2) and a Bruker D8 Quest Eco X-ray diffractometer, respectively, using graphite-monochromated Mo-Kα radiation. The structures were solved with Direct Methods using Shelxs-97 [51] and refinement was carried out against F2 using Shelxl-2016 [51]. All non-hydrogen atoms were refined anisotropically. The C–H and N–H hydrogen atoms were located in difference maps but were positioned with idealized geometry and refined isotropically with Uiso(H) = 1.2 Ueq(C,N) using a riding model. The O–H hydrogen atoms were located in difference maps, their bond lengths were set to ideal values and finally they were refined isotropically with Uiso(H) = 1.5 Ueq(O) using a riding model. Technical details of data acquisition and selected refinement results for 1 and 2 are given in Table 1.

Table 1:

Crystal data and structure refinement for 1 and 2.

Refinement result Compound 1 Compound 2
Empirical formula C8H32Mo2N2Na2O18 C8H32W2N2Na2O18
Formula weight (g mol−1) 682.21 858.03
Temperature (K) 170(2)K 100(2)
Wavelength (Å) 0.71073 0.71073
Crystal system Triclinic Triclinic
Space group P 1 P 1
Unit cell dimensions
a (Å) 7.1728(4) 7.1802(5)
b (Å) 9.0692(5) 9.0454(7)
c (Å) 10.0146(6) 9.9898(8)
α (°) 64.636(4) 64.673(2)
β (°) 70.885(5)° 71.352(2)
Γ (°) 88.878(5) 89.525(2)
Volume (Å3) 550.62(6) 549.37(7)
Z 1 1
Dcalc (g/cm3) 2.06 2.59
Absorption coefficient (mm−1) 1.3 10.6
F(000) 344 408
Crystal size (mm3) 0.05 × 0.09 × 0.17 0.31 × 0.23 × 0.14
θ range for data collection (°) 2.406 to 29.003 3.615 to 28.312
Limiting indices −9 ≤ h ≤ 9, −12 ≤ k ≤ 12, −13 ≤ l ≤ 13 −9 ≤ h ≤ 9, −12 ≤ k ≤ 12, −13 ≤ l ≤ 13
Reflections collected/unique 10615/2925 [R(int) = 0.0514] 10094/2724 [R(int) = 0.0233]
Completeness θ = 25.242° 100.0% 99.50%
Refinement method Full- matrix least-squares on F2 Full-matrix least-squares on F2
Data/restraints/parameters 2925/0/146 2724/0/177
Goodness of fit on F2 1.056 1.179
Final R indices [I > 2σ(I)] R1 = 0.0259, wR2 = 0.0665 R1 = 0.0132, wR2 = 0.0341
R indices (all data) R1 = 0.0282, wR2 = 0.0671 R1 = 0.0134, wR2 = 0.0343
Largest diff. peak and hole (e Å−3) 0.81 and −0.85 0.78 and −1.28
CCDC deposition no 2128432 2128433

2.2 Synthesis of [M2O6(L)2(Na2(H2O)4)]·2H2O (M = Mo 1; M = W 2)

To a solution of sodium molybdate dihydrate (Na2MoO4·2H2O) (2.42 g, 10 mmol) dissolved in ∼10 mL distilled water, 2-amino-2-(hydroxymethyl)propane-1,3-diol (2.42 g, 20 mmol) taken in ∼10 mL of water was added. The reaction mixture was stirred well and the pH was found to be 11.97. Subsequently the clear solution was left undisturbed. After a few days, the crystals that precipitated were isolated by filtration to obtain 3.03 g of 1. The use of Na2WO4·2H2O (3.298 g, 10 mmol) instead of Na2MoO4·2H2O under identical reaction conditions resulted in the formation of 3.64 g of 2.

Anal. Found (Calcd.) for (1) (%): C, 14.08 (13.95); H, 4.60 (4.73); N, 4.05 (4.11); IR data (cm−1): 3550–2950(br), 2261(br), 1761(m), 1652(m), 1563(m), 1280(m), 1081(νC–O), 1044(νC–O), 881(w), 855(s), 760(s), 618, 486. Raman data (cm−1): 3575, 2791, 2554, 1446, 1052, 883(vs), 764, 321; DTA (°C): 132(endo), 456(exo), 534(exo).

Anal. Found (Calcd.) for (2) (%): C, 11.02 (11.20); H, 3.14 (3.80); N, 3.11 (3.26); IR data (cm−1): 3619–3181(br), 2875(m), 2723(w), 2610(w), 2227(br), 1570(s), 1485(s), 1276(m), 1209(m), 1085(νC–O), 1076(νC–O), 1039(νC–O), 902(s), 848(s), 713(w), 684(w), 412(w). Raman data (cm−1): 3029, 2790, 1466, 1274, 1067, 896(vs), 839, 609; DTA (°C): 128(endo), 209(endo), 320(exo), 491(exo), 584(exo).

3 Results and discussion

3.1 Description of the crystal structures of compounds 1 and 2

The compounds [M2O6L2(Na2(H2O)4)]·2H2O (M = Mo 1; M = W 2; L = 2-amino-3-hydroxy-2-(hydroxymethyl)propan-1-olato) are isostructural and crystallize in the centrosymmetric triclinic space group P 1 . The crystal structures consist of a centrosymmetric neutral [M2O6] unit and a bis(μ2-aqua) bridged (Na2(H2O)4)2+ dication, a crystallographically independent 2-amino-3-hydroxy-2-(hydroxymethyl)propan-1-olato ligand abbreviated as L and a lattice water molecule (Figure 1, Figure S1). The midpoint of the line connecting Mo1 and its symmetry generated counterpart Mo1i (for symmetry code see Table 2) is the center of inversion. The geometric parameters of the amino alcohol ligand are in normal range (Table S1).

Figure 1: 
The asymmetric unit of 1 showing hexa- and penta-coordination around the Mo and Na atoms, respectively. Displacement ellipsoids are drawn at 50% probability level for the non-hydrogen atoms. Intramolecular hydrogen bonding is shown by broken lines. For symmetry code See Table 2 (For the structure of the isotypic compound 2 see Figure S1).
Figure 1:

The asymmetric unit of 1 showing hexa- and penta-coordination around the Mo and Na atoms, respectively. Displacement ellipsoids are drawn at 50% probability level for the non-hydrogen atoms. Intramolecular hydrogen bonding is shown by broken lines. For symmetry code See Table 2 (For the structure of the isotypic compound 2 see Figure S1).

Table 2:

Selected bond lengths and bond angles (Å, °) for 1 and 2. Estimated standard deviations are given in parentheses.

Compound 1 Compound 2
Mo1-O3 1.7469(16) W1-O3 1.7612(18)
Mo1-O2 1.7469(16) W1-O2 1.7708(19)
Mo1-O1 1.8239(15) W1-O1 1.8621(18)
Mo1-O11 1.9888(14) W1-O11 1.9727(18)
Mo1-O1i 2.2329(15) W1-O1i 2.1653(18)
Mo1-N1 2.3418(17) W1-N1 2.338(2)
Na1-O4ii 2.3150(18) Na1-O4 2.312(2)
Na1-O13 2.3368(18) Na1-O12ii 2.344(2)
Na1-O5 2.3492(18) Na-O13 2.350(2)
Na1-O12iii 2.3581(18) Na1-O5 2.352(2)
Na1-O4 2.3581(18) Na1-O4iii 2.362(2)
Na···Na 7.145(21) Na···Na 3.565(2)
Na···Na 3.538(15) Na···Na 7.159(10)
O3-Mo1-O2 104.16(8) O3-W1-O2 103.25(9)
O3-Mo1-O1 101.07(7) O3-W1-O1 99.25(8)
O2-Mo1-O1 104.78(7) O2-W1-O1 104.02(9)
O3-Mo1-O11 98.62(7) O3-W1-O11 99.13(8)
O2-Mo1-O11 90.77(7) O2-W1-O11 91.44(8)
O1-Mo1-O11 150.92(7) O1-W1-O11 152.49(8)
O3-Mo1-O1i 166.12(7) O3-W1-O1i 165.18(8)
O2-Mo1-O1i 89.71(7) O2-W1-O1i 91.49(8)
O1-Mo1-O1i 75.12(7) O1-W1-O1i 75.31(8)
O11-Mo1-O1i 80.64(6) O11-W1-O1i 81.77(7)
O3-Mo1-N1 88.94(7) O3-W1-N1 88.29(8)
O2-Mo1-N1 160.79(7) O2- W1-N1 162.44(8)
O1-Mo1-N1 85.98(6) O1-W1-N1 86.80(8)
O11-Mo1-N1 73.16(6) O11-W1-N1 73.48(8)
O1i-Mo1-N1 77.55(6) O1i-W1-N1 77.74(8)
O4ii-Na1-O13 93.59(6) O4-Na1-O13 171.12(9)
O4ii-Na1-O5 91.38(6) O4-Na1-O5 91.42
O13-Na1-O5 116.14(7) O13-Na1-O5 89.29(8)
O4ii-Na1-O12iii 172.44(7) O4-Na1-O12ii 94.58(8)
O13-Na1-O12iii 93.28(6) O13-Na1-O12ii 92.67(8)
O5-Na1-O12iii 88.41(6) O5-Na1-O12ii 117.55(8)
O4ii-Na1-O4 81.60(6) O4-Na1-O4iii 80.59(8)
O13-Na1-O4 115.33(7) O13-Na1-O4iii 92.25(8)
O5-Na1-O4 128.37(7) O5-Na1-O4iii 130.73(9)
O12iii-Na1-O4 92.61(6) O12ii-Na1-O4iii 111.56(8)
  1. Symmetry transformations used to generate equivalent atoms: Symmetry code for compound 1 i) −x + 2, −y + 1, −z; ii) −x + 1, −y + 2, −z; iii) 3 −x + 1, −y + 1, −z + 1. Compound 2 i) −x + 1, −y + 1, −z; ii) −x, −y + 1, −z + 1; iii) −x, −y, −z + 2.

In the neutral dinuclear (M2O6) units the M6+ cation (Mo or W) is bonded to two terminal oxido ligands at shorter bond distances (Mo1-O3 = 1.7469(16) and Mo1-O2 = 1.7469(16) Å; W1-O3 = 1.7612(18) and W1-O2 = 1.7708(19) Å), and to two μ2-bridging O atoms at slightly longer distances (Mo1-O1 = 1.8239(15) and 2.2329(15) Å; W1-O1 = 1.8621(18) and 1.9727(18) Å) (Table 2). The linking of the (M2O6) units with two symmetry related L ligands via the alkoxide oxygen atom (O11) at an intermediate distance (Mo1-O11 = 1.9888(14) Å and W1-O11 = 1.9727(8) Å) and an amino nitrogen atom (N1) at the longest distance (Mo1-N1 = 2.3418 Å; W1-N1 = 2.338(2) Å) completes the hexa-coordination around M. The M–O/N bond angles deviate from the ideal values of 90° (75.12–104.88(7)° in 1; 73.48(7) to 104.02(9)° in 2) and 180° (150.92(7) to 166.12(7)° in 1; 152.49(6) to 165.18(8)° in 2) indicating a distortion of the {MO5N} octahedra (Table 2). The M-M distances are 3.226(12) and 3.194(6) Å in 1 and 2, respectively. The bridging nature of O1 and the coordination of L result in the formation of edge-sharing {M2O8N2} bioctahedra with the anions L acting as handles (Figure S2). In addition, each anionic ligand forms bridges to two symmetry related Na+ cations (through its atoms O12 and O13) with a Na···Na separation of 7.155(21) Å in 1 (7.159(10) Å in 2) (Figure 2, Figure S3) and is thus a μ3-bridging tetradentate ligand.

Figure 2: 
The μ3-bridging tetradentate ligand anion L– in the structure of 1 with a Na···Na separation of 7.155(21) Å (left); the μ2-bridging aqua ligand O4 in 1 with a Na···Na separation of 3.538(15) Å (right). For symmetry code See Table 2.
Figure 2:

The μ3-bridging tetradentate ligand anion L in the structure of 1 with a Na···Na separation of 7.155(21) Å (left); the μ2-bridging aqua ligand O4 in 1 with a Na···Na separation of 3.538(15) Å (right). For symmetry code See Table 2.

The Na+ cations in 1 (or 2) are in a distorted trigonal bipyramidal geometry (Figure S4) and are bonded to a terminal aqua ligand (O5), a pair of bridging aqua (O4) ligands, and two symmetry related L anions via the hydroxy groups of O12 and O13 (Figure 2). Although the Na–O distances scatter in a small range between 2.3150(8) and 2.3581(18) Å in 1 (2.312(2) to 2.362(2) in 2), the O–Na–O angles of the {NaO5} polyhedron deviate considerably from the ideal values expected for a trigonal bipyramid (Table 2). An edge-sharing {Na2O8} polyhedron is generated by two symmetry related aqua ligands (O4) linking neighboring Na+ cations and a Na···Na separation of 3.538(15) and 3.565(7) Å is observed in 1 and 2, respectively (Figure 2).

The bridging modes of the aqua ligand O4 and L organize the Na+ ions into an infinite zig-zag chain extending along the c axis, with alternating Na···Na separations of 3.538(15) and 7.155(21) Å in 1 and 3.565(7) and 7.159(10) Å in 2 (Figure 3). Each Na+ cation in the chain is bonded to a terminal aqua ligand O5 and alternating pairs of Na+ are bridged by a pair of aqua ligands (O4) and a pair of organic ligands (Figure 3).

Figure 3: 
View of the zig-zag chains of the Na+ cations with alternating Na···Na separations in 1 (top); the same arrangement with terminal and bridging aqua ligands and –OH groups of L to show the full chains. For clarity the H atoms attached to C and N are not shown (bottom).
Figure 3:

View of the zig-zag chains of the Na+ cations with alternating Na···Na separations in 1 (top); the same arrangement with terminal and bridging aqua ligands and –OH groups of L to show the full chains. For clarity the H atoms attached to C and N are not shown (bottom).

The ligand L functions as a handle attached on either side of the {Na2O8} polyhedra. Thus the chain in 1 (or 2) consists of alternating bis(μ-aqua) bridged (Na2(H2O)4)2+ units and L anions (Figure 4). The parallel chains are interlinked due to the binding of the alkoxide O11 atom and the amine N1 atom of L to the centrosymmetric {M2O6} moieties extending the connectivity along the a axis and resulting in a two-dimensional heterometallic polymer (Figure 5, Figure S5).

Figure 4: 
Parallel chains of Na+ ions showing alternating {Na2O8} polyhedra and bridging ligand anions L–.
Figure 4:

Parallel chains of Na+ ions showing alternating {Na2O8} polyhedra and bridging ligand anions L.

Figure 5: 
The binding of L− in the chains with (M2O6) clusters results in 2D connectivity in crystals of 1. For the polyhedral representation of the (M2O8N2) bioctahedra and {Na2O8} polyhedra and the 2D connectivity in 2 see Figure S5.
Figure 5:

The binding of L in the chains with (M2O6) clusters results in 2D connectivity in crystals of 1. For the polyhedral representation of the (M2O8N2) bioctahedra and {Na2O8} polyhedra and the 2D connectivity in 2 see Figure S5.

The constituents in the crystal structure of 1 (or 2) viz. (M2O6), L, (Na2(H2O)4)2+ and water molecules are interlinked with the aid of three varieties of hydrogen bonding interactions (Table S2). A search of the Cambridge Structural Database (CSD) [52] reveals many examples of structurally characterized compounds containing the amino alcohol ligand LH (Table 3). In these compounds LH exhibits different bridging modes viz. μ8-, μ7-, μ5-, μ4-, which are classified under five categories, namely Type A, Type B, Type C, Type D and Type E. The μ3-bridging tetradentate binding mode is observed only in the (M2O6) containing compounds 1 and 2.

Table 3:

Binding modes of the amino alcohol (LH) ligand listed in the CSD.

No. Compound Space group Refcode
Type A: binding mode of LH is μ 8 3 : ƞ 3 : ƞ 3 : ƞ 1 ; decadentate (O, N donor)
1 [TBA]5[Mo6O18(L)MnMo6O18(LH)] P 1 DUWSEV
2 {Ag3[MnMo6O18(LH)2(DMSO)5]·3(DMSO)} n P21/n ODERAR
3 {Ag3[MnMo6O18{LH}2(DMSO)6(CH3CN)2]·DMSO} n P21/n ODEREV
Type B: binding mode of LH is μ 7 3 : ƞ 3 : ƞ 3 ; nonadentate (O donor)
4 [TMA]2[GaMo6O18(OH)3(LH2)]·7H2O P 1 YOSMOK
5 [TBA]4{[(LH2)CrMo6O18(OH)4]2·4[TBA]Br2[NH4]Br·15H2O P2 1 ZABKAR
6 Na3[MnMo6O18(LH)2]·3DMF·4H2O P 1 RAKFOB
7 Na3[MnMo6O18(LH)2]·4DMF·5H2O P 1 VEHSOS
8 Na3[MnMo6O18(LH)2]·10DMF P 1 VEHSUY
9 (TMA)Na5[MnMo6O18(LH)2]·14DMF P 1 RAKFER
10 (TMA)3[MnMo6O18(LH)2]·3DMF C2/c RAKGES
11 (TEA)3[MnMo6O18(LH)2]·DMF C2/m RAKGAO
12 (TEA)3[MnMo6O18(LH)2]·2DMF C2/c RAKFUH
13 [TBA]3[(LH)CrMo6O18(OH)3][TBA]Br·2H2O P21 ROQVOK02
14 [TBA]2[CoMo6O18{(LH) }2]·DMA C2/c TEPJEG
15 [TBA]3[(LH)2CrMo6O18(OH)3]·6H2O P21/n ROQTAU
16 [TBA]3[MnMo6O18(LH)2] Pcmn OJEHEQ
17 [TBA]3[FeMo6O18(LH)2] C2/m OJEHIU
18 [TBA]4 {(LH)CrMo6O18(OH)3}·Br·2H2O P21 ROQVOK01
19 [TBA]4{(LH)MnMo6O18 (OH)3}·Br P21 FUDDUF02
20 [TBA]6{[C2H5C(CH2O)3]CrMo6O18(LH)}2·3DMF P212121 FUDFER02
21 [TBA]6{[C2H5C(CH2O)3]MnMo6O18(LH)}2·4DMF P212121 FUDFIV02
22 [TBA]6{[C2H5C(CH2O)3]AlMo6O18(LH)}2·3DMF P212121 FUDFOB
23 (NH4)4{[LH]2CuMo6O18} P42/n DIRTOQ
Type C: binding mode of LH is μ 5 - ƞ 2 3 : ƞ 3 octadentate (O donor)
24 Na[TMA]2[FeMo6O18(OH)3(LH2)](OH)·6H2O P 1 YOSMUQ
25 γ (TBA)3[(LH)2CrMo6O18(OH)4] Br ·5DMF P1 TUFQOC
26 [TBA]2[H] [(LH)CrMo6O18(OH)3]·3DMF·2H2O P 1 TUFPUH01
Type D: binding mode of LH is μ 4 3 : ƞ 2 : ƞ 2 ; heptadentate (O donor)
27 [TBA]4{[(LH2)]CrMo6O18(OH)4}2·4[TBA]Br·2[NH4]Br·14H2O P21 ZUZVOH
28 β-{Cr(LH)2Mo6O18} C2/c BAYTED
29 [NH4] β-{(LH2)2MnMo6O18} C2/c TIWMEU
Type E: Binding mode of LH is μ 3 1 : ƞ 1 : ƞ 1 : ƞ 1 ; tetradentate (O, N donor)
30 [M2O6(L)2Na2(H2O)4]·2H2O P 1 This work
31 [W2O6(L)2Na2(H2O)4]·2H2O P 1 OWIYIF

This work
  1. TBA = tetrabutylammonium; DMSO = dimethylsulfoxide; TMA = tetramethylammonium; DMF = dimethylformamide; TEA = tetraethylammonium; DMA = dimethylacetamide.

3.2 Synthetic aspects, spectral characteristics and thermal studies of 1 and 2

The reaction of an aqueous solution of Na2MoO4·2H2O (or Na2WO4·2H2O) with LH in 1:2 M ratio at room temperature (RT) results in the facile formation of 1 (or 2) containing the neutral dinuclear (M2O6) unit. The conversion of two (MO4)2– units to a (M2O6) moiety requires four H+ ions to remove the oxido ligands as water (Eq. (1)), which are provided by the deprotonation of one of the –OH groups of LH by the alkaline (MO4)2– solution as shown below. The L anion thus formed binds to both M6+ and Na+ cations to afford 1 or 2.

(1) 2 ( MO 4 ) 2 + 4 H + ( M 2 O 6 ) + 2 H 2 O

Equation (1) can explain the requirement of two moles of LH per Mo. The use of less LH (equimolar ratio) also results in product formation but with reduced yields. The compounds thus prepared exhibit satisfactory elemental analytical data for C, H and N. A comparison of the X-ray powder pattern of 1 with that of the pattern calculated from single crystal data confirms the composition of the bulk material (Figure S6). In the case of 2, the unit cell data are in agreement with the data reported by Gumerova et al. [50].

An interesting aspect of the synthesis of 1 (or 2) is that the condensation of (MO4)2– (M = Mo or W) units occurs in alkaline solution to form products comprising a dinuclear unit. Aqueous solutions of Na2MoO4·2H2O (or Na2WO4·2H2O) and LH in 1:2 ratio are alkaline in nature and the pH value is ∼12.0. The Raman spectrum of the solution exhibits a strong signal at 897 cm−1 for Mo (928 cm−1 for W) which can be assigned to the symmetric stretching vibration (υ1) of the tetrahedral (MO4)2− species (Figure 6). Slow evaporation resulted in the formation of crystalline 1 (or 2) which exhibits an intense Raman signal at 883 cm−1 for 1 (896 cm−1 for 2) due to Mo-O (or W-O) vibrations of 1 (or 2) (Figure 6, Figure S7). The absence of the characteristic υ1 signals of (MO4)2− in the Raman spectra reveals the formation of new products. However, when compound 1 (or 2) is dissolved in water the pH is 8.73 (or 8.12). The Raman spectra exhibit a strong band at 897 cm−1 (928 cm−1) indicating that in solution 1 (or 2) dissociate to form the tetrahedral (MO4)2– species (Figure 6). A similar dissociation of 2 to orthotungstate (WO4)2– has been recently reported at a pH of 7.5 [50].

Figure 6: 
Raman spectra of a solid sample of 1 (in blue); compound 1 in water (red); an aqueous mixture containing Na2MoO4·2H2O and LH (black).
Figure 6:

Raman spectra of a solid sample of 1 (in blue); compound 1 in water (red); an aqueous mixture containing Na2MoO4·2H2O and LH (black).

In recent work [49] we have shown that the reaction of LH with in situ generated (NH4)2MoO4 (MoO3 dissolved in excess ammonia) leads to crystallization of a mixed-cation monomolybdate (NH4)(LH2)(MoO4). In contrast, the use of Na2MoO4 containing the oxophilic Na+ cation leads to the formation of the two-dimensional coordination polymer 1 in which (L) functions as a µ3-bridging tetradentate ligand.

A survey of the literature reveals that several compounds (Table 4) containing the (Mo2O6) unit have been synthesized and structurally characterized [53], [54], [55], [56], [57], [58], [59], [60], [61], [62], [63], [64], [65]. Most of these syntheses were performed either under hydrothermal conditions or at high temperatures except for 1 and 2, (TBA)2[Mo2O6(pic)2] and (nHex4N)2[Mo2O6(picOH)2] (entry nos. 6 and 7 in Table 4) which are conducted at RT. A range of Mo containing reagents other than Na2MoO4 have been employed as the starting materials including MoO3, (NH4)2MoO4, (NH4)6Mo7O24·4H2O, [MoO3(2,2′-bipy)], (TBA)4Mo8O26, and [MoO2Cl2(pypzEA)]. In the present study, we have shown that the isostructural W compound 2 can also be synthesized at RT, while in previous work it was obtained by heating an orthotungstate (WO4)2– solution containing LH acidified with HCl to 90 °C, and adjusting the pH of the reaction mixture to 7.5 after cooling to RT [50].

Table 4:

Reaction conditions for synthesis of (Mo2O6) containing compounds.

No Compound Reaction temperature/Mo reagent Ref
1 [Mo2O6(imi)6] Hydrothermal, 150 °C/(NH4)2MoO4 [53]
2 [Mo2O6(di-tBu-bpy)2] Hydrothermal,160 °C/MoO3 [54]
3 (Pr2NH2)4[Mo2O6(C2O4)2] Heated at 60 °C/(NH4)6Mo7O24·4H2O [55]
4 (iPr2NH2)4[Mo2O6(C2O4)2] Heated at 60 °C/(NH4)6Mo7O24·4H2O [55]
5 (HDabco)4[Mo2O6(C2O4)2]·2H2O Heated at 60 °C/(NH4)6Mo7O24·4H2O [55]
6 (TBA)2[Mo2O6(pic)2] RT/(TBA)4Mo8O26 [57]
7 (n-Hex4N)2[Mo2O6(picOH)2] RT/Na2MoO4·2H2O [60]
8 [Mo2O6(2,2′-bipy)] Solid-state synthesis at 300 °C/[MoO3(2,2′-bipy)] [58]
9 [Mo2O6(HpypzA)] Hydrothermal, 100 °C/[MoO2Cl2(pypzEA)] [56]
10 [Mo2O6(tr2pr)] Hydrothermal, 180 °C/(NH4)6Mo7O24·4H2O [59]
11 [Mo2O6(tr2cy)] Hydrothermal,180 °C/MoO3/Co2+Catalyst [59]
12 [Mo2O6(tr2ad)]·6H2O Hydrothermal, 180 °C/MoO3/Co2+Catalyst [59]
13 [Mo2O6(4,9-tr2dia)]·0.5H2O Hydrothermal, 180 °C/Na2MoO4·2H2O/Cu2+ catalyst [59]
14 [Mo2O6(trethbz)2]·H2O Hydrothermal/(NH4)6Mo7O24·4H2O [61]
15 [Mo2O6(trpzH)(H2O)2] H2O Reflux in water (100 °C)/MoO3 [62]
16 [Mo2O6(m-trtzH)(H2O)2] H2O Hydrothermal, 140 °C/MoO3 [62]
17 [Mo2O6(p-trtzH)] H2O Hydrothermal, 140 °C/MoO3 [62]
18 [Mo2O6(tr2ad)]·H2O Hydrothermal, 170 °C/(NH4)6Mo7O24·4H2O [62]
19 [Mo2O6(p-tr2Ph)] H2O Hydrothermal, 140 °C/(NH4)6Mo7O24·4H2O [62]
20 Cu(nic)(nicH)(H2O)2[Mo2O6] n Hydrothermal, 170 °C/MoO3 [63]
21 [Mo2O6(pyrazine)] Hydrothermal, 160 °C/MoO3 [64]
22 [Mo2O6(4,4′-bipy)] Hydrothermal, 150 °C/MoO3 [65]
23 [Mo2O6(1,2,4-triazole)] Hydrothermal, 200 °C/Na2MoO4 [65]
24 [Mo2O6(L)2Na2(H2O)4]·2H2O 1 RT/Na2MO4·2H2O This work
25 [W2O6(L)2Na2(H2O)4]·2H2O 2 RT/Na2WO4·2H2O This work
26 [W2O6(L)2Na2(H2O)4]·2H2O Heated at 90 °C/Na2WO4·2H2O [50]
  1. imi = imidazole; di-tBu-bpy = 4,4′-di-tert-butyl-2,2′-bipyridine; Pr2NH = dipropylamine; iPr2NH = diisopropylamine; Dabco = 1,4-diazabicyclo[2.2.2]octane; TBA = tetrabutylammonium; pic = picolinate; n-Hex4N; tetrahexylammonium; HpicOH = 3-hydroxypicolinic acid; 2,2′-bipy = 2,2′-bipyridine; HpypzA = [3-(pyridinium-2-yl)-1H-pyrazol-1-yl]- acetate; pypzEA = ethyl[3-(pyridin-2-yl)-1H-pyrazol-1-yl]acetate; tr2pr = bis(1,2,4-triazol-4-yl)1,3-propylene; tr2cy = 1,4-bis(1,2,4-triazol-4-yl)cyclohexane; tr2ad = 1,3-bis-(1,2,4-triazol-4-yl)adamantine; 4,9-tr2dia = 9-bis(1,2,4-triazol-4-yl)diamantine; trethbz = (S)-4-(1-phenylpropyl)-1,2,4-triazole; trpzH = 5-dimethylpyrazol-4-yl]-4H-1,2,4-triazole; m-trtzH = m-[4-3’-(tetrazol-5-yl)phenyl]-4H-1,2,4-triazole; p-trtzH = p-[4-3’-(tetrazol-5-yl)phenyl]-4H-1,2,4-triazole; p-tr2Ph = [1,4-bis-(1,2,4-triazol-4-yl)benzene; nic = nicotinate; 4,4′-bipy = 4,4′-bipyridine.

The infrared spectra of solid 1 (or 2) exhibit several peaks in the mid IR region indicating the presence of L (Figure S8). The broad absorption in the 3500 cm−1 region in 1 can be attributed to the O–H stretching vibration of water molecules. The intense bands at around 1081 and 1044 cm−1 can be assigned to the vibrations of C–O groups while the absorptions at ≈881, 855, 760 cm−1, are caused by the vibrations of the dinuclear (Mo2O6) moiety.

The thermal studies of 1 (or 2) revealed that both compounds lose crystal water molecules accompanied by an initial endothermic event at 132 °C (or 128 °C) (Figure S9). Increasing the temperature, exothermic events are observed assignable to the decomposition of L resulting in the formation of 45.0% (62.5) residual mass. The IR spectra of the residues indicate the absence of water and organic components. Although the absence of mass spectral analysis of the emitted fragments precludes a detailed discussion of the exact nature of the decomposition processes, based on isothermal mass loss studies and featureless IR spectra the composition of the residue can be assigned as M2Na2O7. The thermal behavior of 2 was also reported in [50] but without a complementary MS experiment the exact decomposition pathway cannot be described.

4 Conclusions

In this study, we describe a convenient synthetic protocol for the room temperature synthesis of the isostructural heterometallic coordination polymers [M2O63-L)2(Na2(H2O)4)]·2H2O (M = Mo 1; M = W 2) by an aqueous reaction of Na2MO4·2H2O with 2-amino-2-(hydroxymethyl)propane-1,3-diol (LH). The condensation of (MO4)2− units to form the dinuclear moiety (M2O6) occurs in alkaline medium due to deprotonation of LH. The anionic amino alkoxide ligand L exhibits a μ3-tetradentate bridging mode and functions as an organic handle for the centrosymmetric neutral (M2O6) unit and bis(μ2-aqua) bridged [Na2(H2O)4]2+ cations and organizes the Na+ cations into a chain with alternating Na···Na separations. The chains are interlinked by the binding of the (M2O6) units with the organic ligand resulting in 2D connectivity. Compound 1 is a new example of a dinuclear Mo compound based on the (Mo2O6) moiety. The synthesis of the previously reported W analogue 2 has now been performed at room temperature at a suitably adjusted pH value.

5 Supporting information

Deposition numbers CCDC 2128432 (1) CCDC 2128433 (2), contain the supplementary crystallographic data for this paper. These data are provided free of charge by the joint Cambridge Crystallographic Data Centre and Fachinformationszentrum Karlsruhe Access Structures service www.ccdc.cam.ac.uk/structures. Supplementary Data (Figures S1–S9) and Tables (Tables S1 and S2) associated with this article are available in electronic form.


Dedicated to Professor Christian Näther on the occasion of his 60th birthday.



Corresponding author: Bikshandarkoil R. Srinivasan, School of Chemical Sciences, Goa University, Goa 403206, India, E-mail:

Funding source: State of Schleswig-Holstein

Funding source: University Grants Commission, New Delhi

Acknowledgments

Financial assistance to the School of Chemical Sciences (formerly Department of Chemistry), Goa University at the level of DSA-I under the Special Assistance Programme (SAP) by the University Grants Commission, New Delhi is gratefully acknowledged. WB acknowledges financial support by the State of Schleswig-Holstein.

  1. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This work was funded by the University Grants Commission, New Delhi and the State of Schleswig-Holstein.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

1. Cotton, F. A., Wilkinson, G., Murillo, C. A., Bochmann, M. Advanced Inorganic Chemistry, 6th ed.; John Wiley & Sons, Inc.: New York, 1999.Search in Google Scholar

2. Atkins, P., Overton, T., Rourke, J., Weller, M., Armstrong, F., Hagerman, M. Shriver and Atkins Inorganic Chemistry, 5th ed.; W. H. Freeman and Company: New York, 2010; p. 463.Search in Google Scholar

3. Mascaros, J. R. G., Kortz, U. Acta Crystallogr. 2018, C74, 1180–1181; https://doi.org/10.1107/s2053229618015188.Search in Google Scholar

4. Pope, M. T., Sadakane, M., Kortz, U. Celebrating polyoxometalates. Eur. J. Inorg. Chem. 2019, 2019, 340–342; https://doi.org/10.1002/ejic.201801543.Search in Google Scholar

5. Gumerova, N. I., Rompel, A. Polyoxometalates in solution: speciation under spotlight. Chem. Soc. Rev. 2020, 49, 7568–7601; https://doi.org/10.1039/d0cs00392a.Search in Google Scholar

6. Mahnke, L. K., Stehlikova, G., Synnatschke, K., Backes, C., Näther, C., Bensch, W. An exotic layered compound consisting of interconnected arsenato-polyoxovanadate clusters: thermal and magnetic properties and liquid phase exfoliation. Chem. Nano. Mat. 2021, 7, 78–84; https://doi.org/10.1002/cnma.202000563.Search in Google Scholar

7. Horn, M. R., Singh, A., Alomari, S., Goberna-Ferrón, S., Benages-Vilau, R., Chodankar, N., Motta, N., Ostrikov, K., MacLeod, J., Sonar, P., Gomez-Romero, P., Dubal, D. Polyoxometalates (POMs): from electroactive clusters to energy materials. Energy Environ. Sci. 2021, 14, 1652–1700; https://doi.org/10.1039/d0ee03407j.Search in Google Scholar

8. Müller, A., Beckmann, E., Bögge, H., Schmidtmann, M., Dress, A. Inorganic chemistry goes protein size: a Mo368 nano-hedgehog initiating nanochemistry by symmetry breaking. Angew Chem. Int. Ed. Engl. 2002, 41, 1162–1167.10.1002/1521-3773(20020402)41:7<1162::AID-ANIE1162>3.0.CO;2-8Search in Google Scholar

9. Berzelius, J. J. Beitrag zur näheren Kenntniss des Molybdäns. Ann. Phys. 1826, 82, 369–392; https://doi.org/10.1002/andp.18260820402.Search in Google Scholar

10. Pope, M. T., Müller, A. Polyoxometalate chemistry: an old field with new dimensions in several disciplines. Angew Chem. Int. Ed. Engl. 1991, 30, 34–48; https://doi.org/10.1002/anie.199100341.Search in Google Scholar

11. Pope, M. T., Müller, A. Polyoxometalate Chemistry from Topology via Self-Assembly to Applications; Kluver Academic Publishers: New York, 2002.10.1007/0-306-47625-8Search in Google Scholar

12. Wang, S. S., Yang, G. Y. Recent advances in polyoxometalate catalyzed reactions. Chem. Rev. 2015, 115, 4893–4962; https://doi.org/10.1021/cr500390v.Search in Google Scholar

13. Nadal, L. V., Cronin, L. Design and synthesis of polyoxometalates framework materials from cluster precursors. Nat. Rev. 2017, 2, 1–15; https://doi.org/10.1038/natrevmats.2017.54.Search in Google Scholar

14. Jassal, A. K., Rana, L. K., Hundal, G. Structure directing role of amines and water molecules in the self-assembly of polyoxomolybdates. CrystEngComm 2017, 19, 2021–2035; https://doi.org/10.1039/c6ce02640k.Search in Google Scholar

15. Bijelic, A., Aureliano, M., Rompel, A. The antibacterial activity of polyoxometalates: structures, antibiotic effects and future perspectives. Chem. Commun. 2018, 54, 1153–1169; https://doi.org/10.1039/c7cc07549a.Search in Google Scholar PubMed PubMed Central

16. Mahnke, L. K., Kondinski, A., Warzok, U., Näther, C., van Leusen, J., Schalley, C. A., Monakhov, K. Y., Kögerler, P., Bensch, W. Configurational isomerism in polyoxovanadates. Angew. Chem. Int. Ed. 2018, 57, 2972–2975; https://doi.org/10.1002/anie.201712417.Search in Google Scholar PubMed

17. Bijelic, A., Aureliano, M., Rompel, A. Polyoxometalates as potential next-generation metallodrugs in the combat against cancer. Angew. Chem. Int. Ed. 2019, 58, 2980–2999; https://doi.org/10.1002/anie.201803868.Search in Google Scholar PubMed PubMed Central

18. Janusson, E., de Kler, N., Vila`-Nadal, L., Long, D., Cronin, L. Synthesis of polyoxometalate clusters using carbohydrates as reducing agents leads to isomer-selection. Chem. Commun. 2019, 55, 5797–5800; https://doi.org/10.1039/c9cc02361e.Search in Google Scholar PubMed

19. Evans, H. T., Gatehouse, B. M., Leverett, P. Crystal structure of the heptamolybdate(VI) (paramolybdate) ion, [Mo7O24]6−, in the ammonium and potassium tetrahydrate salts. J. Chem. Soc. Dalton Trans. 1975, 505–514; https://doi.org/10.1039/dt9750000505.Search in Google Scholar

20. Ohashi, Y., Yanagi, K., Sasada, Y., Yamase, T. Crystal structure and photochemistry of isopolymolybdates. I. The crystal structures of hexakis(propylammonium) heptamolybdate(VI) trihydrate and hexakis(isopropylammonium) heptamolybdate(VI) trihydrate. Bull. Chem. Soc. Jpn. 1982, 55, 1254–1260; https://doi.org/10.1246/bcsj.55.1254.Search in Google Scholar

21. Gutierrez-Zorrilla, J. M., Roman, P., Esteban-Calderon, C., Martinez-Ripoll, M., Garcia-Blanco, S. Syntheses and crystal structures of aminopyridinium polymolybdates. Acta Crystallogr. 1984, 40A, C229–C231; https://doi.org/10.1107/s0108767384093107.Search in Google Scholar

22. Roman, P., Gutierrez-Zorrilla, J. M., Martinez-Ripoll, M., Garcia-Blanco, S. The crystal structure of 4- aminopyridinium heptamolybdate hexahydrate. Z. Kristallogr. 1985, 173, 283–292; https://doi.org/10.1524/zkri.1985.173.3-4.283.Search in Google Scholar

23. Roman, P., Gutierrez-Zorrilla, J. M., Luque, A., Martinez-Ripoll, M. Crystal structure and spectroscopic study of polymolybdates. The crystal structure and bonding of hexakis(n-pentylammonium) heptamolybdate(VI) trihydrate. J. Crystallogr. Spectrosc. Res. 1988, 18, 117–131.10.1007/BF01181904Search in Google Scholar

24. Roman, P., Luque, A., Aranzabe, A., Gutierrez-Zorrilla, J. M. Reactions of MoO3 with diethylenetriamine (dien): syntheses, solid-state characterization and thermal behaviour. Molecular and crystal structure of a second polymorph of (H3dien)2[Mo7O24]·4H2O. Polyhedron 1992, 11, 2027–2038; https://doi.org/10.1016/s0277-5387(00)83158-7.Search in Google Scholar

25. Roman, P., San Jose, A., Luque, A., Gutierrez-Zorrilla, J. M. Hexakis(tert-butylammonium) heptamolybdate(VI)-water (1/7). Acta Crystallogr. 1994, 50, 1031–1034; https://doi.org/10.1107/s0108270193012831.Search in Google Scholar

26. Kortz, U., Pope, M. T. Cs6[Mo7O24].7H2O. Acta Crystallogr. 1995, C51, 1717–1719; https://doi.org/10.1107/s010827019500494x.Search in Google Scholar

27. Niu, J. Y., You, X. Z., Fun, H. K., Zhou, Z. Y., Yip, B. C. Synthesis, crystal structure and photochromism of tri (N,N,N′,N′tetramethylethylendiammonium) heptamolybdate(VI) tetrahydrate. Polyhedron 1996, 15, 1003–1008; https://doi.org/10.1016/0277-5387(95)00191-t.Search in Google Scholar

28. Gili, P., Lorenzo-Luis, P. A., Mederos, A., Arrieta, J. M., Germain, G., Castineiras, A., Carballo, R. Crystal structures of two new heptamolybdates and of a pyrazole incorporating a γ-octamolybdate anion. Inorg. Chim. Acta 1999, 295, 106–114; https://doi.org/10.1016/s0020-1693(99)00329-1.Search in Google Scholar

29. Pavani, K., Ramanan, A. Influence of 2-aminopyridine on the formation of molybdates under hydrothermal conditions. Eur. J. Inorg. Chem. 2005, 2005, 3080–3087; https://doi.org/10.1002/ejic.200500092.Search in Google Scholar

30. Coue, V., Dessapt, R., Bujoli-Doeuff, M., Evain, M., Jobic, S. Synthesis, characterization, and photochromic properties of hybrid organic-inorganic materials based on molybdate, DABCO, and piperazine. Inorg. Chem. 2007, 46, 2824–2835; https://doi.org/10.1021/ic0621502.Search in Google Scholar PubMed

31. Wutkowski, A., Srinivasan, B. R., Naik, A. R., Schütt, C., Näther, C., Bensch, W. Synthesis, structure and photochemistry of a novel organic heptamolybdate-monomolybdate. Eur. J. Inorg. Chem. 2011, 2011, 2254–2263; https://doi.org/10.1002/ejic.201001154.Search in Google Scholar

32. Shivaiah, V., Kumar, N. T., Das, S. K. A gas-liquid interface synthesis in polyoxometalate chemistry: potential bag filter for volatile organic amines. J. Chem. Sci. 2018, 130, 1–10; https://doi.org/10.1007/s12039-018-1435-2.Search in Google Scholar

33. Khandolkar, S. S., Raghavaiah, P., Srinivasan, B. R. Synthesis, characterization and photochemistry of a new heptamolybdate supported magnesium-aqua coordination complex. J. Chem. Sci. 2015, 127, 1581–1588; https://doi.org/10.1007/s12039-015-0918-7.Search in Google Scholar

34. Srinivasan, B. R., Morajkar, S. M. Sodium paramolybdate revisited. Indian J. Chem. 2016, 55A, 676–680.10.1002/chin.201644017Search in Google Scholar

35. Khandolkar, S. S., Näther, C., Bensch, W., Srinivasan, B. R. Synthesis and structures of two new lithium ̶ heptamolybdates. J. Coord. Chem. 2016, 69, 1166–1178; https://doi.org/10.1080/00958972.2016.1159679.Search in Google Scholar

36. Khandolkar, S. S., Naik, A. R., Näther, C., Bensch, W., Srinivasan, B. R. Synthesis, crystal structure and photochemistry of hexakis(butan-1-aminium) heptamolybdate(VI) tetrahydrate. J. Chem. Sci. 2016, 128, 1737–1744; https://doi.org/10.1007/s12039-016-1168-z.Search in Google Scholar

37. Srinivasan, B. R., Morajkar, S. M., Khandolkar, S. S., Näther, C., Bensch, W. Synthesis, structure and properties of a hexarubidium heptamolybdate with bridging aqua ligands. Indian J. Chem. 2017, 56A, 601–609.Search in Google Scholar

38. Srinivasan, B. R., Khandolkar, S. S., Morajkar, S. M. Structural identification of two differently coordinated heptamolybdate ligands in a hexamagnesium compound. Indian J. Chem. 2020, 59A, 517–525.Search in Google Scholar

39. Pannuru, P., Rani, A., Venkatesu, P., Lee, M. J. The effects of biological buffers TRIS, TAPS, TES on the stability of lysozyme. Int. J. Biol. Macromol. 2018, 112, 720–727; https://doi.org/10.1016/j.ijbiomac.2018.01.203.Search in Google Scholar PubMed

40. Bastos, I. N., Platt, G. M., Andrade, M. C., Soares, G. D. Theoretical study of tris and bistris effects on simulated body fluids. J. Mol. Liq. 2008, 139, 121–130; https://doi.org/10.1016/j.molliq.2007.12.003.Search in Google Scholar

41. Soriano, A. N., Cabahug, D. I. V., Li, M. H. Thermophysical property characterization of tris(hydroxymethyl)aminomethane. J. Chem. Thermodyn. 2011, 43, 186–189; https://doi.org/10.1016/j.jct.2010.08.016.Search in Google Scholar

42. Ibrahim-Hashim, A., Abrahams, D., Enriquez-Navas, P. M., Luddy, K., Gatenby, R. A., Gillies, R. J. Tris-base buffer: a promising new inhibitor for cancer progression and metastasis. Cancer Med. 2017, 6, 1720–1729; https://doi.org/10.1002/cam4.1032.Search in Google Scholar PubMed PubMed Central

43. Long, D. L., Kögerler, P., Farrugia, L. J., Cronin, L. Restraining symmetry in the formation of small polyoxomolybdates: new building blocks of unprecedented topology resulting from shrink-wrapping [H2Mo16O52]10− type clusters. Angew. Chem. Int. Ed. 2003, 42, 4180–4183; https://doi.org/10.1002/anie.200351615.Search in Google Scholar PubMed

44. Fernández-Navarro, L., Nunes-Collado, A., Artetxe, B., Ruiz-Bilbao, E., San Felices, L., Reinoso, S., San José Wéry, A., Gutiérrez-Zorrilla, J. M. Isolation of the elusive heptavanadate anion with trisalkoxide ligands. Inorg. Chem. 2021, 60, 5442–5445; https://doi.org/10.1021/acs.inorgchem.1c00448.Search in Google Scholar PubMed

45. Blazevic, A., Al-Sayed, E., Roller, A., Giester, G., Rompel, A. Tris−functionalized hybrid Anderson polyoxometalates: synthesis, characterization, hydrolytic stability and inversion of protein surface charge. Chem. Eur. J. 2015, 21, 4762–4771; https://doi.org/10.1002/chem.201405644.Search in Google Scholar PubMed

46. Gumerova, N. I., Roller, A., Rompel, A. Synthesis and characterization of the first Ni(II)-centered single-side tris-functionalized Anderson-type polyoxomolybdate. Eur. J. Inorg. Chem. 2016, 2016, 5507–5511; https://doi.org/10.1002/ejic.201601198.Search in Google Scholar

47. Gumerova, N. I., Roller, A., Rompel, A. [Ni(OH)3W6O18(OCH2)3CCH2OH]4−: the first tris-functionalized Anderson-type heteropolytungstate. Chem. Commun. 2016, 52, 9263–9266; https://doi.org/10.1039/c6cc04326g.Search in Google Scholar PubMed PubMed Central

48. Gumerova, N. I., Caldera Fraile, T., Roller, A., Giester, G., Pascual-Borràs, M., Ohlin, C. A., Rompel, A. The direct single- and double-side triol-functionalization of the mixed type Anderson polyoxotungstate [Cr(OH)3W6O21]6−. Inorg. Chem. 2019, 58, 106–113; https://doi.org/10.1021/acs.inorgchem.8b01740.Search in Google Scholar PubMed PubMed Central

49. Srinivasan, B. R., Morajkar, S. M., Khandolkar, S. S., Gobre, V. V., Apreyan, R. A. Synthesis, crystal structure and properties of a noncentrosymmetric tetraoxidomolybdate(VI). J. Mol. Struct. 2020, 1204, 127518; https://doi.org/10.1016/j.molstruc.2019.127518.Search in Google Scholar

50. Gumerova, N. I., Prado-Roller, A., Rambaran, M. A., André Ohlin, C., Rompel, A. The smallest polyoxotungstate retained by TRIS-stabilization. Inorg. Chem. 2021, 60, 12671–12675; https://doi.org/10.1021/acs.inorgchem.1c01188.Search in Google Scholar PubMed PubMed Central

51. Sheldrick, G. M. Crystal structure refinement with Shelxl. Acta Crystallogr. 2015, C71, 3–8; https://doi.org/10.1107/s2053229614024218.Search in Google Scholar

52. Groom, C. R., Bruno, I. J., Lightfoot, M. P., Ward, S. C. The Cambridge structural database. Acta Crystallogr. 2016, B72, 171–179; https://doi.org/10.1107/s2052520616003954.Search in Google Scholar

53. Lin, Y. W., Tong, Y., Yang, C., Lin, Y. R. Synthesis, X-ray structure, spectroscopic properties and theoretical investigations of dinuclear imidazole molybdenum(VI) complex with long Mo=O bonds. Inorg. Chem. Commun. 2009, 12, 252–254; https://doi.org/10.1016/j.inoche.2009.01.001.Search in Google Scholar

54. Amarante, T. R., Neves, P., Paz, F. A. A., Pillinger, M., Valente, A. A., Gonçalves, I. S. A dinuclear oxomolybdenum(VI) complex, [Mo2O6(4,4′-di-tert-butyl-2,2′-bipyridine)2] displaying the MoO2(μ-O)2MoO2}0 core, and its use as a catalyst in olefin epoxidation. Inorg. Chem. Commun. 2012, 20, 147–152; https://doi.org/10.1016/j.inoche.2012.02.038.Search in Google Scholar

55. Sarr, B., Mbaye, A., Diop, C. A. K., Melin, M. S. F., Hellwig, P., Maury, F., Senocq, F., Guionneau, P. One pot-synthesis of the fourth category of dinuclear molybdenum(VI) oxalate series: structure and study of thermal and redox properties. Inorg. Chim. Acta 2019, 491, 84–92; https://doi.org/10.1016/j.ica.2019.03.037.Search in Google Scholar

56. Figueiredo, S., Gomes, A. C., Neves, P., Amarante, T. R., Paz, F. A. A., Soares, R., Lopes, A. D., Valente, A. A., Pillinger, M., Goncalves, I. S. Synthesis, structural elucidation, and application of a pyrazolylpyridine−molybdenum oxide composite as a heterogeneous catalyst for olefin epoxidation. Inorg. Chem. 2012, 51, 8629–8635; https://doi.org/10.1021/ic301405r.Search in Google Scholar PubMed

57. Liang, H., Chen, X. B., Chen, Z. F., Hu, R. X. One-dimensional anion-chain containing [Mo2O6(pic)2]2− unit connected by unclassical hydrogen bon. Chin. J. Chem. 2002, 20, 650–656.10.1002/cjoc.20020200706Search in Google Scholar

58. Bruno, S. M., Nogueira, L. S., Gomes, A. C., Valente, A. A., Gonçalves, I. S., Pillinger, M. High-yield synthesis and catalytic response of chainlike hybrid materials of the [(MoO3)m(2,2′–bipyridine)n] family. New J. Chem. 2018, 42, 16483–16492; https://doi.org/10.1039/c8nj02668h.Search in Google Scholar

59. Lysenko, A. B., Senchyk, G. A., Lincke, J., Lassig, D., Fokin, A. A., Butova, E. D., Schreiner, P. R., Krautscheid, H., Domasevitch, K. V. Metal oxide-organic frameworks (MOOFs), a new series of coordination hybrids constructed from molybdenum(VI) oxide and bitopic 1,2,4-triazole linkers. Dalton Trans. 2010, 39, 4223–4231; https://doi.org/10.1039/b922732f.Search in Google Scholar PubMed

60. Quintal, S. M. O., Nogueira, H. I. S., Carapuça, H. M., Félix, V., Drew, M. G. B. Polynuclear molybdenum and tungsten complexes of 3-hydroxypicolinic acid and the crystal structures of (nBu4N)2[Mo4O12(picOH)2] and (nHex4N)2[Mo2O6(picOH)2]. J. Chem. Soc., Dalton Trans. 2001, 3196–3201; https://doi.org/10.1039/b103348b.Search in Google Scholar

61. Lysenko, A. B., Senchyk, G. A., Domasevitch, K. V., Kobalz, M., Krautscheid, H., Cichos, J., Karbowiak, M., Neves, P., Valente, A. A., Goncalves, I. S. Triazolyl, imidazolyl, and carboxylic acid moieties in the design of molybdenum trioxide hybrids: photophysical and catalytic behavior. Inorg. Chem. 2017, 56, 4380–4394; https://doi.org/10.1021/acs.inorgchem.6b02986.Search in Google Scholar PubMed

62. Lysenko, A. B., Senchyk, G. A., Domasevitch, K. V., Hauser, J., Fuhrmann, D., Kobalz, M., Krautscheid, H., Neves, P., Valente, A. A., Goncalves, I. S. Synthesis and structural elucidation of triazolylmolybdenum(VI)oxide hybrids and their behavior as oxidation catalysts. Inorg. Chem. 2015, 54, 8327–8338; https://doi.org/10.1021/acs.inorgchem.5b01007.Search in Google Scholar PubMed

63. Gschwind, F., Jansen, M. Synthesis and characterization of a new infinite 1D polyoxomolybdate polymer further connected via Cu(I) nicotinate subunits. Z. Naturforsch. 2011, 66b, 351–354; https://doi.org/10.1515/znb-2011-0402.Search in Google Scholar

64. Xu, Y., Lu, J., Goh, N. K. Hydrothermal assembly and crystal structures of three novel open frameworks based on molybdenum(VI) oxides. J. Mater. Chem. 1999, 9, 1599–1602; https://doi.org/10.1039/a901592b.Search in Google Scholar

65. Hagrman, P. J., LaDuca, R. L.Jr., Koo, H. J., Rarig, R. S.Jr., Haushalter, R. C., Whangbo, M. H., Zubieta, J. Ligand influences on the structures of molybdenum oxide networks. Inorg. Chem. 2000, 39, 4311–4317; https://doi.org/10.1021/ic000496l.Search in Google Scholar PubMed


Supplementary Material

The online version of this article offers supplementary material (https://doi.org/10.1515/znb-2021-0183).


Received: 2021-12-15
Accepted: 2021-12-22
Published Online: 2022-01-05
Published in Print: 2022-05-25

© 2021 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 14.5.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2021-0183/html
Scroll to top button