You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/257585704

Crystal Structure of a New Ammonium Potassium Selenomolybdate


Compound K3.31(NH4)0.69[Se2Mo5O21]·2H2O

Article  in  Journal of Cluster Science · December 2012


DOI: 10.1007/s10876-012-0508-5

CITATIONS READS

3 78

4 authors:

Meriem Ayed Ichraf Nagazi


UNIVERSITY OF JEDDAH Faculty of Sciences and Arts Khulais University of Monastir
11 PUBLICATIONS   17 CITATIONS    10 PUBLICATIONS   22 CITATIONS   

SEE PROFILE SEE PROFILE

Brahim Ayed Amor Haddad


University of Monastir Higher Institute of Applied Sciences and Technology, Mahdia, Tunisia.
39 PUBLICATIONS   90 CITATIONS    86 PUBLICATIONS   193 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

synthesis of polyoxometallates View project

Polyoxometalates synthesis View project

All content following this page was uploaded by Meriem Ayed on 02 October 2016.

The user has requested enhancement of the downloaded file.


J Clust Sci (2012) 23:1133–1142
DOI 10.1007/s10876-012-0508-5

ORIGINAL PAPER

Crystal Structure of a New Ammonium Potassium


Selenomolybdate Compound
K3.31(NH4)0.69[Se2Mo5O21]2H2O

Meriem Ayed • Ichraf Nagazi • Brahim Ayed •

Amor Haddad

Received: 28 May 2012 / Published online: 9 August 2012


Ó Springer Science+Business Media, LLC 2012

Abstract An inorganic compound formulated as K3.31(NH4)0.69[Se2Mo5O21]


2H2O has been synthesized by conventional solution method and characterized by
scanning electron microscopy, IR, UV-Vis spectroscopies behaviors. The structure
of the title compound has been determined from a single-crystal X-ray diffraction.
It crystallizes in the monoclinic space group P21/n, with a = 9.9371(2) Å,
b = 23.3545(2) Å, c = 10.5179(2) Å, b = 114.12(3)°, V = 2227.7(5) Å3 and
Z = 4. It was revealed that the Strandberg-type polyoxoselenomolybdate cluster can
be considered as a ring formed by five distorted edge- and corner-sharing MoO6
octahedra, capped on both poles by a selenate pyramids sharing three vertices with
the ring molybdenum centers. The Strandberg clusters are connected with ammo-
nium ions and water molecules through hydrogen-bonding interactions which
ensure the cohesion of the structure into a three-dimensional network.

Keywords Polyoxometalate  X-ray diffraction  Crystal structure  Infrared


spectroscopy

Introduction

Over the past decades, inorganic materials based on polyoxometalates through


crystal engineering have been attracting considerable interest as a promising new
class of materials not only because of the structural and topological novelty of such
engineered solids, but also due their potential applications in different areas such as
catalysis, medicine, electrical conductivity, magnetism and photochemistry [1–9].
To date, many examples of extended architectures in which well-characterized

M. Ayed  I. Nagazi  B. Ayed (&)  A. Haddad


Laboratoire de Matériaux et Cristallochimie (LMC), Département de Chimie,
Faculté des Sciences Monastir, Monastir, Tunisia
e-mail: brahimayed@yahoo.fr

123
1134 M. Ayed et al.

POMs, such as Keggin [10–13], Wells-Dawson [14, 15], Silverton [16] and
Lindquist [17] type polyoxoanions serve as the building blocks have been
successfully reported. However, in contrast, the role of Strandberg-type polyoxoa-
nions as the inorganic building units to construct such extended frameworks has not
been so extensively studied. Hence, our current synthetic strategy is to acquire new
extended compounds constructed from Strandberg-type polyoxoanions in the
presence of alkali metal cations. Until very recently, the structures of the Strandberg
type were mainly known by the phosphate groups [18–21]. Analogous structures
have been found are similar for [Se2Mo5O21]4- [22] and [Se2Mo5O21]4- [23–27]. In
the present paper, we successfully prepared the three dimensional (3D) structure
based on Strandberg-type polyoxoanions. The compound is built up of
[Se2Mo5O21]4- clusters as the structural motif covalently which are linked through
K?, NH4? ions and water molecules, hydrogen-bonding interactions form the 3D
supramolecular structure.

Experimental

Materials and Measurements

All chemicals were purchased commercially and used without any further purification.
Qualitative analysis by scanning electron microscopy (SEM) probe was performed
using an apparatus of the type JOEL JSM-5400 (JOEL Ltd, Tokyo, Japan). Infrared
spectrum was recorded at a room temperature on a Nicolet 470 FTIR spectrophotometer
as KBr pellets in the 4,000–400 cm-1 region. UV–Vis spectrum was measured using a
Perkin-Elmer Lambda 19 spectrophotometer in the 180–800 nm range. The TGA-DTA
thermograms were obtained with 4.99 mg. Samples were placed in an open platinum
crucible and heated, under air, from room temperature to 450 °C with 3 °C/min heating
rate; an empty crucible was used as reference.

Chemical Preparation

Single crystals of K3.31(NH4)0.69[Se2Mo5O21]2H2O were obtained from an aqueous


solution. At room temperature 0.28 g K2CO3 (Fluka, 99 %), 2.47 g (NH4)6Mo7O24
4H2O (Panreac, 99 %) and 0.37 g SeO2 (Aldrich, 98 %) were slowly added to 40 ml of
water. The resulting mixture was stirred and the pH was adjusted to 3.5 with acetic acid.
The solution was kept for 6 days at ambient conditions, and then colorless crystals
suitable for X-ray crystallography were obtained. The qualitative analysis by SEM
spectroscopy of one of the colorless crystals obtained revealed the presence of Mo, Se,
N, K and O elements (Fig. 1). Elem. Anal. Found: Mo, 41.45; Se, 13.78; K, 11.32; N,
0.80 (%). Calcd: Mo, 41.68; Se, 13.72; K, 11.23; N, 0.86 (%).

X-Ray Data Collection

A suitable transparent single crystal with dimensions 0.2 9 0.1 9 0.1 mm3 was
selected for the structure determination and refinement. The X-ray data collection

123
Crystal Structure 1135

Fig. 1 SEM spectrum of K3.31(NH4)0.69[Se2Mo5O21]2H2O

Table 1 Crystal data and structure refinement for K3.31(NH4)0.69[Se2Mo5O21]2H2O


Chemical formula K3.31(NH4)0.69[Se2Mo5O21]2H2O
Formula weight 1151.52 g mol-1
Crystal system Monoclinic
Space group P21/n
a (Å) 9.9371(2)
b (Å) 23.3545(2)
c (Å) 10.5179(2)
b (°) 114.12(3)
V (Å3) 2227.7(5)
Z 4
T (K) 293(2)
DC (g/cm3) 3.437
lMo Ka (mm-1) 6.715
R1 [I [ 2r(I)] 0.0340
WR2 0.0834

was carried out on Enraf-Nonius CAD-4 four circle diffractometer using with
Mo Ka radiation (k = 0.71069 Å). The crystal structure was solved by direct
method using SHELXS-97 [28]. Full-matrix F2 least-squares refinement and
subsequent Fourier synthesis procedures were performed by SHELXL-97 [29]
program included in the WINGX software package [30]. The molybdenum,
selenium and potassium atoms were located and the remaining oxygen atoms were
found from successive difference Fourier maps. Refinement of all atoms, except of
nitrogen atom led to R = 0.053 and wR = 0.134. Moreover the K4–O distances as

123
1136 M. Ayed et al.

Fig. 2 An ORTEP drawing of the asymmetric unit of K3.31(NH4)0.69[Se2Mo5O21]2H2O

well as the displacement parameters led us to consider that the position is occupied
simultaneously by potassium K4 and nitrogen. The refinement of the occupancies
factors was then performed leading to 69 and 31 % for N and K4, respectively, and
to an improvement of the agreement factors: R = 0.034 and wR = 0.083; the
corresponding formula is thus K3.31(NH4)0.69[Se2Mo5O21]2H2O. The positions of
the hydrogen atoms attached to oxygen water and nitrogen were determined from a
difference Fourier map and were refined isotropically. The crystal data and structure
solution and refinement details are given in Table 1. The structural figures were
carried out with Diamond 2.1 supplied by Crystal Impact [31].

Results and Discussion

Structure Description and Discussion

The asymmetric unit of the crystal structure consists of 3.31 K? cations, 0.69 NH4?
cations two water molecules and one [Se2Mo5O21]4- cluster anion. A projection of

123
Crystal Structure 1137

Table 2 Hydrogen-bonding geometry (Å, °)


D-HA D-H (Å) HA (Å) DA (Å) D-HA (°)

OW1–H1W1–O20 0.865 2.067 2.888 158.23


OW1–H2W1–OW2 0.874 2.286 3.160 179.58
OW2–H1W2–OW1 0.862 2.298 3.160 178.96
OW2–H2W2–O13 0.860 2.298 3.158 178.62
OW2–H2W2–O15 0.860 2.630 3.055 111.67
N–H1N–O8 0.893 2.498 3.082 123.45
N–H2N–OW1 1.027 1.987 2.950 154.98
N–H3N–O8 1.056 2.051 2.845 129.85
N–H4N–OW2 1.141 1.872 2.861 142.33
N–H4N–O1 1.141 2.440 3.033 110.41

Fig. 3 The potassium environments

the structure, showing the displacement ellipsoids, is presented in Fig. 2. The


[Se2Mo5O21]4- Strandberg-type cluster has approximate C2 symmetry, the axis
passing through the unique corner-sharing oxygen (O3) atom and the opposite
molybdenum (Mo5) center. In the structure of the compound [Se2Mo5O21]4- cluster
formed by close packing of oxygen atoms with Mo and Se atoms in the distorted
octahedral and trigonal pyramidal respectively and is similar to those reported in
MoVI
5 clusters with SO4
2-
and PO43- [32–36]. However, the latter anions have
hetero atom P’s tetrahedrally coordinated by four oxygen atoms; that is, each
phosphorus has one unshared atom, while the trigonal pyramidal SeO3 has no such
terminal oxygen. The bond lengths within the anion are listed in Fig. 2.
Each selenite subunit shares three oxo-groups with the molybdate ring. In turn,
one of these oxo-groups adopts the l2-bridging mode, linking one molybdenum site
and the selenium, and the other two adopts the l3-bridging mode, linking two
molybdenum sites and the selenium. In the compound, Mo–O distances were in the
range of 1.706(5)–1.731(4) Å for terminal oxygen (Ot), 1.912(4)–1.979(4) Å for Ob
bonded to two Mo atoms, and 2.203(4)–2.381(4) Å for Oc bonded to one Mo atom
and one Se atom.
These values are almost equal to the corresponding lengths in the [P2Mo5O23]6-
anions [37–39]. The trigonal pyramidal SeO3 exhibit distorted geometry, with Se–O
bond distances ranging from 1.659(5)–1.736(4) Å and O–Se–O bond angles of
between 99.06(2)° and 102.76(2)°.

123
1138 M. Ayed et al.

Fig. 4 A packing view of compound along the a axis, showing the arrangement of the [Se2Mo5O21]4-
clusters, potassium and ammonium cations and the lattice water molecules in the crystal structure. The
dashed lines denote the hydrogen interactions

The calculated average values of the distortion indices [40] corresponding to the
different angles and distances in SeO3 pyramids and MoO6 octahedra, show that
these groups exhibit a compact assembly of oxygen atoms in which selenium atoms
are slightly displaced from the center of gravity then molybdenum atoms
[DI(OSeO) = 0.007–0.010; DI(SeO) = 0.010–0.016; DI(OMoO) = 0.20–0.21
and DI(MoO) = 0.093–0.117].
BVS calculations [41] revealed that all the molybdenum atoms have valence
sums ranging from 5.858 to 5.9319, with an average value of 5.908, close to the
ideal value of 6 for MoVI. The calculated average value of selenium (?IV), oxygen
(-II) and potassium (?I) atoms are, respectively, 4.060, 1.970 and 0.964.
The K? ions occupy four crystallographically distinct sites. As often occurs for
alkali ions, the environment of this site consists of a wide range of cation–oxygen
distances, thus it is very difficult to distinguish between bonding and non-bonding
interactions. A simple criterion is to consider all distances which are shorter than the
shortest distance between K? and its nearest cation. Assuming this criterion, the K2
and K3 are linked to nine oxygen atom with K1–O ranging from 2.702(6) to
3.209(5) Å and K3–O ranging from 2.712(5) to 3.325(5) Å with an average value of
2.940(5) and 2.977(5) Å respectively. The environment of K1 consists of eight
atoms oxygen with K1–O distances scattering from 2.688(6) to 3.005(5) Å. The K4
and N atoms are statistically distributed and have partial occupancies of 0.31 and
0.69 respectively. K4 is surrounded by seven oxygen atoms with an average value of
2.991(4) Å (Fig. 3).

123
Crystal Structure 1139

Fig. 5 The two curves corresponding to the DTA and TGA analysis in argon of
K3.31(NH4)0.69[Se2Mo5O21]2H2O

The arrangement of the [Se2Mo5O21]4- clusters, K? cations NH4? and the lattice
water molecules in the crystal structure along the a axis is displayed in Fig. 4, which
shows extensive hydrogen bonding. As listed in Table 2, OW–HO, OW–HOW,
N–HN–O and N–HN–OW hydrogen bonds between the solvent water molecules and the
clusters have interatomic OO distances ranging from 2.888 to 3.160 Å and NO
distances varying between 2.845 and 3.082 Å. These hydrogen bonds hold the
components together into a 3D network and make the crystal structure of 1 more stable.

Thermal Analysis

The two curves corresponding to the DTA and TGA analysis in argon of
K3.31(NH4)0.69[Se2Mo5O21]2H2O are given in Fig. 5. The DTA curve shows that

123
1140 M. Ayed et al.

Fig. 6 FT-IR spectrum of K3.31(NH4)0.69[Se2Mo5O21]2H2O between 4,000 and 400 cm-1

this compound undertakes a series of endothermic and exothermic peaks in a wide


temperature range (190–500 °C). The most important one appears at about 450 °C.
It corresponds to a melting point of the compound, which is in good agreement with
the result obtained by the capillarity tube method. From 190 °C, the DTA curve
shows two endothermic peaks characterized by an important weight loss observed
on the TGA curve. This thermal phenomenon corresponding to the loss of two water
molecules. The peaks exothermic at 285 and 331 °C are attributed to the loss of two
SeO2 [42, 43].

Spectroscopic Characterization

Vibrational Spectrum

As shown in Fig. 6, the strong band at 905 cm-1 in the IR spectrum of the tittle
compound is due to the stretching vibration of the Mo–Ot groups of [Se2Mo5O21]4-
clusters, those at 851–835, 634 cm-1 can be attributed to m (Mo–(l-O)) and m (Mo–
O–Mo). The bands around 758 and 455 cm-1 can be assigned to the vibrations of
the selenite groups. The compound also possesses an intense band at 1625, 1455 and
1385 cm-1 are attributed to the bending vibration of water and ammonium
molecules. The symmetric and asymmetric O–H and N–H stretchings appear
respectively around 3388, 3507 and 3220 cm-1.

UV–Vis Absorption Spectrum

The UV–Vis absorption behavior of our compound was analyzed in the


190–700 nm range using an aqueous solution. The obtained spectrum (Fig. 7)
reveals an absorption band centered at 209 nm attributed to the ligand-to-metal

123
Crystal Structure 1141

Fig. 7 UV–Vis absorption spectrum of K3.31(NH4)0.69[Se2Mo5O21]2H2O

charge transfers from terminal oxygen to molybdenum center, where electrons are
promoted from the low-energy electronic states, mainly comprised of oxygen 2p
orbitals, to the high-energy states, which are mainly comprised of metal d orbitals
[44, 45].

Conclusion

It is important to remark that few Strandberg-type anions with selenite groups have
been previously described. In this paper we described the chemical preparation and
crystal structure for a new selenomolybdate K3.31(NH4)0.69[Se2Mo5O21]2H2O. This
compound crystallizes in the monoclinic system, space group P21/n and the main
geometrical feature of this structure is the existence of [Se2Mo5O21]4- anionic
clusters, these groups are linked via ionic bonds or hydrogen bonds to form 3D
frameworks.

Supplementary Material

CCDC 849017 contains the supplementary crystallographic data for this paper.
These data can be obtained free of charge at www.ccdc.cam.ac.uk/conts/
retrieving.html [Or from the Cambridge Crystallographic Data Centre (CCDC),
12 Union Road, Cambridge CB2 1EZ, UK; Fax: ?44(0)1223-336033; email:
deposit@ccdc.cam.ac.uk].

123
1142 M. Ayed et al.

References

1. A. Müller, H. Reuter, and S. Dillinger (1995). Angew. Chem. Int. Ed. Engl. 34, 2328.
2. M. T. Pope and A. Müller (1991). Angew. Chem. Int. Ed. Engl. 30, 34.
3. D. Hagrman, R. C. Haushalter, and J. Zubieta (1998). Chem. Mater. 10, 361.
4. B. B. Xu, Z. H. Peng, Y. G. Wei, and D. R. Powell (2003). Chem. Commun. 20, 2562.
5. S. L. Zheng, J. H. Yang, X. L. Yu, X. M. Chen, and W. T. Wong (2004). Inorg. Chem. 43, 830.
6. M. Sasa, K. Tanaka, X. H. Bu, M. Shiro, and M. J. Shionoya (2001). Am. Chem. Soc. 123, 10750.
7. B. Q. Ma, D. S. Zhang, S. Gao, T. Z. Jin, C. H. Yan, and G. X. Xu (2000). Angew. Chem. Int. Ed.
Engl. 39, 3644.
8. C. M. Liu, D. Q. Zhang, M. Xiong, and D. B. Zhu (2002). Chem. Commun. 13, 1416.
9. J. J. Lu, Y. Xu, N. K. Goh, and L. S. Chia (1998). Chem. Commun. 24, 2733.
10. S. Reinoso, P. Vitoria, L. Lezama, A. Luque, and J. M. Gutierrez- Zorrilla (2003). Inorg. Chem. 42,
3709.
11. Y. Lu, Y. Xu, E. B. Wang, J. Lu, C. W. Hu, and L. Xu (2005). Cryst. Growth Des. 5, 257.
12. F. Bonhomme, J. P. Larentzos, T. M. Alam, E. J. Maginn, and M. Nyman (2005). Inorg. Chem. 44,
1774.
13. M. Vasylyeva, R. Popovitz-Birob, L. J. W. Shimonc, and R. Neumanna (2003). J. Mol. Struct. 656,
27.
14. J. Y. Niu, M. L. Wei, J. P. Wang, and D. B. Dang (2004). Eur. J. Inorg. Chem. 1, 160.
15. J. P. Wang, J. W. Zhao, and J. Y. Niu (2004). J. Mol. Struct. 697, 191.
16. C. D. Wu, C. Z. Lu, H. H. Zhuang, and J. S. Huang (2002). J. Am. Chem. Soc. 124, 3836.
17. P. J. Hagrman, D. Hagrman, and J. Zubieta (1999). Angew. Chem. Int. Ed. 38, 3165.
18. Y. F. Li, W. Cui, G. D. Zhu, S. L. Qiu, Q. R. Fang, and C. L. Wang (2003). Chem. J. Chin. Univ. 24,
394.
19. A. Aranzabe, A. S. J. Wery, S. Martin, J. M. Guiterrz-Zorrilla, A. Luque, M. Martinez-Ripoll, and P.
Roman (1997). Inorg. Chim. Acta 35, 255.
20. Z. Zhao, H. H. Zhang, and C. C. Huang (2002). Chem. J. Chin. Univ. 23, 521.
21. Y. Lu, J. Lu, E. B. Wang, Y. Q. Guo, X. X. Xu, and L. Xu (2005). J. Mol. Struct. 159, 740.
22. K. Y. Matsumoto, M. Kato, and Y. Sasaki (1976). Bull. Chem. Soc. Jpn. 49, 106.
23. G.-Q. Huang, S.-W. Zhang, and M.-C. Shao (1995). Gaodeng Xuexiao Huaxue Xuebao 16, 670.
24. H. Ichida, H. Fukushima, and Y. Sasaki (1986). Nippon Kagaku Kaishi. 1521.
25. F. Kong, C.-L. Hu, X. Xu, T.-H. Zhou, and J.-G. Mao (2012). Dalton Trans. 41, 5687.
26. M.-L. Feng and J.-G. Mao (2004). Eur. J. Inorg. Chem. 3712.
27. I. Nagazi and A. Haddad (2011). Mater. Res. Bull. 47, 356.
28. G. M. Sheldrick SHELXS-97, A Program for Crystal Structure Determination (University of
Göttingen, Gottingen, 1997).
29. G. M. Sheldrick SHELXL-97, A Program for the Refinement of Crystal Structures (University of
Göttingen, Gottingen, 1997).
30. L. J. Farrugia (1999). J. Appl. Crystallogr. 32, 837.
31. K. Brandenburg (1997). DIAMOND Visual Crystal Structure Information System.
32. T. Hori, S. Himeno, and O. J. Tamada (1992). Chem. Soc. Dalton Trans. 275.
33. E. Burholder, V. Golub, C. J. O’Connor, and J. Zubieta (2003). Chem. Commun. 17, 2128.
34. R. B. Fu, X. T. Wu, S. M. Hu, J. J. Zhang, Z. Y. Fu, W. X. Du, and S. Q. Xia (2003). Eur. J. Inorg.
Chem. 9, 1798.
35. E. Burholder, V. Golub, C. J. O’Connor, and J. Zubieta (2003). Inorg. Chem. 42, 6729.
36. V. Shinvaiah, T. Arumuganathan, and S. K. Das (2004). Inorg. Chem. Commun. 7, 365.
37. R. Strandberg (1973). Acta Chem. Scand. 27, 1004.
38. J. Fischer, L. Ricard, and P. Toledano (1974). J. Chem. Soc. Dalton Trans. 3, 941.
39. B. Hedman (1973). Acta Chem. Scand. 27, 3335.
40. W. H. Baur (1974). Acta Crystallogr. B 30, 1195.
41. Softbv web page by Pr. Stefan Adams. http://kristall.uni.mki.gwdg.de/softbv.
42. J. Ling and T. E. Albrecht-Schmitt (2007). J. Solid State Chem. 180, 1601.
43. Y. Porter and P. Shiv Halasyamani (2003). J. Solid State Chem. 174, 441.
44. T. Yamase (1998). Chem. Rev. 98, 307.
45. X. M. Zhang, B. Z. Shen, X. Z. You, and H. K. Fun (1997). Polyhedron 16, 95.

123
View publication stats

You might also like