11.03.2014 Views

School of Engineering and Science - Jacobs University

School of Engineering and Science - Jacobs University

School of Engineering and Science - Jacobs University

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Hybrid Organic-Inorganic Polyoxometalates<br />

Functionalized by Diorganotin Groups<br />

by<br />

Firasat Hussain<br />

A thesis submitted in partial fulfillment<br />

<strong>of</strong> the requirements for the degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy<br />

Approved, Thesis Committee:<br />

Pr<strong>of</strong>. Ulrich Kortz (Mentor), IUB<br />

Pr<strong>of</strong>. Ryan M Richards, IUB<br />

Dr. Michael H Dickman, IUB<br />

Pr<strong>of</strong>. Michael T Pope<br />

Georgetown <strong>University</strong>, U.S.A.<br />

Pr<strong>of</strong>. Emmanuel Cadot<br />

Université de Versailles, France<br />

Date <strong>of</strong> defense: 19 May 2006<br />

<strong>School</strong> <strong>of</strong> <strong>Engineering</strong> <strong>and</strong> <strong>Science</strong>


To my beloved parents


Abstract<br />

Polyoxometalates (POMs) are a well-known class <strong>of</strong> inorganic metal-oxygen clusters<br />

with an unmatched structural variety combined with a multitude <strong>of</strong> properties. The search<br />

for novel POMs is predominantly driven by exciting catalytic, medicinal, material science<br />

<strong>and</strong> bioscience applications. However, the mechanism <strong>of</strong> action <strong>of</strong> most polyoxoanions is<br />

not selective towards a specific target. In order to improve selectivity it appears highly<br />

desirable to attach organic functionalities covalently to the surface <strong>of</strong> polyoxoanions.<br />

The hydrolytic stability <strong>of</strong> the Sn-C bond enables the synthesis <strong>of</strong> a novel class <strong>of</strong><br />

polyoxoanions via attachment <strong>of</strong> organometallic functionalities based on Sn(IV) to the<br />

surface <strong>of</strong> lacunary polyoxoanion precursors.<br />

By reacting (CH 3 ) 2 SnCl 2 with Na 9 (α-XW 9 O 33 ) (X = As III , Sb III ) in aqueous acidic<br />

medium leads to the formation <strong>of</strong> 2-D solid-state structures with inorganic <strong>and</strong> organic<br />

surface, which are rare examples <strong>of</strong> discrete polyoxoanions. (CsNa 4 {(Sn(CH 3 ) 2 ) 3 O(H 2 O) 4<br />

(β-AsW 9 O 33 )}·5H 2 O) ∞ (CsNa-1) <strong>and</strong> the isostructural (CsNa 4 [(Sn(CH 3 ) 2 ) 3 O(H 2 O) 4 (<br />

β-SbW 9 O 33 )]·5H 2 O) ∞ (CsNa-2)It has been synthesized <strong>and</strong> characterized by multinuclear<br />

NMR spectroscopy, FTIR spectroscopy <strong>and</strong> elemental analysis. They crystallizes<br />

in the orthorhombic system, space group P na2 1 , with identical unit cell parameters a =<br />

26.118(2) Å, b = 16.064(1) Å, c = 13.776(1) Å, <strong>and</strong> Z = 1. Multinuclear NMR ( 183 W,<br />

119 Sn, 13 C, 1 H) showed that CsNa-1 <strong>and</strong> CsNa-2 decomposes in solution leading to the<br />

monomeric species [{Sn(CH 3 ) 2 (H 2 O) 2 } 3 (β-XW 9 O 33 )] 3− (X = As III (1), Sb III (2). Polyanions<br />

1 <strong>and</strong> 2 consist <strong>of</strong> a (β-XW 9 O 33 ) fragment which is stabilized by three dimethyltin<br />

fragments. The three dimethyltin groups <strong>of</strong> 1 <strong>and</strong> 2 are grafted onto the polyanion via<br />

two Sn-O(W) bonds involving the terminal O atoms on the side <strong>of</strong> the hetero atom lone<br />

pair. This polyanion has a nominal C s symmetry.<br />

i


Reacting (CH 3 ) 2 SnCl 2 with the superlacunary polyanion [H 2 P 4 W 24 O 94 ] 22− resulted in<br />

a dimeric hybrid organic-inorganic polyanion, [{Sn(CH 3 ) 2 } 4 (H 2 P 4 W 24 O 94 ) 2 ] 36− (3).It has<br />

been characterized by multinuclear NMR spectroscopy, FTIR spectroscopy <strong>and</strong> elemental<br />

analysis. Single-crystal X-ray analysis <strong>of</strong> K 17 Li 19 [{Sn(CH 3 ) 2 } 4 (H 2 P 4 W 24 O 94 ) 2 ]·51H 2 O<br />

showed that it crystallizes in the tetragonal system, space group P 4 2 /nmc, a = b =<br />

21.5112(17) Å, c = 27.171(3) Å, Z = 2. Polyanion 3 is composed <strong>of</strong> two (H 2 P 4 W 24 O 94 )<br />

fragments that are linked by four equivalent diorganotin groups. The unprecedented assembly<br />

3 has D 2d symmetry <strong>and</strong> contains a hydrophobic pocket at its center.<br />

The trimeric, hybrid organic-inorganic tungstophophate(V)<br />

[{Sn(CH 3 ) 2 (OH)} 3 (Sn(CH 3 ) 2 ) 3 {A-PW 9 O 34 } 3 ] 18− (4) has been synthesized in acidic medium.<br />

It has been characterized by FTIR spectroscopy, elemental analysis <strong>and</strong> cyclic voltametry.<br />

Single-crystal X-ray analysis <strong>of</strong> the compound showed that it crystallizes in rhombohedral<br />

system, space group R 3m, a = b = 29.7445(7) Å, c = 15.5915(7) Å, Z = 3. The polyanion<br />

is composed <strong>of</strong> three (A-α-PW 9 O 34 ) fragments that are linked by six dimethyltin groups,<br />

where the outer three ones are connected by µ 2 − OH bridges leading to a cyclic core<br />

which act as a cap leading to a bowl type structure. This arrangement results in a cyclic,<br />

trimeric, unprecedented polyanion assembly with C 3v symmetry.<br />

The tetrameric, hybrid organic-inorganic tungstoarsenate(III) [{Sn(CH 3 ) 2 (H 2 O)} 2<br />

{Sn(CH 3 ) 2 }As 3 (α-AsW 9 O 33 ) 4 ] 12− (5) has been characterized by multinuclear NMR spectroscopy,<br />

FTIR spectroscopy <strong>and</strong> elemental analysis. Single-crystal X-ray analysis <strong>of</strong> the<br />

compound showed that it crystallizes in the monoclinic system, space group P 21/c, with<br />

a = 22.612(2) Å, b = 19.954(2) Å, c = 41.099(4) Å, Z = 4. Polyanion 5 is composed <strong>of</strong> four<br />

(B-α-AsW 9 O 33 ) fragments that are linked by three dimethyltin groups <strong>and</strong> three As(III)<br />

atoms resulting in an unprecedented, chiral polyoxoanion assembly with C 1 symmetry.<br />

Interaction <strong>of</strong> (CH 3 ) 2 SnCl 2 with Na 9 [A-XW 9 O 34 ] (X = P V , As V ) in aqueous acidic<br />

medium resulted in the dodecameric, ball-shaped anions [{Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12<br />

(A-XW 9 O 34 ) 12 ] 36− (X = P V , As V ) (6-7). They has been characterized by multinuclear<br />

NMR spectroscopy, FTIR spectroscopy, scanning tunneling microscopy <strong>and</strong> elemental<br />

analysis. Both compounds crystallize as mixed cesium-sodium salts <strong>and</strong> are isomorphous.<br />

Single crystal Single-crystal X-ray analysis <strong>of</strong> the compound showed that it crystallizes in<br />

ii


the cubic system (space group I m¯3 with a = b = c = 32.7441(4) Å, Z = 2 . The spherical<br />

structure <strong>of</strong> the polyanion is spectacular in terms <strong>of</strong> geometry <strong>and</strong> size (diameter <strong>of</strong> 30<br />

Å) <strong>and</strong> is unprecedented in polyoxotungstate chemistry. This supermolecular assembly<br />

is composed <strong>of</strong> 12 trilacunary [A-XW 9 O 34 ] 9− (X= As V , P V ) Keggin fragments which are<br />

linked by 36 dimethyltin groups [12 inner (CH 3 ) 2 Sn 2+ <strong>and</strong> 24 outer (CH 3 ) 2 (H 2 O)Sn 2+<br />

groups] resulting in a polyanion with T h symmetry. The polyanion contains 1000 atoms<br />

with a molar mass <strong>of</strong> around 33000 gmol −1 .<br />

The bis-phenyltin substituted lone pair containing tungstoarsenate [(C 6 H 5 Sn) 2 As 2 W 19 O 67<br />

(H 2 O)] 8− (8) has been synthesized <strong>and</strong> characterized by multinuclear NMR, FTIR spectroscopy<br />

<strong>and</strong> elemental analysis. Single-crystal X-ray analysis <strong>of</strong><br />

(NH 4 ) 7 Na[(C 6 H 5 Sn) 2 As 2 W 19 O 67 (H 2 O)]·17.5H 2 O showed that it crystallizes in the monoclinic<br />

system, space group P 2 1 /c, with, a = 18.3127(17) Å, b = 24.403(2) Å, c =<br />

22.965(2) Å, β= 106.223(2) ◦ , <strong>and</strong> Z = 4. Polyanion 8 consists <strong>of</strong> two B-α-(As III W 9 O 33 )<br />

Keggin moieties linked via WO(H 2 O) fragment <strong>and</strong> two (SnC 6 H 5 ) 3+ groups leading to a<br />

s<strong>and</strong>wich-type structure with nominal C 2v symmetry.<br />

The titanium(IV), disubstituted lone pair containing tungstoarsenate<br />

[(TiOH) 2 WO(H 2 O)As 2 W 19 O 67 ] 8− (9) has been synthesized <strong>and</strong> characterized by FTIR<br />

spectroscopy <strong>and</strong> elemental analysis. Single-crystal X-ray analysis was carried out on<br />

Cs 8 [(TiOH) 2 WO(H 2 O)As 2 W 19 O 67 ]·10.5H 2 O which crystallizes in the monoclinic system,<br />

space group P 2 1 /m , with a = 12.7764(19)Å, b = 19.425(3) Å, c = 18.149(3) Å,<br />

β=110.23(3) ◦ , <strong>and</strong> Z = 2. Polyanion (9) consists <strong>of</strong> two B-α-(As III W 9 O 33 ) Keggin moieties<br />

linked via one [WO(H 2 O)] 4+ fragment <strong>and</strong> two Ti 4+ ions leading to a s<strong>and</strong>wich-type<br />

structure with nominal C 2v symmetry.<br />

Interaction <strong>of</strong> solid TiO(SO 4 ) with K 14 [P 2 W 19 O 69 (OH 2 )] in an aqueous acidic medium<br />

resulted in a novel, trimeric polyoxometalate . The compound has been characterized<br />

by FTIR spectroscopy <strong>and</strong> elemental analysis. Single-crystal X-ray analysis <strong>of</strong> the compound<br />

showed that it crystallizes in rhombohedral system, space group R 3m, a = b =<br />

29.7444(7) Å, c = 13.6254(9) Å, Z = 3. The polyanion (10) is composed <strong>of</strong> three α-<br />

Ti 3 PW 9 O 34 Keggin fragments that are linked via Ti-O-Ti bridges leading to a trimeric<br />

assembly. Interestingly, one <strong>of</strong> the titanium from each fragment is connected to the phosiii


phate leading to a trimeric-capped type structure with nominal C 3v symmetry.<br />

Interaction <strong>of</strong> solid TiO(SO 4 ) with [γ-SiW 10 O 36 ] 8− in an aqueous acidic medium resulted<br />

in a novel, tetrameric polyoxometalate [β-Ti 2 SiW 10 O 39 ] 4 ] 24− (11).It has been characterized<br />

by multinuclear NMR spectroscopy, FTIR spectroscopy <strong>and</strong> elemental analysis.<br />

Single-crystal X-ray analysis <strong>of</strong> the compound showed that it crystallizes in monoclinic<br />

system, space group P 2 1 /n, a = 12.5188(13) Å, b = 18.864(2) Å, c = 41.075(4) Å, β<br />

= 97.450(2) ◦ Z = 2. The polyanion is composed <strong>of</strong> four β-Ti 2 SiW 10 O 39 Keggin fragments<br />

that are linked via Ti-O-Ti bridges leading to a cyclic assembly. The solid-state<br />

structure <strong>of</strong> K 24 [(β-Ti 2 SiW 10 O 39 ) 4 ]·50H 2 O shows face-by-face self assembled polyanions<br />

(K-11) along the crystallographic ‘a’ axis, which leads to a nanotube-like arrangement.<br />

The cadmium(II)-substituted tungstoarsenate [Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ] 12− (12) has<br />

been synthesized <strong>and</strong> characterized in the solid state by XRD, FTIR spectroscopy, elemental<br />

analysis <strong>and</strong> 183 W <strong>and</strong> 111 Cd NMR spectroscopy. Single crystal X-ray analysis<br />

<strong>of</strong> Cs 4 K 3 Na 5 [Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ]·20H 2 O (CsKNa-12), shows that it crystallizes in<br />

the monoclinic system, space group P 2 1 /n, with a = 13.1402(12) Å, b = 19.0642(17) Å,<br />

c = 17.5666(15) Å, β = 90.274(2) ◦ <strong>and</strong> Z = 2. Polyanion (9) consists <strong>of</strong> two lacunary<br />

[B-α-AsW 9 O 34 ] 10− Keggin moieties linked via a rhomb like (Cd 4 O 14 Cl 2 ) cluster leading to<br />

a s<strong>and</strong>wich-type structure. Interestingly, the two external Cd atoms each have a terminal<br />

Cl lig<strong>and</strong>, but the derivative have terminal water lig<strong>and</strong>s, [Cd 4 (H 2 O) 2 (B-α-AsW 9 O 34 ) 2 ] 10−<br />

(13), which has been proved by 183 W <strong>and</strong> 111 Cd NMR spectroscopy.<br />

The indium(III)-substituted polyanions [In 3 Cl 2 (B-α-PW 9 O 34 ) 2 ] 11− (14), [In 3 Cl 2 (P 2 W 15<br />

O 56 ) 2 ] 17− (15), [In 4 (H 2 O) 10 ( β-AsW 9 O 32 OH) 2 ] 4− (16) <strong>and</strong> [In 4 (H 2 O) 10 (β-SbW 9 O 33 ) 2 ] 6−<br />

(17) have been synthesized <strong>and</strong> characterized in the solid state by FTIR spectroscopy<br />

<strong>and</strong> elemental analysis <strong>and</strong> 183 W-NMR. Single-crystal X-ray analysis showed that D,L-<br />

(NH 4 ) 11 [In 3 Cl 2 (B-α-PW 9 O 34 ) 2 ]·16H 2 O (NH4-14) showed that it crystallizes in the monoclinic<br />

system, space group P 2 1 /n, with a = 17.5090(10) Å, b = 12.7361(7) Å, c =<br />

18.7955(11) Å, β = 107.3030(10) ◦ , <strong>and</strong> Z = 2;<br />

D,L- (NH 4 ) 9 Na 8 [In 3 Cl 2 (P 2 W 15 O 56 ) 2 ]·39H 2 O (NH4Na-15) crystallizes in the triclinic system,<br />

space group P ¯1, with a = 13.0643(5) Å, b = 14.8901(6) Å, c = 19.8603(8) Å, α<br />

= 92.2920(10) ◦ , β = 90.8680(10) ◦ , γ = 100.5630(10) ◦ , <strong>and</strong> Z = 1; RbNa 3 [In 4 (H 2 O) 10 (βiv


AsW 9 O 32 OH) 2 ]·36H 2 O (RbNa-16) crystallizes in the triclinic system, space group P ¯1,<br />

with a = 12.8142(5) Å, b = 12.8672(5) Å, c = 16.1794(7) Å, α = 91.1370(10) ◦ , β =<br />

105.9450(10) ◦ , γ = 104.0980(10) ◦ <strong>and</strong> Z = 1; K 4 Na 2 [In 4 (H 2 O) 10 (β-SbW 9 O 33 ) 2 ]·30H 2 O<br />

(KNa-17) crystallizes also in the triclinic system, space group P ¯1, with a = 12.2306(9)<br />

Å, b = 12.7622(10) Å, c = 16.1639(12) Å, α = 73.9890(10) ◦ , β = 76.5550(10) ◦ , γ =<br />

86.2130(10) ◦ <strong>and</strong> Z = 1. Synthesis <strong>of</strong> polyanions 14-17 have been accomplished by reaction<br />

<strong>of</strong> In 3+ ions with the lacunary precursors [B-α-PW 9 O 34 ] 9− , [P 2 W 15 O 56 ] 12− <strong>and</strong><br />

[α-XW 9 O 33 ] 9− (X = As III , Sb III ), respectively, in aqueous, acidic medium. The chiral<br />

polyanions 14 <strong>and</strong> 15 are composed <strong>of</strong> three In(III) ions connected to two (B-α-PW 9 O 34 )<br />

<strong>and</strong> (P 2 W 15 O 56 ) fragments, respectively. Only one <strong>of</strong> the inner sites <strong>of</strong> the central rhombus<br />

is occupied by an indium atom <strong>and</strong> the two outer indium atoms contain a terminal<br />

Cl lig<strong>and</strong>. The structural stability <strong>of</strong> the polyanions 14-17 in solution were studied by<br />

31 P <strong>and</strong> 183 W NMR.<br />

v


Acknowledgements<br />

I extend my sincere gratitude <strong>and</strong> appreciation to many people who made this doctoral<br />

thesis possible. Special thanks are due to my mentor Pr<strong>of</strong>. Ulrich Kortz for his consistent<br />

guidance, advice <strong>and</strong> cooperation. Thanks are also due to my lab mates <strong>and</strong> heartfelt<br />

thanks to Mr. Markus Reicke for cooperation in lab work also thanks to Dr. Li-Hua Bi<br />

for her fruitful suggestions.<br />

I would like to thank Dr. M. H. Dickman,(IUB) who instructed me on X-ray crystallography,<br />

starting from all the basics; how to mount a crystal to solving the crystal structure.<br />

Several compounds were re-collected during the instruction <strong>and</strong> also compound 10 was<br />

collected here at IUB in Pr<strong>of</strong>.Ulrich Kortz’s lab.<br />

I thank my mentor Pr<strong>of</strong>.Ulrich Kortz again who brought valuable gifts (crystal structures)<br />

from Florida State <strong>and</strong> Georgetown <strong>University</strong>, U.S.A. during his visits. X-ray data for<br />

compounds 1, 3-8,11-17 were collected in the Chemistry department <strong>of</strong> Florida State<br />

<strong>and</strong> Georgetown <strong>University</strong>, U.S.A. X-ray data for compound 2 was collected during his<br />

visit to Hamburg <strong>University</strong>.<br />

X-ray data for compounds 10 <strong>and</strong> 20 were collected in Kortz’s lab, IUB.<br />

Thanks are due to Pr<strong>of</strong>. R. M. Richards,(IUB) for being one <strong>of</strong> the examiners <strong>of</strong> the<br />

thesis.<br />

I would like to thank Pr<strong>of</strong>. M. T. Pope, (Georgetown <strong>University</strong>, U.S.A.) for his time to<br />

time suggestions <strong>and</strong> also for being an external examiner <strong>of</strong> the thesis.<br />

I would like to thank Pr<strong>of</strong>. E. Cadot, (Université de Versailles, France.) for being an<br />

external examiner <strong>of</strong> the thesis.<br />

I would like to thank Pr<strong>of</strong>. M. Winterhalter <strong>and</strong> his group for HPPS measurements.<br />

I would like to thank our collaborators Dr. B. Keita, <strong>and</strong> Pr<strong>of</strong>. L. Nadjo, (Université<br />

vi


Paris-Sud, Orsay Cedex, France) for electrochemistry studies.<br />

I would also like to thank our collaborators Pr<strong>of</strong>. P. Müller <strong>and</strong> his coworkers Mr. M. S.<br />

Alam, V. Dremov, Physikalisches Institut III, (Universität Erlangen-Nürnberg, Germany)<br />

for STM studies.<br />

I am highly indebted to Pr<strong>of</strong>. G. Haerendel for giving me an opportunity to carry out my<br />

research career in IUB.<br />

I would also like to acknowledge IUB for the fellowship.<br />

My special thanks goes to my brother, parents, friends <strong>and</strong> Katja Knoop (Office secretary)<br />

for their cooperation, support <strong>and</strong> encouragement during the time <strong>of</strong> my studies.<br />

vii


Table <strong>of</strong> Contents<br />

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

List <strong>of</strong> Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

List <strong>of</strong> Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

i<br />

xi<br />

xvi<br />

1 Introduction 1<br />

1.1 Historical perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.1.1 The clathrate-like structure . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.1.2 Chemical elements taking part in POMs . . . . . . . . . . . . . . . 4<br />

1.1.3 Features <strong>of</strong> POMs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

1.2 Geometric Structures <strong>of</strong> representative types <strong>of</strong> Polyanions: Keggin, Lindqvist,<br />

Anderson-Evans <strong>and</strong> Wells-Dawson . . . . . . . . . . . . . . . . . . . . . . 6<br />

1.3 Lacunary Species . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

2 Experimental 18<br />

2.0.1 Reagents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.1 Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.1.1 Infrared spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.1.2 Single crystal X-ray diffraction . . . . . . . . . . . . . . . . . . . . . 18<br />

2.1.3 Multinuclear magnetic resonance spectroscopy . . . . . . . . . . . . 19<br />

2.1.4 Cyclic voltametry . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2.1.5 Elemental analyses <strong>and</strong> thermogravimetric analysis . . . . . . . . . 19<br />

2.2 Preparation <strong>of</strong> starting materials . . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.2.1 Synthesis <strong>of</strong> Na 9 [AsW 9 O 33 ]·27H 2 O . . . . . . . . . . . . . . . . . . 20<br />

2.2.2 Synthesis <strong>of</strong> Na 9 [SbW 9 O 33 ]·27H 2 O . . . . . . . . . . . . . . . . . . . 20<br />

2.2.3 Synthesis <strong>of</strong> K 14 [As 2 W 19 O 67·(H 2 O)] . . . . . . . . . . . . . . . . . . 20<br />

2.2.4 Synthesis <strong>of</strong> A & B Na 8 [HAsW 9 O 34 ]·11H 2 O (A & B-Type AsW 9 O 34 ) 21<br />

2.2.5 Synthesis <strong>of</strong> A-Na 9 [PW 9 O 34 ]·7H 2 O . . . . . . . . . . . . . . . . . . 21<br />

2.2.6 Synthesis <strong>of</strong> Cs 6 [P 2 W 5 O 23 ]·H 2 O . . . . . . . . . . . . . . . . . . . . 21<br />

2.2.7 Synthesis <strong>of</strong> Cs 7 [PW 10 O 36 ]·H 2 O . . . . . . . . . . . . . . . . . . . . 22<br />

2.2.8 Synthesis <strong>of</strong> Na 20 [P 6 W 18 O 79 ]·37.5H 2 O . . . . . . . . . . . . . . . . . 22<br />

viii


2.2.9 Synthesis <strong>of</strong> K 12 [H 2 P 2 W 12 O 48 ]·24H 2 O . . . . . . . . . . . . . . . . . 22<br />

2.2.10 Synthesis <strong>of</strong> K 16 Li 2 [H 6 P 4 W 24 O 94 ]·33H 2 O . . . . . . . . . . . . . . . 23<br />

2.2.11 Synthesis <strong>of</strong> K 28 Li 5 [H 7 P 8 W 48 O 184 ]·92H 2 O . . . . . . . . . . . . . . . 23<br />

2.2.12 Synthesis <strong>of</strong> Na 27 [NaAs 4 W 40 O 140 ]·60H 2 O . . . . . . . . . . . . . . . 23<br />

2.2.13 Synthesis <strong>of</strong> K 8 [γ-SiW 10 O 36 ]·20H 2 O . . . . . . . . . . . . . . . . . . 23<br />

2.2.14 Synthesis <strong>of</strong> K 7 [PW 11 O 39 ]·14H 2 O . . . . . . . . . . . . . . . . . . . 24<br />

2.2.15 Synthesis <strong>of</strong> K 14 [P 2 W 19 O 69 (H 2 O)]·24H 2 O . . . . . . . . . . . . . . . 24<br />

3 Results 26<br />

3.1 The hybrid organic-inorganic 2-D material<br />

(CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-XW 9 O 33 )]·5H 2 O) ∞<br />

(X = As III , Sb III ) <strong>and</strong> its solution properties . . . . . . . . . . . . . . . . 26<br />

3.1.1 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

3.1.2 Results <strong>and</strong> discussion . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3.1.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

3.2 The tetrakis-dimethyltin containing<br />

tungstophosphate ({Sn(CH 3 ) 2 } 4 {H 2 P 4 W 24 O 92 } 2 ) 28−<br />

evidence for lacunary Preyssler ion . . . . . . . . . . . . . . . . . . . . . . 35<br />

3.2.1 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

3.2.2 Results <strong>and</strong> discussion . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

3.2.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

3.3 The trimeric-dimethyltin containing bowl shaped tungstophosphate(V):<br />

Cs 12 Na 6 {Sn(CH 3 ) 2 (OH)} 3<br />

(Sn(CH 3 ) 2 ) 3 {A-PW 9 O 34 } 3·14H 2 O . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.3.1 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.3.2 Results <strong>and</strong> discussion . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

3.3.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

3.4 The tetrameric, chiral tungstoarsenate(III),<br />

({Sn(CH 3 ) 2 (H 2 O)} 2 {Sn(CH 3 ) 2 }As 3 {α-AsW 9 O 33 } 4 ) 21− . . . . . . . . . . . 51<br />

3.4.1 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

3.4.2 Results <strong>and</strong> discussion . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

3.4.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

3.5 The gigantic, ball-shaped heteropolytungstates<br />

({Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-XW 9 O 34 ) 12 ) 36− (X= P V , As V ) . . . . 56<br />

3.5.1 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56<br />

3.5.2 Results <strong>and</strong> discussions . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

3.5.3 STM studies <strong>of</strong> Cs-6 . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

3.5.4 HPPS measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

ix


3.5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

3.6 The bis-phenyltin substituted, lone pair containing tungstoarsenate:<br />

({C 6 H 5 Sn} 2 As 2 W 19 O 67 (H 2 O)) 8− . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

3.6.1 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

3.6.2 Results <strong>and</strong> discussion . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

3.6.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

3.7 Conclusions <strong>and</strong> Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

3.7.1 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

3.7.2 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

3.8 The di-titanium substituted, lone pair containing tungstoarsenate:<br />

{(TiOH) 2 WO(H 2 O)(B-α-AsW 9 O 33 ) 2 } 8− . . . . . . . . . . . . . . . . . . . 82<br />

3.8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

3.8.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

3.8.3 Results <strong>and</strong> discussion . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

3.8.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

3.9 The cyclic trimeric-titanium substituted,<br />

tungstophosphate: [{Ti 3 O 4 (A-α-PW 9 O 34 )} 3 (PO 4 )] 13− . . . . . . . . . . . . 88<br />

3.9.1 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

3.10 Structural control on the nanomolecular scale: Self- assembly <strong>of</strong> the polyoxotungstate<br />

wheel<br />

({β-Ti 2 SiW 10 O 39 } 4 ) 24− . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

3.10.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

3.10.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

3.10.3 Results <strong>and</strong> discussion . . . . . . . . . . . . . . . . . . . . . . . . . 93<br />

3.10.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

3.11 Structure <strong>and</strong> solution properties <strong>of</strong> the<br />

cadmium(II)-substituted tungstoarsenate:<br />

(Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ) 12− . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

3.11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

3.11.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

3.11.3 Results <strong>and</strong> discussion . . . . . . . . . . . . . . . . . . . . . . . . . 104<br />

3.11.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107<br />

3.12 Some indium(III)-substituted polyoxotungstates <strong>of</strong> the Keggin <strong>and</strong> Dawson<br />

type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108<br />

3.12.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108<br />

3.12.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110<br />

3.12.3 Results <strong>and</strong> discussion . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

3.12.4 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122<br />

x


Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129<br />

NMR Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130<br />

Incomplete Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143<br />

Curriculum Vitae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152<br />

xi


List <strong>of</strong> Figures<br />

1.1 Polyhedral models <strong>of</strong> the three possible unions between two MO 6 octahedral<br />

units. A) corner-sharing,B) edge-sharing <strong>and</strong> C) face-sharing . . . . . . . . 3<br />

1.2 Ball <strong>and</strong> stick represention <strong>of</strong> the α Keggin type cluster; the black balls<br />

represent tungsten <strong>and</strong> the red balls represent oxygen . . . . . . . . . . . . 5<br />

1.3 Ball/stick <strong>and</strong> polyhedral representation <strong>of</strong> the alpha isomer <strong>of</strong> the Keggin<br />

structure with different types <strong>of</strong> oxygen. Color codes: The blue polyhedra<br />

represent the tungsten <strong>and</strong> the red balls represent oxygen . . . . . . . . . . 7<br />

1.4 Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong> [Cu 3 Na 3 (H 2 O) 9 (α-<br />

As III W 9 O 33 ) 2 ] 9− . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

1.5 Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong> {(O 3 POPO 3 )Mo 6 O 18 } 4− 8<br />

1.6 Ball <strong>and</strong> stick representation <strong>of</strong> the Lindqvist structure . . . . . . . . . . . 9<br />

1.7 Polyhedron representation <strong>of</strong> the Anderson structure . . . . . . . . . . . . 9<br />

1.8 Ball <strong>and</strong> stick representation <strong>of</strong> the Wells-Dawson structure . . . . . . . . . 10<br />

1.9 Schematic representation <strong>of</strong> the formation <strong>of</strong> Keggin-type lacunary. Courtesy:<br />

Dr. Santiago C. Reinoso . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

1.10 Schematic representation <strong>of</strong> the formation <strong>of</strong> Keggin- <strong>and</strong> Wells-Dawson<br />

type lacunary species in tungsto-phosphate system with respect to pH.Courtesy:<br />

Dr. Santiago C. Reinoso . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

1.11 Monomeric (left) <strong>and</strong> dimeric species (right) <strong>of</strong> monoorganotin containing<br />

polyoxotungstates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

3.1 FTIR spectra <strong>of</strong> compound CsNa-1(red) <strong>and</strong> Na 9 [α-AsW 9 O 33 ] (blue) . . 27<br />

3.2 FTIR spectra <strong>of</strong> compound CsNa-2(red) <strong>and</strong> Na 9 [α-SbW 9 O 33 ](blue) . . . 28<br />

3.3 W 183 NMR spectra <strong>of</strong> (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 ( β-XW 9 O 33 )]·5H 2 O)<br />

([X = As (top)CsNa-1, Sb (bottom)CsNa-2] . . . . . . . . . . . . . . . . 30<br />

3.4 Sn 119 NMR spectra <strong>of</strong> (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 ( β-XW 9 O 33 )]·5H 2 O)<br />

[X = As (top)CsNa-1, Sb(bottom)CsNa-2] . . . . . . . . . . . . . . . . . 31<br />

xii


3.5 Left: combined polyhedral <strong>and</strong> ball/stick representation <strong>of</strong> the 2-D solid<br />

state structure <strong>of</strong> (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-XW 9 O 33 )]·5H 2 O) ∞ (X<br />

= As, CsNa-1; Sb, CsNa-2). The WO 6 octahedra are purple <strong>and</strong> the<br />

balls represent tin (green), arsenic/antimony (blue), oxygen (red) <strong>and</strong> carbon<br />

(yellow). Hydrogen atoms are omitted for clarity (left),Right: Side<br />

view <strong>of</strong> the 2-D solid state structure <strong>of</strong> (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-<br />

XW 9 O 33 )]·5H 2 O) ∞ (X = As, CsNa-1; Sb, CsNa-2). . . . . . . . . . . . . 32<br />

3.6 Left:Ball <strong>and</strong> stick representation <strong>of</strong> [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-XW 9 O 33 )] 3−<br />

(X = As,1; Sb,2.(A) The balls represent tungsten (black), tin (green), arsenic/antimony<br />

(blue), oxygen (red), carbon (yellow) <strong>and</strong> hydrogen (small<br />

black); Right:Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong> the<br />

[{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-XW 9 O 33 )] 3− (X = As, 1; Sb, 2).(B) The WO 6<br />

octahedra are red <strong>and</strong> the color codes <strong>of</strong> balls are same as above . . . . . . 33<br />

3.7 FTIR spectra <strong>of</strong> compound KLi-3(red) <strong>and</strong> K 16 Li 2 [H 6 P 4 W 24 O 94 ] (blue) . . 35<br />

3.8 Ball <strong>and</strong> stick (left), Polyhedron (right) representation <strong>of</strong> polyanion 3 . . . 37<br />

3.9 Ball <strong>and</strong> stick representation <strong>of</strong> the asymmetric unit <strong>of</strong> polyanion 3 . . . . 38<br />

3.10 Cyclic voltammogram <strong>of</strong> 2 × 10 −4 M solution <strong>of</strong> 3 in a pH 4 medium (1<br />

M CH 3 COOLi + CH 3 COOH). The scan rate was 10 mV.s −1 , the working<br />

electrode was glassy carbon <strong>and</strong> the reference electrode was SCE. (A) The<br />

whole voltammetric pattern(left), (B) The voltammetric pattern restricted<br />

to the first redox processes (right) . . . . . . . . . . . . . . . . . . . . . . . 40<br />

3.11 Comparison <strong>of</strong> the cyclic voltammograms <strong>of</strong> 2 × 10 −4 M solution <strong>of</strong> 3 in<br />

a pH 4 medium (1 M CH 3 COOLi + CH 3 COOH). The scan rate was 10<br />

mV.s −1 , the working electrode was glassy carbon <strong>and</strong> the reference electrode<br />

was SCE. (A) Comparison <strong>of</strong> the voltammetric patterns <strong>of</strong> 3 <strong>and</strong><br />

[H 2 P 4 W 24 O 94 ] 22− (left)(B) Comparison <strong>of</strong> the voltammetric patterns <strong>of</strong> 3<br />

<strong>and</strong> [H 7 P 8 W 48 O 184 ] 33 (right) . . . . . . . . . . . . . . . . . . . . . . . . . . 41<br />

3.12 FTIR spectra <strong>of</strong> compound CsNa-4(red) <strong>and</strong> Na 9 [(A-α-PW 9 O 34 )](blue,) . 44<br />

3.13 Ball/stick (left) <strong>and</strong> Polyhedron (right) representation <strong>of</strong> polyanion 4 . . . 46<br />

3.14 The central core <strong>of</strong> Sn(CH 3 ) 2 2+ connected to each other by three (µ 2 -OH)<br />

bridges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

3.15 Cyclic voltammogram <strong>of</strong> 2 × 10 −4 M solution <strong>of</strong> 4 in a pH 4 medium (1<br />

M CH 3 COOLi + CH 3 COOH). The scan rate was 10 mV.s −1 , the working<br />

electrode was glassy carbon <strong>and</strong> the reference electrode was SCE. (A) Superposition<br />

<strong>of</strong> the CVs restricted to the first redox pattern for trimer <strong>and</strong><br />

A-α-PW 9 O 34 respectively.(left), (B) Complete CV for the trimer showing<br />

the presence <strong>of</strong> a second irreversible wave close to the electrolyte discharge.<br />

(right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49<br />

xiii


3.16 FTIR spectra <strong>of</strong> compound KNH4-5(red) <strong>and</strong> Na 9 [α-AsW 9 O 33 ](blue) . . 51<br />

3.17 Top view <strong>of</strong> [{Sn(CH 3 ) 2 (H 2 O)} 2 {Sn(CH 3 ) 2 }As 3 (α -AsW 9 O 33 ) 4 ] 21− (5).(left),the<br />

unique AsW 9 fragment <strong>and</strong> its three associated As linkers are not shown for<br />

clarity.The octahedra represent WO 6 <strong>and</strong> the balls are tin (green), arsenic<br />

(yellow), carbon (blue) <strong>and</strong> oxygen (red). Hydrogen atoms are omitted for<br />

clarity.(right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

3.18 Side view <strong>of</strong> [{Sn(CH 3 ) 2 (H 2 O)} 2 {Sn(CH 3 ) 2 }As 3 (α-AsW 9 O 33 ) 4 ] 21− The color<br />

code is the same as in above. . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

3.19 FTIR spectra <strong>of</strong> compound Cs-6(red) <strong>and</strong> Na 9 [A-PW 9 O 34 ](blue) . . . . . 57<br />

3.20 FTIR spectra <strong>of</strong> compound Cs-7(red)<strong>and</strong> Na 8 H[A-AsW 9 O 34 ](blue) . . . . 57<br />

3.21 Ball <strong>and</strong> stick representation <strong>of</strong> the asymmetric unit <strong>of</strong> Cs-6 <strong>and</strong> Cs-7<br />

showing the same labeling scheme as both compounds are isostructural . . 59<br />

3.22 Left:Ball <strong>and</strong> stick representation <strong>of</strong> Cs-6 <strong>and</strong> Cs-7 including the 14 cesium<br />

counter ions. The color code is as follows: tungsten (black), tin (blue),<br />

phosphorus/arsenic (yellow), oxygen (red), carbon (green) <strong>and</strong> cesium (purple).<br />

No hydrogens are shown for clarity, Right:Polyhedral representation<br />

<strong>of</strong> Cs-6 <strong>and</strong> Cs-7. The WO 6 octahedra are red <strong>and</strong> the XO 4 tetrahedra<br />

(X = P, As) are yellow. Otherwise, the labeling scheme is the same as in<br />

Figure 3.21A. No hydrogens <strong>and</strong> cesiums shown for clarity . . . . . . . . . 59<br />

3.23 Representation <strong>of</strong> the icosahedron spanned by the hetero atoms <strong>of</strong> Cs-6(P)<br />

<strong>and</strong> Cs-7(As) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

3.24 Left:Representation <strong>of</strong> the polyhedron spanned by the 12 inner tin atoms <strong>of</strong><br />

6 <strong>and</strong> 7.Right: Representation <strong>of</strong> the polyhedron spanned by the 24 outer<br />

tin atoms <strong>of</strong> Cs-6 <strong>and</strong> Cs-7 . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.25 Representation <strong>of</strong> the hexa-capped cube spanned by the 14 cesium ions <strong>of</strong><br />

Cs-6 <strong>and</strong> Cs-7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.26 Ball <strong>and</strong> stick representation <strong>of</strong> Cs-6 <strong>and</strong> Cs-7 highlighting the different<br />

polyhedral shells (12 inner Sn atoms: red, 12 hetero atoms: purple, 24<br />

outer Sn atoms: yellow)Courtesy: Dr. H. Bögge, <strong>University</strong> <strong>of</strong> Bielefeld,<br />

Germany. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

3.27 31 P(Left), <strong>and</strong> 183 W NMR spectra <strong>of</strong><br />

Cs 14 Na 22 {Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-PW 9 O 34 ) 12 ]·149H 2 O . . . . . 63<br />

3.28 STM pictures <strong>of</strong> {Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-PW 9 O 34 ) 12 ] 36− . . . . 66<br />

3.29 Size distribution by intensity <strong>and</strong> number <strong>of</strong> polyanion 6 . . . . . . . . . . 67<br />

3.30 Size distribution by intensity <strong>and</strong> number <strong>of</strong> polyanion 7 . . . . . . . . . . 67<br />

3.31 Size distribution by intensity <strong>and</strong> number <strong>of</strong> polyanion Cs-6 . . . . . . . . 68<br />

3.32 FTIR spectra <strong>of</strong> compound NH4-8(red) <strong>and</strong> K 14 [As 2 W 19 O 67 (H 2 O)](blue) . 71<br />

xiv


3.33 Combined polyhedron <strong>and</strong> ball/stick representations <strong>of</strong> [(C 6 H 5 Sn) 2 As 2 W 19 O 67<br />

(H 2 O)] 8− (left),Top view (right) . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

3.34 Projection <strong>of</strong> the crystal packing on the bc plane showing the 2-D arrangement<br />

<strong>of</strong> compound 8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

3.35 183 W NMR spectra <strong>of</strong> compound 8 . . . . . . . . . . . . . . . . . . . . . . 76<br />

3.36 119 Sn NMR spectra <strong>of</strong> compound 8 . . . . . . . . . . . . . . . . . . . . . . 76<br />

3.37 Ball/stick (left), polyhedral (right) representations <strong>of</strong> 9 . . . . . . . . . . . 84<br />

3.38 FTIR spectra <strong>of</strong> compound RbK-10(red) <strong>and</strong> K 14 [P 2 W 19 O 69 (H 2 O)](blue) 88<br />

3.39 Ball/stick <strong>and</strong> polyhedral representation <strong>of</strong> polyanion 10, color codes: blue<br />

polyhedra (W), yellow balls (Ti), red balls (O) <strong>and</strong> pink polyhedra (PO4) 89<br />

3.40 FTIR spectra <strong>of</strong> compound K-11(red) <strong>and</strong> K 8 [γ-SiW 10 O 36 ](blue) . . . . . 92<br />

3.41 Polyhedral (left) <strong>and</strong> ball <strong>and</strong> stick (right) representations <strong>of</strong><br />

({β-Ti 2 SiW 10 O 39 } 4 ) 24− . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92<br />

3.42 Side-view <strong>of</strong> compound 11 including the potassium ions (purple) inside the<br />

central cavity (K8) <strong>and</strong> above <strong>and</strong> below the cavity (K9, <strong>and</strong> K9’). . . . . . 94<br />

3.43 Dimer <strong>of</strong> Dimer <strong>of</strong> β(1,10)-Ti 2 (OH) 2 SiW 10 O 38 ] 6− . . . . . . . . . . . . . . . 95<br />

3.44 FTIR spectra <strong>of</strong> compound (CsKNa-12)(red) <strong>and</strong> Na 8 [A-HAsW 9 O 34 ]·11H 2 O<br />

(blue) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100<br />

3.45 FTIR spectra <strong>of</strong> compound (CsNa-13)(red) <strong>and</strong> Na 8 [A-HAsW 9 O 34 ]·11H 2 O<br />

(blue) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

3.46 183 W NMR spectra <strong>of</strong> [Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ] 12− (12), top, <strong>and</strong> [Cd 4 (H 2 O) 2 (Bα-AsW<br />

9 O 34 ) 2 ] 10− (13) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

3.47 111 Cd NMR spectra <strong>of</strong> [Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ] 12− 12, top, <strong>and</strong> [Cd 4 (H 2 O) 2 (Bα-AsW<br />

9 O 34 ) 2 ] 10− 13, bottom. . . . . . . . . . . . . . . . . . . . . . . . . . 104<br />

3.48 Combined polyhedral <strong>and</strong> ball <strong>and</strong> stick representation <strong>of</strong> [Cd 4 Cl 2 (B-α-<br />

AsW 9 O 34 ) 2 ] 12− . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

3.49 Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong> [In 3 Cl 2 (PW 9 O 34 ) 2 ] 11−<br />

14The red octahedra represent WO 6 , the blue tetrahedra represent PO 4<br />

<strong>and</strong> the balls represent indium (green) <strong>and</strong> chlorine (yellow). . . . . . . . 114<br />

3.50 183 W NMR spectrum <strong>of</strong> [In 3 Cl 2 (B-α-PW 9 O 34 ) 2 ] 11− . . . . . . . . . . . . . 116<br />

3.51 Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong> [In 3 Cl 2 (P 2 W 15 O 56 ) 2 ] 17−<br />

15. The color code is the same as in Figure 3.40 . . . . . . . . . . . . . . 116<br />

3.52 Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong><br />

[In 4 (H 2 O) 10 (β-AsW 9 O 32 OH) 2 ] 4− (16) <strong>and</strong> [In 4 (H 2 O) 10 (β-SbW 9 O 33 ) 2 ] 6− (17).<br />

The WO 6 octahedra are shown in red <strong>and</strong> the balls represent indium<br />

(green), arsenic/antimony (blue) <strong>and</strong> water molecules (red). . . . . . . . . 118<br />

3.53 183 W NMR spectrum <strong>of</strong> [In 4 (H 2 O) 10 (β-AsW 9 O 32 OH) 2 ] 4− at 293 K . . . . 120<br />

3.54 13 C NMR spectrum <strong>of</strong> polyanion 1 . . . . . . . . . . . . . . . . . . . . . . 130<br />

xv


3.55 13 C NMR spectrum <strong>of</strong> polyanion 2 . . . . . . . . . . . . . . . . . . . . . . 131<br />

3.56 31 P NMR spectrum <strong>of</strong> polyanion 3 . . . . . . . . . . . . . . . . . . . . . . 132<br />

3.57 119 Sn NMR spectrum <strong>of</strong> polyanion 3 . . . . . . . . . . . . . . . . . . . . . 133<br />

3.58 13 C NMR spectrum <strong>of</strong> polyanion 3 . . . . . . . . . . . . . . . . . . . . . . 134<br />

3.59 1 H NMR spectrum <strong>of</strong> polyanion 3 . . . . . . . . . . . . . . . . . . . . . . 135<br />

3.60 1 H NMR spectrum <strong>of</strong> polyanion 5 . . . . . . . . . . . . . . . . . . . . . . 136<br />

3.61 13 C NMR spectrum <strong>of</strong> polyanion 5 . . . . . . . . . . . . . . . . . . . . . . 137<br />

3.62 119 Sn NMR spectrum <strong>of</strong> polyanion 5 . . . . . . . . . . . . . . . . . . . . . 138<br />

3.63 1 H NMR spectrum <strong>of</strong> polyanion 6 . . . . . . . . . . . . . . . . . . . . . . 138<br />

3.64 13 C NMR spectrum <strong>of</strong> polyanion 6 . . . . . . . . . . . . . . . . . . . . . . 139<br />

3.65 119 Sn NMR spectrum <strong>of</strong> polyanion 6 . . . . . . . . . . . . . . . . . . . . . 139<br />

3.66 1 H NMR spectrum <strong>of</strong> polyanion 7 . . . . . . . . . . . . . . . . . . . . . . 140<br />

3.67 13 C NMR spectrum <strong>of</strong> polyanion 7 . . . . . . . . . . . . . . . . . . . . . . 140<br />

3.68 183 W NMR spectrum <strong>of</strong> polyanion 7 . . . . . . . . . . . . . . . . . . . . . 141<br />

3.69 1 H NMR spectrum <strong>of</strong> polyanion 8 . . . . . . . . . . . . . . . . . . . . . . 141<br />

3.70 13 C NMR spectrum <strong>of</strong> polyanion 8 . . . . . . . . . . . . . . . . . . . . . . 142<br />

3.71 119 Sn NMR spectrum <strong>of</strong> polyanion 8 in presence <strong>of</strong> sodium chloride . . . . 142<br />

3.72 FTIR spectra <strong>of</strong> compound K-18(red) <strong>and</strong> Na 9 [A-PW 9 O 34 ](blue) . . . . . 143<br />

3.73 31 P NMR spectrum <strong>of</strong> polyanion 18 . . . . . . . . . . . . . . . . . . . . . 145<br />

3.74 Size distribution by intensity <strong>of</strong> polyanion 18 . . . . . . . . . . . . . . . . 145<br />

3.75 Size distribution by number <strong>of</strong> polyanion 18 . . . . . . . . . . . . . . . . . 146<br />

3.76 Size distribution by volume <strong>of</strong> polyanion 18 . . . . . . . . . . . . . . . . . 146<br />

3.77 Ball/stick <strong>and</strong> polyhedron representation <strong>of</strong> solid state <strong>of</strong> 18 showing 1-D<br />

chain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146<br />

3.78 FTIR spectra <strong>of</strong> compound K-19(red) <strong>and</strong> K 12 [H 2 P 2 W 12 O 48 ](blue) . . . . 147<br />

3.79 Ball/stick <strong>and</strong> polyhedron representation <strong>of</strong> solid state <strong>of</strong> 19 showing 1-D<br />

chain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149<br />

3.80 FTIR spectra <strong>of</strong> compound K-20(red) <strong>and</strong> K 14 [P 2 W 19 O 69 (H 2 O)](blue) . . 150<br />

3.81 Ball/stick <strong>and</strong> polyhedral representation <strong>of</strong> polyanion 20 . . . . . . . . . . 151<br />

xvi


List <strong>of</strong> Tables<br />

1.1 Selected M-M distances (in angstrom units) <strong>of</strong> corner- <strong>and</strong> edgesharing<br />

octahedra in POMs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

1.2 List <strong>of</strong> common metal cations, M n+ , taking part in POM frameworks. We<br />

especially highlight W <strong>and</strong> Mo for being the most typical as addenda atoms 4<br />

3.1 Crystal Data <strong>and</strong> Structure Refinement for compounds CsNa-1 <strong>and</strong> CsNa-<br />

2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

3.2 Crystal Data <strong>and</strong> Structure Refinement for compound KLi-3 . . . . . . . . 36<br />

3.3 Reduction peak potentials measured from CVs <strong>and</strong> number (n 1 or n 2 ) <strong>of</strong><br />

electrons corresponding to each wave <strong>of</strong> 3, [H 2 P 4 W 24 O 94 ] 22− <strong>and</strong> [H 7 P 8 W 48 O 184 ] 33−<br />

The scan rate was 10 mV s −1 , the working electrode was glassy carbon <strong>and</strong><br />

the reference electrode was SCE . . . . . . . . . . . . . . . . . . . . . . . . 41<br />

3.4 Crystal Data <strong>and</strong> Structure Refinement for compound CsNa-4 . . . . . . 45<br />

3.5 Crystal Data <strong>and</strong> Structure Refinement for compound KNH4-5 . . . . . . 52<br />

3.6 Crystal Data <strong>and</strong> Structure Refinement for compounds Cs-6 <strong>and</strong> Cs-7 . . 58<br />

3.7 Crystal Data <strong>and</strong> Structure Refinement for compound NH4-8 . . . . . . . 72<br />

3.8 Crystal Data <strong>and</strong> Structure Refinement for compound Cs-9 . . . . . . . . 85<br />

3.9 Crystal Data <strong>and</strong> Structure Refinement for compound RbK-10 . . . . . . 90<br />

3.10 Crystal Data <strong>and</strong> Structure Refinement for compound K-11 . . . . . . . . 93<br />

3.11 Crystal Data <strong>and</strong> Structure Refinement for compound CsKNa-12 . . . . . 101<br />

3.12 Crystal Data <strong>and</strong> Structure Refinement for compounds NH4-14, NH4Na-<br />

15, RbNa-16 <strong>and</strong> KNa-17 . . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

3.13 Selected bond lengths <strong>and</strong> angles in polyanions 14 <strong>and</strong> 15 . . . . . . . . . 115<br />

3.14 Selected bond lengths <strong>and</strong> angles in polyanions 16 <strong>and</strong> 17 . . . . . . . . . 119<br />

3.15 Crystal Data <strong>and</strong> Structure Refinement for compound K-18 . . . . . . . . 144<br />

3.16 Crystal Data <strong>and</strong> Structure Refinement for compound K-19 . . . . . . . . 148<br />

3.17 Crystal Data <strong>and</strong> Structure Refinement for compound K-20 . . . . . . . . 150<br />

xvii


Chapter 1<br />

Introduction<br />

The great majority <strong>of</strong> inorganic compounds are constructed <strong>of</strong> metallic atoms as principal<br />

entities. Inorganic molecules have great potential because the number <strong>of</strong> elements<br />

in purely inorganic molecules, combined with structural diversity, make them more powerful,<br />

particularly as far as their application is concerned. In fact, the search for new<br />

properties puts more importance on the elements in a framework than on the structure<br />

itself. Polyoxometalates (POMs) are polynuclear metal oxygen clusters usually formed <strong>of</strong><br />

Mo, W or V that form a unique class <strong>of</strong> inorganic compounds because it is unmatched<br />

in terms <strong>of</strong> structural versatility <strong>and</strong> properties [1–6]. They have potential applications<br />

in many fields including medicine, catalysis, multifunctional materials, chemical analysis,<br />

imaging, etc.<br />

1.1 Historical perspectives<br />

The polyoxometalates have been known since the work <strong>of</strong> Berzelius [7] on the ammonium<br />

12-molybdophosphate in 1826, however, the study <strong>of</strong> polyoxoanion chemistry did not accelerate<br />

until the discovery <strong>of</strong> the tungstosilicic acids <strong>and</strong> their salts by Marignac [8] in<br />

1862, when analytical compositions <strong>of</strong> such heteropoly acids were precisely determined.<br />

Thereafter, the field developed rapidly, so that over 60 different types <strong>of</strong> heteropoly acids<br />

(giving rise to several hundred salts) had been described by the first decade <strong>of</strong> this century.<br />

Pauling [9] proposed a structure for 12:1 complexes based on an arrangement <strong>of</strong> twelve<br />

1


MO 6 octahedra surrounding a central XO 4 tetrahedron. After Pauling’s proposal, in the<br />

early 1930’s Keggin [10, 11] solved the structure <strong>of</strong> [H 3 PW 12 O 40 ]·5H 2 O by powder X-ray<br />

diffraction <strong>and</strong> showed that the anion was indeed based on WO 6 octahedral units as suggested,<br />

these octahedra being linked by shared edges as well as corners. The application<br />

<strong>of</strong> X-ray crystallography to the determination <strong>of</strong> polyoxometalate structures accelerated<br />

the development <strong>of</strong> polyoxometalate chemistry.<br />

Polyoxometalates are macro-molecular nanometer sized cluster composed <strong>of</strong> transition<br />

metal centers <strong>of</strong> groups V <strong>and</strong> VI in high oxidation states (mainly Mo, W <strong>and</strong> V <strong>and</strong><br />

also Nb <strong>and</strong> Ta) <strong>and</strong> normally oxygen atoms. although some derivatives with S,[12–16]<br />

F, [17] Br [18] <strong>and</strong> other p-block elements are known. In general, POMs can be described<br />

as composed <strong>of</strong> MO n units, where ‘n’ indicates the coordination number <strong>of</strong> M (n = 4,5,6<br />

or 7). Usually the distorted octahedral coordination (n = 6)is observed. Apart from M<br />

<strong>and</strong> O, other elements (heteroatoms) which are usually labeled as “X” can be part <strong>of</strong> the<br />

POM framework. As a general rule, they are tetra or hexa coordinated <strong>and</strong> lie in the<br />

center <strong>of</strong> the M x O y shell.<br />

According to their composition, the polyoxometalates can be classified in two groups:<br />

•Isopolyoxometalates, <strong>of</strong> general formula [ M m O y ] n− , which contain only metal <strong>and</strong> oxygen.<br />

•Heteropolyoxometalates, <strong>of</strong> general formula [X x M m O y ] n− , in addition to metal <strong>and</strong> oxygen,<br />

contain another element that acts as a heteroatom.<br />

In general, restrictions for heteroatoms “X” do not exist, so that around 70 elements from<br />

all the groups in the periodic Table, except the noble gases, are known to be able to play<br />

the heteroatom role. Heteroatoms “X” can be classified as: primary, which are essential<br />

to maintain the structure <strong>of</strong> the polyanion <strong>and</strong> therefore, they are not susceptible to<br />

chemical interchange; secondary, which can be eliminated to generate a stable independent<br />

polyanion, so that the combination can be considered as a coordination compound<br />

in which the polyanion acts as a lig<strong>and</strong>. More specifically, the central heteroatoms in<br />

Keggin <strong>and</strong> Wells-Dawson type polyoxometalates can be nonmetallic atom or a transition<br />

metal atoms which shows tetrahedral geometry. Nevertheless, due to the increasing<br />

diversity <strong>of</strong> structures, classified as in the literature, the unequivocal definition <strong>of</strong> this<br />

2


group <strong>of</strong> compounds becomes gradually more <strong>and</strong> more diffuse. Many reports,[19] books<br />

[4] <strong>and</strong> reviews [2, 20] have been published on this topic, showing an enormous molecular<br />

diversity amongst the inorganic family <strong>of</strong> molecules. This diversity is a consequence <strong>of</strong><br />

the rich, unprecedented <strong>and</strong> unusual properties associated to POMs. Many authors claim<br />

that they can be regarded as packed arrays <strong>of</strong> pyramidal MO 5 <strong>and</strong> octahedral MO 6 units.<br />

These entities are then the fundamental structural units <strong>and</strong> are somewhat similar to the<br />

-CH 2 - unit in organic molecules. The very important MO 6 units (<strong>and</strong> the MO 5 partner<br />

as well but to a lesser extent) are packed to form countless shapes. They join to one<br />

another apparently, in accordance with a few simple rules (as -CH 2 - does in organic molecules).<br />

Observing a representative set <strong>of</strong> POM clusters <strong>and</strong> identifying the MO 6 blocks,<br />

one can notice that the molecule as a whole is built by edge- <strong>and</strong>/or corner-sharing MO 6<br />

octahedra. Figure 1.1 shows these simple unions. The most stable unions between two<br />

Fig. 1.1: Polyhedral models <strong>of</strong> the three possible unions between two MO 6 octahedral units. A) cornersharing,B)<br />

edge-sharing <strong>and</strong> C) face-sharing<br />

octahedra are the corner- <strong>and</strong> edge-sharing models,[21–23] in which the M n+ ions are far<br />

away from each other, <strong>and</strong> their mutual repulsion is modest (Table 1.1).<br />

Table 1.1: Selected M-M distances (in angstrom units) <strong>of</strong> corner- <strong>and</strong> edgesharing octahedra in POMs.<br />

Metal(ON) corner-sharing edge-sharing<br />

W(VI) 3.7 3.4<br />

Mo(VI) 3.7 3.4<br />

V(V) 3.5 3.2<br />

In Figure 1.1C, the MO 6 octahedra are face-sharing, in such a way that the metallic<br />

centers are closer than in any <strong>of</strong> the other two cases (A, B) <strong>and</strong> at such distances, the<br />

repulsion is not balanced by the stabilization due to the chemical bonding in the 2-block<br />

unit. The latter form <strong>of</strong> union is uncommon <strong>and</strong> all the structures presented in this text<br />

only deal with derivatives containing combinations <strong>of</strong> pairs <strong>of</strong> types A <strong>and</strong> B.<br />

3


1.1.1 The clathrate-like structure<br />

Clathrate-like systems are molecular or supramolecular arrangements in which an internal<br />

unit is encapsulated by an external core. In metal-oxide polynuclear clusters, it is a<br />

common phenomenon that has been discussed by experimentalists [24–26] <strong>and</strong> computational<br />

chemists [27–29]. A formulation for molecules that accomplish the requirements <strong>of</strong><br />

a clathrate-like system was introduced in the way I-E, in which I <strong>and</strong> E are the internal<br />

<strong>and</strong> the external fragments. Keggin, Wells-Dawson <strong>and</strong> Lindqvist anions were also formulated<br />

as [XO 4 ] n− - M 12 O 36 - Keggin, [XO 4 ] n− - M 18 O 54 - Wells-Dawson or O 2− - M 6 O 18<br />

- Lindqvist<br />

1.1.2 Chemical elements taking part in POMs<br />

The addenda atoms ‘M’ have been identified as the most important entities in POMs.<br />

All the clusters included in this classification contain MO n units, so the characteristics<br />

<strong>of</strong> “M” deserve further discussion. Many ‘M’ elements are known to form octahedral<br />

coordination compounds with oxygen, but not so many can take part as MO 6 units <strong>of</strong><br />

packed polynuclear metal-oxide aggregates. Therefore, structures <strong>of</strong> polyanions appear<br />

to be governed by the electrostatic charge (‘q’) <strong>and</strong> ionic radius (‘r’) principles <strong>of</strong> metal<br />

centers, that is, only selected values <strong>of</strong> the charge/radius ratio are observed in M n+ in<br />

combination with O 2− lig<strong>and</strong>s, thus forming POMs. Few M’s are found in POM a structure<br />

since these physical limitations control the stability <strong>of</strong> the metal-oxide framework<br />

(see the values listed in Table 1.2 below).<br />

Table 1.2: List <strong>of</strong> common metal cations, M n+ , taking part in POM frameworks. We especially highlight<br />

W <strong>and</strong> Mo for being the most typical as addenda atoms<br />

Metal ion Octahedral radius (Å) Observed coordination numbers in POMs<br />

W 6+ 0.7 4, 6<br />

Mo 6+ 0.7 3 4, 6, 7<br />

V 5+ 0.68 4, 5, 6, 7<br />

Ta 5+ 0.78 6<br />

Nb 5+ 0.78 6<br />

The table contains early transition metal elements, from the left <strong>of</strong> the periodic table.<br />

In fact, there are other ions that have values <strong>of</strong> ‘q’ <strong>and</strong> ‘r’ with limits similar to those<br />

4


shown in Table 1.2. Apart from other transition metals, some p-block elements could<br />

be, at least in principle, good c<strong>and</strong>idates for being included in MO 6 units as addenda.<br />

However, charge <strong>and</strong> radius are not the only considerations to rule these assemblages <strong>of</strong><br />

metal-centerd units. Actually, an additional parameter to be considered, related to ‘M’,<br />

is the ability to form metal-oxygen pπ-dπ-bonds. This parameter affects the stability<br />

<strong>of</strong> these clusters, as well. It was observed long ago that, in octahedral MO 6 blocks,<br />

the metallic center is not in the middle <strong>of</strong> the polyhedron but somewhat displaced from<br />

the geometrical center towards one <strong>of</strong> the corners (see Figure 1.2). More precisely, it<br />

Fig. 1.2: Ball <strong>and</strong> stick represention <strong>of</strong> the α Keggin type cluster; the black balls represent tungsten <strong>and</strong><br />

the red balls represent oxygen<br />

is displaced towards the corner which is not shared with another octahedron, so that<br />

the oxygen at this position usually forms M=O double bond with the metal. (see the<br />

cluster in Figures 2 <strong>and</strong> 3). Despite all <strong>of</strong> the above, after nearly two centuries <strong>of</strong> POM<br />

chemistry, almost all the elements <strong>of</strong> the periodic table have been incorporated into a<br />

polyoxometalate framework [1]. This accounts for the chemical variability <strong>of</strong> this field.<br />

1.1.3 Features <strong>of</strong> POMs<br />

Heteropoly- <strong>and</strong> isopolyanions are routinely prepared <strong>and</strong> isolated from both aqueous<br />

<strong>and</strong> non-aqueous solutions. The most common method <strong>of</strong> synthesis involves dissolving<br />

[MO n ] m− oxoanions which after acidification assemble to yield a packed molecular array<br />

<strong>of</strong> MO 6 units. For example,<br />

5


7(MoO 4 ) 2− + 8H + → [Mo 7 O 24 ] 6− + 4H 2 O (Eq = 1)<br />

(PO 4 ) 3− + 12(WO 4 ) 2− + 24H + → [PW 12 O 40 ] 3− + 12 H 2 O (Eq = 2)<br />

(WO 4 ) 2− + H 3 PO 4 (excess) + H + → [P 2 W 18 O 62 ] 6− + P 5 W 30 (Eq = 3)<br />

Care must be taken with pH conditions so that the reaction can be controlled. The<br />

sequence in which the reagents are added to the reaction media is also important. One <strong>of</strong><br />

the most important steps in synthetic procedures <strong>of</strong> POMs are the isolation <strong>of</strong> crystals so<br />

that their features can be studied in greater depth. Clusters are precipitated or crystallized<br />

by adding countercations (alkali metals, organic cations like TBA, etc.) <strong>and</strong> subsequent<br />

separation. Usually the solid state structure <strong>of</strong> polyoxometalates is preserved in solution<br />

<strong>and</strong> an appropriate choice <strong>of</strong> the counterion allows the redissolution <strong>of</strong> a polyoxoanion in<br />

aqueous, organic or mixed aqueous/organic solvents. The structure <strong>of</strong> polyoxoanions is<br />

determined by single-crystal X-ray diffraction <strong>and</strong> the diamagnetic nature enables the use<br />

<strong>of</strong> multinuclear NMR, which is the most powerful technique for studying the structure in<br />

solutions. As far as preparation <strong>and</strong> storage conditions are concerned, it is worth noting<br />

that POMs are hydrolytically <strong>and</strong> thermally stable.<br />

1.2 Geometric Structures <strong>of</strong> representative types <strong>of</strong><br />

Polyanions: Keggin, Lindqvist, Anderson-Evans<br />

<strong>and</strong> Wells-Dawson<br />

In 1933, Keggin left his name on the structure by correctly deducing its geometry based<br />

on powder X-ray diffraction patterns. The α isomer <strong>of</strong> the Keggin structure is shown in<br />

Figure 1.3. The structure has T d symmetry <strong>and</strong> it is composed <strong>of</strong> a central heteroatom<br />

tetrahedron (XO 4 ) surrounded by twelve peripheral or “addenda” atom octahedra (MO 6 ).<br />

The peripheral MO 6 units form four trimetallic clusters (triads). Four edge ‘O’ atoms hold<br />

the trimetallic cluster together, one linking to the central heteroatom (O i - ‘i’ for inner),<br />

<strong>and</strong> three bridging peripheral MO 6 octahedra (O e - ‘e’ for edge shared). The trimetallic<br />

clusters are linked to each other by corner-shared oxygen atoms (O c - ‘c’ for corner-shared),<br />

<strong>and</strong> each metal atom has a single terminal “oxo” lig<strong>and</strong> with double bonding character<br />

(O t - ‘t’ for terminal). Thus there are four non-degenerate oxygen species present in each<br />

6


Fig. 1.3: Ball/stick <strong>and</strong> polyhedral representation <strong>of</strong> the alpha isomer <strong>of</strong> the Keggin structure with<br />

different types <strong>of</strong> oxygen. Color codes: The blue polyhedra represent the tungsten <strong>and</strong> the red balls<br />

represent oxygen<br />

Keggin unit. The Keggin unit is usually anionic, <strong>and</strong> is balanced by cations in solid state.<br />

The cations may be protons (typically coordinated with water as [H 5 O 2 ] + or [H 9 O 4 ] + ), in<br />

which case the complex is acidic, <strong>and</strong> it is identified as a Heteropolyacid. Other typical<br />

charge-balancing cations generally used are alkali metals <strong>and</strong> ammonium derivatives. As<br />

mentioned above, the tetrahedral heteroatom at the center <strong>of</strong> the Keggin unit can be any<br />

<strong>of</strong> a wide array <strong>of</strong> elemental cations. The most commonly studied structures are those<br />

containing, As V , P V , Si IV , Ge IV as the heteroatom but other transition metals <strong>and</strong> non<br />

metals have been observed playing this role. Moreover, trigonal-pyramidal (e.g. AsO 3− 3 ,<br />

Figure 1.4) or ditetrahedral (e.g. O 3 POPO 4− 3 ,Figure 1.5) hetero groups as additional<br />

building blocks allows further structural versatility. Furthermore, few types <strong>of</strong> addenda<br />

atoms have been observed in Keggin structures; typically, the addenda atoms are molybdenum<br />

or tungsten. Vanadium based structures are also reported, but its questionable<br />

whether they form 1:12 heteroatom:addenda complexes; it appears more likely that they<br />

form Keggin-like 1:13 or 1:14 complexes. Clusters with p-block elements (P, Si, Al, Ga <strong>and</strong><br />

Ge), transition metal elements (Fe(II/III), Co(I/II), Ni(II/IV), Zn(II)), <strong>and</strong> even two H +<br />

have been synthesized. This position can be either tetrahedrally coordinated (as in Keggin<br />

<strong>and</strong> Wells-Dawson anions) or octahedrally coordinated (as in the Anderson structure).<br />

Figure 1.6 shows a [M 6 O 19 ] n− isopolyanion (A) <strong>and</strong> two heteropolyanions <strong>and</strong> Figures 1.7<br />

<strong>and</strong> 1.8 illustrates two different types <strong>of</strong> heteropolyanions. The [M 6 O 19 ] n− isopolyanion<br />

7


Fig. 1.4: Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong> [Cu 3 Na 3 (H 2 O) 9 (α-As III W 9 O 33 ) 2 ] 9−<br />

Fig. 1.5: Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong> {(O 3 POPO 3 )Mo 6 O 18 } 4−<br />

representative <strong>of</strong> the Lindqvist type structure [30]is characterized by the presence <strong>of</strong> one<br />

terminal M-O bond at each metal center <strong>and</strong> thus, they belong to the type-I category<br />

in Pope’s classification scheme. Each metal center is octahedrally coordinated <strong>and</strong> the<br />

structure consists <strong>of</strong> a compact assemblage <strong>of</strong> edgesharing octahedra, leading to an overall<br />

octahedral cluster, with full Oh symmetry. The pronounced distortion <strong>of</strong> each MO 6 unit<br />

<strong>and</strong> the three different oxygen coordination sites are shown in Figure 1.6. One can see<br />

8


Fig. 1.6: Ball <strong>and</strong> stick representation <strong>of</strong> the Lindqvist structure<br />

that the metals occupy equivalent octahedral sites, so that they make one short terminal<br />

M-O bond <strong>and</strong> a rather long M-O bond to the translocated high-coordinate oxygen<br />

site. This cluster is known as the Lindqvist type structure, e.g. [M 6 O 19 ] n− (n= 8 for<br />

M= Nb, Ta) <strong>and</strong> (n = 2 for M= Mo, W ). In Figure 1.7, the polyanion is based on an<br />

Fig. 1.7: Polyhedron representation <strong>of</strong> the Anderson structure<br />

arrangement <strong>of</strong> seven edge shared octahedra where the heteroatom “X” is surrounded<br />

by six-oxo lig<strong>and</strong>s in a pseudo-octahedral symmetry <strong>and</strong> it occupies the central position.<br />

9


This planar structure was originally proposed by Anderson for the heptamolybdate anion,<br />

[Mo 7 O 24 ] 6− , but it was observed for the first time when Evans reported the structure <strong>of</strong><br />

[TeMo 6 O 24 ] 6− , which was isostructural to that <strong>of</strong> the 6-molybdo-anion [IMo 6 O 24 ] 5− , It<br />

is generally know as ‘Anderson-Evans’ structure [31]. In Figure 1.8, the complete anion<br />

Fig. 1.8: Ball <strong>and</strong> stick representation <strong>of</strong> the Wells-Dawson structure<br />

consist <strong>of</strong> two identical ‘half-units’ related by a plane <strong>of</strong> symmetry perpendicular to the<br />

trigonal axis. The ‘half-units’ are linked together by six oxygen atoms situated in the<br />

plane <strong>of</strong> symmetry, so that these atoms are shared equally by the two halves. The two<br />

heteroatoms ‘X’ are tetrahedrally coordinated <strong>and</strong> surrounded by nine WO 6 octahedra<br />

linked together by edge sharing <strong>and</strong> corner sharing. This type <strong>of</strong> cluster is generally<br />

known as ‘Wells-Dawson structure’ [32].<br />

1.3 Lacunary Species<br />

Although all polyoxoanions are ultimately decomposed to monomeric species in basic solution,<br />

controlling the pH can lead to the formation <strong>and</strong> isolation <strong>of</strong> ‘defect’ or ‘lacunary’<br />

structures. These are best illustrated with the α isomers <strong>of</strong> the Keggin-type polyanion,<br />

the Keggin anion has five different rotational isomers known as ‘Baker-Figgis’(α,β,γ,δ,ɛ).<br />

10


β,γ,δ <strong>and</strong> ɛ are derived from the α-isomer, a cluster already mentioned above, by rotation<br />

<strong>of</strong> 60 ◦ <strong>of</strong> 1, 2, 3, <strong>and</strong> 4 different triads respectively. which can form [XM 11 O 39 ] n− <strong>and</strong><br />

Fig. 1.9: Schematic representation <strong>of</strong> the formation <strong>of</strong> Keggin-type lacunary. Courtesy: Dr. Santiago C.<br />

Reinoso<br />

[XM 9 O 34 ] n− species by the ‘removal’ <strong>of</strong> one or three adjacent MO 6 octahedra respectively.<br />

An intermediate structure in which two adjacent octahedra have been removed has not<br />

been observed for the α <strong>and</strong> β isomers but observed only for the γ isomer; in the former<br />

two cases the structure would contain an MO 6 octahedron with three fac terminal oxygen<br />

atoms <strong>and</strong> is therefore expected to be quite reactive. Two kinds <strong>of</strong> the XM 9 anions can be<br />

formed, both having C 3v symmetry: A-type, in which three corner-shared octahedra have<br />

been lost, <strong>and</strong> B-type, in which three edge-shared octahedra have been lost. Neither <strong>of</strong><br />

these sructures appears to be thermodynamically stable in solution, although both have<br />

been isolated as solids salts, <strong>and</strong> used to synthesize derivatives. In addition, both A-<br />

<strong>and</strong> B-type XM 9 structural units are recognizable in larger polyoxoanions. Examples are:<br />

11


[X 2 M 18 O 62 ] 6− , the D 3h Wells-Dawson structure formed by fusion <strong>of</strong> two A-XM 9 units;<br />

[As 2 W 21 O 69 ] 6− <strong>and</strong> [P 2 W 21 O 71 ] 6− , which contain two B- <strong>and</strong> A-type XW 9 units respectively<br />

separated by a ‘belt’ <strong>of</strong> three WO 6 octahedra; <strong>and</strong> [(NH 4 )As 4 W 40 O 140 ] 27− , containing<br />

four B-type AsW 9 units. The large tungstoarsenate(III) [As 6 W 65 O 217 (H 2 O) 7 ] 26− is the<br />

only example <strong>of</strong> a polyanion that contains both (α-AsW 9 O 33 ) <strong>and</strong> (β-AsW 9 O 33 ) isomers<br />

in the same structure [33]. Many <strong>of</strong> these larger structures can be further converted to<br />

lacunary species, e.g. X 2 M 18 (X= P, As) (‘Wells-Dawson structure’) yields X 2 M 17 , X 2 M 15 ,<br />

<strong>and</strong> X 2 W 12 anions.<br />

Salts <strong>of</strong> the anions [AsW 9 O 33 ] 9− <strong>and</strong> [SbW 9 O 33 ] 9− were first reported by Rosenheim, <strong>and</strong><br />

structurally confirmed by Krebs [34] <strong>and</strong> Tourné [35]. The anions appear to be stable at<br />

pH 7.5 to 9.0 <strong>and</strong> the potassium salts crystallize in a face centered cubic form isotypic<br />

with several lacunary Keggin salts mentioned earlier. This suggests an incomplete Keggin<br />

structure probably <strong>of</strong> B-type statistically oriented in the cubic cell. In contrast to the<br />

A-type 1:9 anions (Si, Ge, etc.), As III W 9 <strong>and</strong> Sb III W 9 do not react with tungstate to<br />

give As III W 11 <strong>and</strong> Sb III W 11 . However, solutions <strong>of</strong> the anions react with electrophiles to<br />

give complexes.<br />

These lacunary species derived from the Keggin type structure, are obtained from the<br />

three Baker-Figgis (α, β, γ) isomers by means <strong>of</strong> the elimination <strong>of</strong> a variable number <strong>of</strong><br />

octahedra, so that a total <strong>of</strong> nine species are known whose structures are shown in above<br />

Figure 1.9<br />

In the case <strong>of</strong> phospho-tungstate system, the reaction patterns are similar to those <strong>of</strong><br />

the silico-tungstates, but a series <strong>of</strong> remarkable differences exists, which can be visualized<br />

from the Figure 1.10<br />

The β-isomers are much more unstable <strong>and</strong> isomerize to the α-isomers, with the exception<br />

<strong>of</strong> the trivacant species A- β-[PW 9 O 34 ] 9− , which forms on acidifying solutions <strong>of</strong><br />

tungstate <strong>and</strong> phosphate to pH around 9. When the reaction is carried out at 0 ◦ C, the<br />

A-α-isomer is obtained but it isomerizes to the A-β-isomer at room temperature. In addition,<br />

the dilacunary species γ-[PW 10 O 36 ] 7− is not synthesized directly, but by overnight<br />

refluxing <strong>of</strong> polyanion [P 2 W 5 O 23 ] 6− , overnight at pH = 7 <strong>and</strong> it is only isolated with<br />

12


Fig. 1.10: Schematic representation <strong>of</strong> the formation <strong>of</strong> Keggin- <strong>and</strong> Wells-Dawson type lacunary species<br />

in tungsto-phosphate system with respect to pH.Courtesy: Dr. Santiago C. Reinoso<br />

a cesium counterion. On the other h<strong>and</strong>, a series <strong>of</strong> heteropolyanions can be obtained<br />

which has no equivalent in silico-tungstate system. Thus, when the trivacant species<br />

A-α-[PW 9 O 34 ] 9− is acidified in excess <strong>of</strong> potassium counterion, instead <strong>of</strong> the Keggin<br />

ion formation, association <strong>of</strong> two trivacant species takes place by means <strong>of</strong> a bridging<br />

tungsten in the belt. If the pH is decreased different heteropolyanions are obtained,<br />

13


[P 2 W 19 O 68 (HO)] 14− , [P 2 W 20 O 70 (H 2 O) 2 ] 6− <strong>and</strong> [P 2 W 21 O 71 (HO) 3 ] 6− , which show one, two<br />

<strong>and</strong> three {WO(HO)} 4+ groups in the central belt, respectively. In addition, when a<br />

solution <strong>of</strong> tungstate is acidified until pH below 2 in the presence <strong>of</strong> an excess <strong>of</strong> phosphate,<br />

mixture <strong>of</strong> α <strong>and</strong> β-isomers <strong>of</strong> the Wells-Dawson [P 2 W 18 O 62 ] 6− heteropolyanion<br />

is obtained instead <strong>of</strong> the Keggin. The Wells-Dawson heteropolyanion shows a similar<br />

behaviour to that <strong>of</strong> the Keggin, since the basicity <strong>of</strong> its dissolution produces hydrolytic<br />

cleavage <strong>of</strong> the M-O bonds to give rise to monovacant [P 2 W 17 O 61 ] 10− lacunary species,<br />

only if the pH is in between 4 <strong>and</strong> 6, <strong>and</strong> to the trilacunary [P 2 W 15 O 56 ] 12− at pH around<br />

10. These trivacant species undergo irreversible processes <strong>of</strong> transformation to form the<br />

hexa lacunary species [P 2 W 12 O 48 ] 14− . Lacunary anions with more than one surface ‘vacancy’<br />

may also be derivatized by cation complexation. Reaction <strong>of</strong> a stable, lacunary<br />

polyoxometalate with transition metal ions usually leads to a product with the unchanged<br />

heteropolyanion framework. Depending upon the coordination requirement <strong>and</strong> the size<br />

<strong>of</strong> a given transition metal ion, the geometry <strong>of</strong> the reaction product can therefore <strong>of</strong>ten<br />

be predicted. At the same time it must be pointed out that the mechanism <strong>of</strong> formation<br />

<strong>of</strong> polyoxometales is not well understood <strong>and</strong> commonly described as self assembly.<br />

Therefore, the synthesis <strong>of</strong> polyoxoanions with novel shapes <strong>and</strong> sizes is a very difficult<br />

task. But, according to Müller <strong>and</strong> Pope, POM structure are governed by two general<br />

principles.<br />

•Polyanions are generated by linking MO n polyhedra via corners <strong>and</strong> edges leading to<br />

different types <strong>of</strong> faces on the surfaces.<br />

•Each metal atom forms an MO n coordination polyhedron (most commonly an octahedron<br />

or a square pyramid) in which the metal atoms are displaced, as a result <strong>of</strong> M-Oπ<br />

bonding, towards the terminal polyhedral vertices forming the surface <strong>of</strong> the structure.<br />

Transition metal substituted polyoxometales can also be <strong>of</strong> interest owing to their magnetic<br />

properties. Structures which contain more than one paramagnetic transition metal<br />

ion in close proximity may exhibit exchange-coupled spins leading to large spin ground<br />

states [36, 37]. The polyoxometalate matrix may be considered as a diamagnetic host<br />

encapsulating <strong>and</strong> thereby isolating a magnetic cluster <strong>of</strong> transition metals. Polyoxometalate<br />

chemistry has received a great research interest in the area <strong>of</strong> oxidation catalysis<br />

14


<strong>and</strong> most publications in the field deal with the evaluation <strong>of</strong> the catalytic activity <strong>of</strong><br />

polyoxoanion salts (oxidation catalysis) or the free acids (acid catalysis) [3–5, 38]. Polyoxoanions<br />

have been shown to activate small molecules (e.g. O 2 , H 2 O 2 ) which are highly<br />

desired oxidants in the chemical industries for the catalysis <strong>of</strong> organic reactions (e.g. epoxidation,<br />

hydroxylation). Currently polyoxoanions are being used as catalysts in different<br />

processes on an industrial scale worldwide. Substitution <strong>of</strong> one or more addenda atoms<br />

by redox-active transition metal ions (e.g. Fe 3+ , Ru 3+ ) allows to fine tune the catalytic<br />

activity <strong>of</strong> polyoxoanions [39].<br />

The interactions <strong>of</strong> polyanions with enzymes <strong>and</strong> other biomolecules is also being studied<br />

<strong>and</strong> it allows for a better underst<strong>and</strong>ing <strong>of</strong> the antitumor/viral activities <strong>of</strong> this class<br />

<strong>of</strong> compounds [40–42]. It is also possible to graft organic groups to the surface <strong>of</strong> the<br />

polyoxoanions via incorporation <strong>of</strong> an organo-metal (PhSn) or an organo non-metal (e.g.<br />

RSi)[43] fragment in a lacunary polyoxometalate precursor. The monoorgano tin chemistry<br />

<strong>of</strong> polyoxometalates has been studied mainly by Pope <strong>and</strong> by few other groups.<br />

It is known that the size, shape <strong>and</strong> charge density <strong>of</strong> many polyoxoanions are <strong>of</strong> interest<br />

for pharmaceutical applications. However, the mechanism <strong>of</strong> action <strong>of</strong> many polyoxoanions<br />

is not selective towards a specific target. In order to improve selectivity it appears<br />

desirable slightly to modify a given polyoxoanion core structure . However, such attempts<br />

frequently result in a different polyoxoanion framework. Therefore the most straightforward<br />

<strong>and</strong> promising approach towards systematic derivatization <strong>of</strong> polyoxoanions involves<br />

attachment <strong>of</strong> organic groups to the surface <strong>of</strong> the metal-oxo framework. In order to be<br />

attractive for pharmaceutical applications, the functionalized polyoxoanions should be<br />

water-soluble <strong>and</strong> fairly stable at physiological pH.<br />

To date the number <strong>of</strong> water-soluble polyoxoanions with tightly bound organic functionalities<br />

is rather small. Pope <strong>and</strong> co-workers [44–47] were the first to study the interaction<br />

<strong>of</strong> monoorganotin groups (e.g. n-C 4 H 9 Sn 3+ , C 6 H 5 Sn 3+ ) <strong>and</strong> polyoxoanions. They reacted<br />

different organotin halide precursors (e.g. n-C 4 H 9 SnCl 3 , C 6 H 5 SnCl 3 ) with a large number<br />

<strong>of</strong> lacunary heteropolytungstates in aqueous solution <strong>and</strong> they were able to identify<br />

novel (mostly dimeric) polyoxoanion structures. Single-crystal X-ray diffraction studies<br />

revealed that these compounds contain tightly anchored organotin fragments. The com-<br />

15


Fig. 1.11: Monomeric (left) <strong>and</strong> dimeric species (right) <strong>of</strong> monoorganotin containing polyoxotungstates.<br />

pounds were also studied in solution by multinuclear NMR spectroscopy. This technique<br />

is also very valuable in the evaluation <strong>of</strong> the stability <strong>of</strong> polyoxoanions at physiological<br />

pH <strong>and</strong> low concentration for medicinal applications. Liu <strong>and</strong> coworkers [48–52] have also<br />

synthesized some monoorganotin substituted polyoxotungstates. They introduced ester<br />

functionalities to the organotin fragments <strong>and</strong> tested the biological (antitumor) activity<br />

<strong>of</strong> their products. Most recently, the groups <strong>of</strong> Pope <strong>and</strong> Hasenknopf showed independently<br />

that monoorganotin containing polyanions can be further derivatized by peptide<br />

or ester functions [53–55]. Haiduc et al. also reported on monoorganotin derivatives <strong>of</strong><br />

polyoxotungstates [56]. However, the proposed structures <strong>of</strong> Liu <strong>and</strong> Haiduc need to be<br />

confirmed by X-ray diffraction.<br />

Interestingly, there were no reports on diorganotin substituted polyoxotungstates. Such<br />

species could allow for more flexibility for the interaction with <strong>and</strong> binding to other organic<br />

compounds <strong>and</strong> biomolecules. Therefore we investigated the interaction <strong>of</strong> di-organotin<br />

groups with the entire arsenal <strong>of</strong> lacunary polyoxotungstates.<br />

Introduction <strong>of</strong> organic groups into the POM framework could greatly increase the number<br />

<strong>of</strong> compounds available for screening, together with a potential modulation <strong>of</strong> essential<br />

features such as stability, bioavailability, recognition, etc [57]. While there is a signifi-<br />

16


cant amount <strong>of</strong> work on hybrid polyoxometales,[58–60] there have been only few reports<br />

<strong>of</strong> the reactivity <strong>of</strong> the side chain <strong>of</strong> such functionalized POMs [61–63]. Many <strong>of</strong> these<br />

organically derivatized POMs are unstable in water. Starting from hydrolytically more<br />

stable cyclopentadienyltitanium substituted POMs, [64, 65] Keana reported the preparation<br />

<strong>and</strong> reactivity <strong>of</strong> derivatives with various functional groups. [66, 67] We decided to<br />

re-examine hetropolytungstates with an alkyltin group, because these compounds might<br />

have promising antitumor activities [68]. It is important to have access to a wide variety<br />

<strong>of</strong> organic groups on the POM to meet specific requirements in biological applications.<br />

Different groups like -NH 2 for instance will enable the POM to cross a membrane or to<br />

reach a specific receptor depending on their precise nature.<br />

17


Chapter 2<br />

Experimental<br />

2.0.1 Reagents<br />

All chemicals were purchased from well known chemical companies, <strong>and</strong> used as received:<br />

(CH 3 ) 2 SnCl 2 <strong>and</strong> InCl 3 (anhydrous)was purchased from Fluka Chemie, TiOSO 4 was purchased<br />

from E.Merck AG, C 6 H 5 SnCl 3 was purchased from Aldrich Chem.Co. D 2 O was<br />

purchased from AppliChem. CdCl 2·H 2 O, was purchased from Riedel-de haën.The purity<br />

<strong>of</strong> the dimethyl tin was 95% <strong>and</strong> that <strong>of</strong> monophenyl tin was 98%.<br />

2.1 Instrumentation<br />

2.1.1 Infrared spectroscopy<br />

Infrared spectra with 4 cm −1 resolution were recorded on a Nicolet Avatar 370 FT-IR<br />

spectrophotometer as KBr pellet samples. The following abbreviation was used to assign<br />

the peak intensities: w = weak; m = medium; s = strong; vs = very strong; b = broad;<br />

sh = shoulder.<br />

2.1.2 Single crystal X-ray diffraction<br />

X-ray diffraction data collection was carried out on a Bruker D8 SMART APEX CCD<br />

single crystal diffractometer equipped with a sealed Mo anode tube.The SHELX s<strong>of</strong>tware<br />

package was used in order to solve <strong>and</strong> refine the structures. Direct method solutions<br />

18


located the heaviest atoms <strong>and</strong> remaining atoms were found in subsequent Fourier difference<br />

syntheses. Refinements were full-matrix least-squares on F 2 for structures having not<br />

more than 1200 parameters, <strong>and</strong> were block-diagonal least-squares on F 2 for structures<br />

having more than 1200 parameters. Routine Lorentz <strong>and</strong> polarization corrections were<br />

applied <strong>and</strong> absorption corrections were performed using the SADABS program. a R =<br />

∑ ||Fo |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

2.1.3 Multinuclear magnetic resonance spectroscopy<br />

NMR spectra were recorded on a JEOL 400 ECX spectrometer operating at 9.39 T (400<br />

MHz for proton) magnetic field. The resonance frequencies were 161.834 MHz for 31 P,<br />

149.081 MHz for 119 Sn <strong>and</strong> 16.656 MHz for 183 W. Chemical shifts are given with respect<br />

to external st<strong>and</strong>ard 85% H 3 PO 4 for 31 P, (C 2 H 5 ) 4 Sn for 119 Sn, 2 M Na 2 WO 4 for 183 W. All<br />

aqueous 183 W NMR spectra were collected on highly concentrated solution.<br />

2.1.4 Cyclic voltametry<br />

All cyclic voltammogramms (CV) were recorded on a EG & G 273 A voltameter driven by<br />

a PC with the M270 s<strong>of</strong>tware. Potentials are coated against a saturated calomel electrode<br />

(SCE). The counter electrode was a platinum gauze <strong>of</strong> large surface area. All experiments<br />

were performed at room temperature.<br />

2.1.5 Elemental analyses <strong>and</strong> thermogravimetric analysis<br />

All elemental analyses were performed by Kanti Labs Ltd. in Missisunga, Canada. Thermogravimetric<br />

analysis (TGA): Water contents were determined using a TGA Thermalgravimetric<br />

Analyzer with 15-20 mg samples in 100 µL alumina pans, under a 100 mL<br />

min −1 N 2 flow <strong>and</strong> with heating rates <strong>of</strong> 5 ◦ C min −1 .<br />

19


2.2 Preparation <strong>of</strong> starting materials<br />

2.2.1 Synthesis <strong>of</strong> Na 9 [AsW 9 O 33 ]·27H 2 O<br />

To a hot (∼95 ◦ C) solution <strong>of</strong> 110g <strong>of</strong> Na 2 WO 4·2H 2 O in 117 mL <strong>of</strong> distilled water, 3.67<br />

g <strong>of</strong> As 2 O 3 were added. Then 27.7 mL concentrated HCl were added dropwise with in<br />

2 minutes with continuous stirring for a period <strong>of</strong> 10 minutes. The solution was cooled<br />

for a while <strong>and</strong> than filtered into a beaker <strong>and</strong> covered with parafilm. The formation <strong>of</strong><br />

crystals starts when the solution cools in the period <strong>of</strong> time. The filtrate was left in an<br />

open beaker until the solution reached the mark <strong>of</strong> crystals. The crystals were collected<br />

in a bruchner funnel <strong>and</strong> dried at ∼50 ◦ C in an air oven overnight. The final compound<br />

was characterized by FTIR spectroscopy <strong>and</strong> compared to the reported spectrum [69].<br />

2.2.2 Synthesis <strong>of</strong> Na 9 [SbW 9 O 33 ]·27H 2 O<br />

Solution A : 1.96 g <strong>of</strong> Sb 2 O 3 were dissolved in 10 mL conc. HCl (may or may not<br />

dissolve completely). Solution B : 40g <strong>of</strong> Na 2 WO 4·2H 2 O were dissolved in 80 mL <strong>of</strong><br />

distilled water (∼90 ◦ C). Solution A was transferred into the beaker containing Solution<br />

B. including the non dissolved Sb 2 O 3 drop wise <strong>and</strong> then, the whole reaction mixture was<br />

refluxed for approximately an hour. The solution was cooled <strong>and</strong> filtered, crystals start<br />

forming immediately. Solution was left open until the it reached the level <strong>of</strong> crystals. The<br />

final compound was characterized by FTIR spectroscopy <strong>and</strong> compared to the reported<br />

spectrum [70].<br />

2.2.3 Synthesis <strong>of</strong> K 14 [As 2 W 19 O 67·(H 2 O)]<br />

To a solution <strong>of</strong> 94g (285 mmol) <strong>of</strong> Na 2 WO 4·2H 2 O in 250 mL <strong>of</strong> distilled water. 4.45g<br />

(22.5 mmol) <strong>of</strong> As 2 O 3 were added subsequently. The solution was stirred for few minutes<br />

<strong>and</strong> pH was adjusted to 6.3 by addition <strong>of</strong> 12 M HCl (37%).The solution was heated to<br />

∼80 ◦ C for 10 mins, <strong>and</strong> after cooling 35g (45 mmol) <strong>of</strong> KCl was added to the solution at<br />

room temperature. The solution was stirred again for 15 mins <strong>and</strong> the formed precipitate<br />

was filtered <strong>of</strong>f <strong>and</strong> dried at ∼80 ◦ C in an air oven overnight. It was characterized by<br />

FTIR spectroscopy <strong>and</strong> compared to the reported spectrum [71].<br />

20


2.2.4 Synthesis <strong>of</strong> A & B Na 8 [HAsW 9 O 34 ]·11H 2 O (A & B-Type<br />

AsW 9 O 34 )<br />

30 g <strong>of</strong> Na 2 WO 4·2H 2 O <strong>and</strong> 2.3 g <strong>of</strong> As 2 O 5 were dissolved in 40 mL <strong>of</strong> distilled water with<br />

stirring. Glacial CH 3 COOH was added drop wise until the pH changed to 8.1 or 8.3 the<br />

solution turned milky on addition <strong>of</strong> CH 3 COOH but after stirring for a period <strong>of</strong> 30 min<br />

a heavy white precipitate was formed. The precipitate was filtered on a frit <strong>and</strong> air dried.<br />

The product obtained is A-AsW 9 O 34 . If the above product is kept in the oven for a period<br />

<strong>of</strong> 2hrs at ∼140 ◦ C it isomerizes to give B-AsW 9 O 34 . These synthesis was done with slight<br />

modification to the published method. They were characterized by FTIR spectroscopy<br />

<strong>and</strong> compared to the A & B reported spectrum [72].<br />

2.2.5 Synthesis <strong>of</strong> A-Na 9 [PW 9 O 34 ]·7H 2 O<br />

120 g (0.36 mol) <strong>of</strong> Na 2 WO 4·2H 2 O were dissolved in 150 g <strong>of</strong> distilled water. H 3 PO 4 (85%)<br />

was added dropwise with stirring. After the completion <strong>of</strong> addition, the pH <strong>of</strong> the solution<br />

was measured to be 8.9. Glacial CH 3 COOH was added dropwise with vigorous stirring.<br />

Large quantities <strong>of</strong> white precipitate were formed during the addition. The final pH <strong>of</strong><br />

the solution was 7.8. The solution was stirred at least for an hour <strong>and</strong> the precipitate<br />

was collected on a medium frit. Heating <strong>of</strong> the crude product at ∼120 ◦ C induces a solid<br />

state isomerization from A-type to B-type. These compounds were characterized by FTIR<br />

spectroscopy <strong>and</strong> 31 P-NMR spectroscopy <strong>and</strong> compared to the FTIR reported spectrum<br />

[73].<br />

2.2.6 Synthesis <strong>of</strong> Cs 6 [P 2 W 5 O 23 ]·H 2 O<br />

60 g ( 0.24 mol) <strong>of</strong> H 2 WO 4 was slurried with 200 g <strong>of</strong> water. Approximately 110 mL<br />

<strong>of</strong> a 50% aqueous CsOH solution was added drop wise with vigorous stirring. The turbid<br />

solution was filtered through a cake <strong>of</strong> 10 g celite <strong>and</strong> to the clear colorless filtrate<br />

H 3 PO 4 (85%) was added drop wise while stirring to adjust the pH to 7.0. The solution<br />

was again stirred for an hour <strong>and</strong> filtered again. The filtrate was cooled to ∼0 ◦ C in<br />

refrigerator for 24 hours <strong>and</strong> the white crystalline solid was formed which was identified<br />

21


as Cs 6 [P 2 W 5 O 23 ]·7H 2 O by FTIR spectroscopy [73].<br />

2.2.7 Synthesis <strong>of</strong> Cs 7 [PW 10 O 36 ]·H 2 O<br />

75 g <strong>of</strong> Cs 6 [P 2 W 5 O 23 ]·H 2 O synthesized by the above procedure was dissolved in 150 g <strong>of</strong><br />

water <strong>and</strong> the resulting was refluxed for 24 hour. The solution was filtered hot through a<br />

medium frit to obtain the crude product Cs 7 [PW 10 O 36 ]·H 2 O. The filtrate was cooled for<br />

48 hr at ∼0 ◦ C, <strong>and</strong> filtered to recover the unconverted Cs 6 [P 2 W 5 O 23 ]·H 2 O, which was<br />

again used to synthesize Cs 7 [PW 10 O 36 ]·H 2 O. The product was characterized by FTIR<br />

spectroscopy [73].<br />

2.2.8 Synthesis <strong>of</strong> Na 20 [P 6 W 18 O 79 ]·37.5H 2 O<br />

To a solution <strong>of</strong> 50 g <strong>of</strong> Na 2 WO 4·2H 2 O in 50 mL <strong>of</strong> distilled water. 3.5 mL <strong>of</strong> (85%)<br />

H 3 PO 4 was added to the solution subsequently. This resulting solution was boiled until<br />

the final volume reached 50 mL. The solution was cooled <strong>and</strong> on cooling white crystalline<br />

precipitate was formed, which was recrystallized from water to get pure crystalline material.<br />

It was than characterized by FTIR <strong>and</strong> 31 P-NMR spectroscopies <strong>and</strong> was compared<br />

with the published data [74].<br />

2.2.9 Synthesis <strong>of</strong> K 12 [H 2 P 2 W 12 O 48 ]·24H 2 O<br />

83 g <strong>of</strong> K 6 [P 2 W 18 O 62 ]·X H 2 O was dissolved in 300 mL <strong>of</strong> distilled water, <strong>and</strong> than a<br />

solution <strong>of</strong> 48.4 g (0.4 mol) <strong>of</strong> tris base in 200 mL water was added. The solution was left<br />

at room temperature for 30 minutes <strong>and</strong> then 80 g <strong>of</strong> solid KCl was added. After complete<br />

dissolution, a solution <strong>of</strong> 55.3 g (0.4 mol) <strong>of</strong> K 2 CO 3 in 200 mL <strong>of</strong> water was added. The<br />

resulting mixture was stirred for 15 minutes <strong>and</strong> a white precipitate appeared in due<br />

course <strong>of</strong> stirring. It was collected on a coarse sintered frit, dried under suction for 12<br />

hours <strong>and</strong> washed couple <strong>of</strong> times with ethanol. The precipitate was then air dried for 3<br />

days. It was characterized by FTIR <strong>and</strong> 31 P-NMR spectroscopies <strong>and</strong> was compared with<br />

the published data [75].<br />

22


2.2.10 Synthesis <strong>of</strong> K 16 Li 2 [H 6 P 4 W 24 O 94 ]·33H 2 O<br />

8 g <strong>of</strong> K 12 [α-H 2 P 2 W 12 O 48 ]·24H 2 O was dissolved in 250 mL <strong>of</strong> 1 M aqueous solution <strong>of</strong> LiCl<br />

acidified by 0.7 mL <strong>of</strong> CH 3 COOH. The solution was left for 4 hours at room temperature<br />

<strong>and</strong> then 50 mL <strong>of</strong> saturated KCl solution was added. The formed white precipitate was<br />

filtered <strong>of</strong>f <strong>and</strong> washed with KCl solution <strong>and</strong> twice with ethanol. The precipitate was air<br />

dried overnight <strong>and</strong> It was than characterized by FTIR <strong>and</strong> 31 P-NMR spectroscopies <strong>and</strong><br />

was compared with the published data [76].<br />

2.2.11 Synthesis <strong>of</strong> K 28 Li 5 [H 7 P 8 W 48 O 184 ]·92H 2 O<br />

28 g <strong>of</strong> K 12 H 2 P 2 W 12 O 48·24H 2 O was dissolved in 1 litre <strong>of</strong> a mixture <strong>of</strong> LiCl (0.5 mol),<br />

LiOAc (0.5 mol) <strong>and</strong> CH 3 COOH (0.5 mol) in water. The solution was left open instead <strong>of</strong><br />

being closed as published earlier. After 1 day white crystalline needles started appearing.<br />

The solution was left for a week for further crystallization. The crystalline material was<br />

filtered in a frit <strong>and</strong> kept for air drying. It was than characterized by FTIR <strong>and</strong> 31 P-NMR<br />

spectroscopies <strong>and</strong> was compared with the published data [76].<br />

2.2.12 Synthesis <strong>of</strong> Na 27 [NaAs 4 W 40 O 140 ]·60H 2 O<br />

132 g (0.4 mol) <strong>of</strong> Na 2 WO 4·2H 2 O <strong>and</strong> 5.2 g (40 mmol) Na 2 AsO 3 were dissolved in 200 mL<br />

<strong>of</strong> distilled water at ∼80 ◦ C. 6 M HCl was added slowly with vigorous stirring until the<br />

final pH reached to 4.0. The solution was cooled <strong>and</strong> kept in refrigerator at ∼80 ◦ C , the<br />

solid crystalline product crystallizes slowly. One day is required for complete deposition.<br />

The white solid is collected on a filter paper <strong>and</strong> air dried. The synthesized product was<br />

characterized using FTIR spectroscopy <strong>and</strong> was compared with the published data [77].<br />

2.2.13 Synthesis <strong>of</strong> K 8 [γ-SiW 10 O 36 ]·20H 2 O<br />

To prepare γ-SiW 10 , at first β 2 -SiW 11 was prepared as follows 5.5 g (25 mMol)<strong>of</strong> Na 2 SiO 3·5H 2 O<br />

was dissolved in 50 mL <strong>of</strong> distilled water. (solution A) 91 g (0.25 mMol)<strong>of</strong> Na 2 WO 4·2H 2 O<br />

was dissolved in 150 mL <strong>of</strong> H 2 O in 1 liter beaker.(Solution B) 82.5 mL <strong>of</strong> 4 M HCl was<br />

added in 1 mL portion over 10 min. Solution A was added to solution B <strong>and</strong> the pH<br />

23


was adjusted between 5-6 by addition <strong>of</strong> 4 M HCl for 100 min.( 1hr, 40 min) Later 4.45<br />

g solid KCl was added to the resulting solution <strong>and</strong> stirred for 15 min to obtain solid<br />

white precipitate. This uncharacterized product is K salt <strong>of</strong> βSiW 11 . In order to obtain<br />

γ-SiW 10 , 15 g <strong>of</strong> K salt <strong>of</strong> βSiW 11 was dissolved in 150 mL <strong>of</strong> distilled water. Insoluble<br />

impurity was removed by filtering on a frit containing celite. The pH <strong>of</strong> the solution was<br />

adjusted to 9.1 with 2 M aqueous solution <strong>of</strong> K 2 CO 3 solution. The pH <strong>of</strong> this solution<br />

was kept at this value for exactly 16 minutes. 85 gm <strong>of</strong> KCl was added to it <strong>and</strong> stirred<br />

for 10 minutes. The pH was still maintained at 9.1. by adding 2 M aqueous solution <strong>of</strong><br />

K 2 CO 3 solution. 40 g <strong>of</strong> solid KCl was added to obatin K-salt <strong>of</strong> γ-SiW 10 . The precipitate<br />

was dried overnight at 50 degree in hot air oven. It was characterized using FTIR<br />

spectroscopy <strong>and</strong> was compared with the published data [78].<br />

2.2.14 Synthesis <strong>of</strong> K 7 [PW 11 O 39 ]·14H 2 O<br />

To a solution <strong>of</strong> 181.5 g Na 2 WO 4·2H 2 O in 300 mL H 2 O was slowly added 50 mL <strong>of</strong> 1M<br />

H 3 P0 4 <strong>and</strong> 88 mL glacial acetic acid. This solution was refluxed for 1 hour <strong>and</strong> then<br />

cooled to room temperature. Addition <strong>of</strong> 60 g solid KCl leads to white precipitate which<br />

was collected after 10 minutes <strong>and</strong> air dried. The solid product was characterized using<br />

FTIR spectroscopy <strong>and</strong> was compared with the published data [79].<br />

2.2.15 Synthesis <strong>of</strong> K 14 [P 2 W 19 O 69 (H 2 O)]·24H 2 O<br />

The precursor was synthesized from the mixtures <strong>of</strong> K 7 PW 11 O 39 (aq) <strong>and</strong> Na 8 [HPW 9 O 34 ]<br />

(aq). A solution <strong>of</strong> Na 8 [HPW 9 O 34 ] (aq) (10.65 g , 3.75 mmol) was added to K 7 PW 11 O 39<br />

(aq) (4.0 g 1.25 mmol), the pH reduced to 6-6.5 <strong>and</strong> the mixture was stirred <strong>and</strong> heated<br />

to 50 C. Solid KCl was added until a fine crystalline precipitate appeared . Further<br />

solid KCl about 3-4 g was then added. A white crystalline powder separated. Stirring<br />

was maintained until room temperature was reached. The product was then filtered <strong>of</strong>f<br />

<strong>and</strong> washed with chilled water (10 mL). The solid product was characterized using FTIR<br />

spectroscopy <strong>and</strong> was compared with the published data [80].<br />

24


Part-I<br />

Dimethyltin containing POMs


Chapter 3<br />

Results<br />

3.1 The hybrid organic-inorganic 2-D material<br />

(CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-XW 9 O 33 )]·5H 2 O) ∞<br />

(X = As III , Sb III ) <strong>and</strong> its solution properties<br />

3.1.1 Experimental<br />

The lacunary precursors Na 9 [α-AsW 9 O 33 ] <strong>and</strong> Na 9 [α-SbW 9 O 33 ] were synthesized according<br />

to published procedures <strong>and</strong> their purity was confirmed by infrared spectroscopy<br />

[69, 70]. All other reagents were used as purchased without further purification.<br />

Syntheses<br />

0.58g (2.64 mmols) Sn(CH 3 ) 2 Cl 2 was dissolved in 40 mL distilled H 2 O followed by addition<br />

<strong>of</strong> 2.00 g (0.80 mmol) Na 9 [α-AsW 9 O 33 ] This solution at pH 3.0 was heated to ∼80 ◦ C for 1<br />

h <strong>and</strong> then cooled to room temperature <strong>and</strong> filtered. A few drops <strong>of</strong> 0.1 M CsCl were added<br />

<strong>and</strong> then the solution was allowed to evaporate in an open beaker at room temperature.<br />

After 1-2 days a white crystalline product started to appear. Evaporation was allowed to<br />

continue until the solvent level had approached the solid product, which was filtered <strong>of</strong>f<br />

<strong>and</strong> air-dried. A total <strong>of</strong> 1.9 g (yield 76%) <strong>of</strong> crystalline product was obtained. Elemental<br />

analysis calculated. (found) for (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-AsW 9 O 33 )]·5H 2 O) ∞<br />

(CsNa-1)(MW = 3107.0): Cs 4.3 (4.0), Na 3.0 (2.8), Sn 11.5 (10.9), As 2.4 (2.1), W 53.3<br />

26


(54.0)% FTIR spectra for (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-AsW 9 O 33 )]·5H 2 O) ∞ (CsNa-<br />

1)(KBr disk): 954, 909, 870, 829, 760, 727, 688, 518, 476, 430 cm −1<br />

(CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-SbW 9 O 33 )]·5H 2 O) ∞ (CsNa-2) The synthesis was iden-<br />

Fig. 3.1: FTIR spectra <strong>of</strong> compound CsNa-1(red) <strong>and</strong> Na 9 [α-AsW 9 O 33 ] (blue)<br />

tical to that <strong>of</strong> CsNa-1, with the exception that 2.0 g (0.80 mmol) Na 9 [α-SbW 9 O 33 ] was<br />

used instead <strong>of</strong> Na 9 [α-AsW 9 O 33 ]. In this case a total <strong>of</strong> 2.0 g (yield 79%) <strong>of</strong> crystalline<br />

product was obtained. Elemental analysis calcd. (found) for (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-<br />

SbW 9 O 33 )]·5H 2 O) ∞ (MW = 3153.8): Cs 4.2 (3.9), Na 2.9 (2.8), Sn 11.3 (11.2), Sb 3.9<br />

(3.8), W 52.5 (53.4)%. FTIR spectra for (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 ( β-SbW 9 O 33 )]·5H 2 O) ∞<br />

(KBr disk): 951, 865, 820, 747, 677, 520, 473, 457, 422 cm −1 . All elemental analyses were<br />

performed by Kanti Labs Ltd. in Mississauga, Canada. The FTIR spectra were recorded<br />

on a Nicolet Avatar FTIR spectrophotometer in a KBr pellet.<br />

X-ray Crystallography<br />

Crystal data <strong>and</strong> structure refinement details for (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-AsW 9 O 33 )]<br />

·5H 2 O) ∞ CsNa-1 <strong>and</strong> (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-SbW 9 O 33 )]·5H 2 O) ∞ CsNa-2 are<br />

summarized in Table 3.1. The respective crystals were mounted on a glass fiber for indexing<br />

<strong>and</strong> intensity data collection at 173 K on a Bruker D8 SMART APEX CCD single-<br />

27


Fig. 3.2: FTIR spectra <strong>of</strong> compound CsNa-2(red) <strong>and</strong> Na 9 [α-SbW 9 O 33 ](blue)<br />

crystal diffractometer using Mo K α radiation (λ = 0.71073 Å). Direct methods were used<br />

to solve the structure <strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining<br />

atoms were found from successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong><br />

polarization corrections were applied <strong>and</strong> an absorption correction was performed using<br />

the SADABS program.[81] Crystallographic data are summarized in Table 3.1.<br />

Solution NMR<br />

Solution NMR appears to be the most elegant technique to prove or disprove the existence<br />

<strong>of</strong> 1 <strong>and</strong> 2 in solution. It is well known that NMR <strong>of</strong> the addenda atoms<br />

(in this case tungsten) is the most sensitive analytical tool to obtain structural information<br />

<strong>of</strong> polyoxoanions in solution. The polymeric compounds CsNa-1 <strong>and</strong> CsNa-<br />

2 are ideal for a multinuclear NMR study, because they contain several NMR-active,<br />

spin 1/2 nuclei <strong>and</strong> they are diamagnetic. Therefore we performed 183 W-, 119 Sn-, 1 H-<br />

<strong>and</strong> 13 C-NMR studies on (a) freshly synthesized solutions <strong>of</strong> CsNa-1 <strong>and</strong> CsNa-2 (see<br />

Experimental Section) <strong>and</strong> (b) solid CsNa-1 <strong>and</strong> CsNa-2 redissolved in water. Interestingly<br />

we obtained exactly the same results in both cases. NMR (D 2 O, 293 K) for<br />

28


Table 3.1: Crystal Data <strong>and</strong> Structure Refinement for compounds CsNa-1 <strong>and</strong> CsNa-2<br />

formula AsC 6 CsH 36 Na 4 O 43 Sn 3 W 9 (1) C 6 CsH 36 Na 4 O 43 SbSn 3 W 9 (2)<br />

fw (g/mol) 3107 3153.8<br />

crystal color colorless colorless<br />

crystal system orthorhombic orthorhombic<br />

crystal size (mm 3 ) 0.10 × 0.06 × 0.02 0.11 × 0.06 × 0.04<br />

space group (No.) P na2 1 (33) P na2 1 (33)<br />

unit cell dim.<br />

a (Å) 26.118(2) 26.118(2)<br />

b (Å) 16.064(1) 16.064(1)<br />

c (Å) 13.776(1) 13.776(1)<br />

vol (Å 3 ) 5779.4(7) 5779.4(7)<br />

Z 1 1<br />

dcalc (Mg m −3 ) 3.561 3.672<br />

abs. coeff. (mm −1 ) 20.548 20.595<br />

Reflections (unique) 14373 14370<br />

Reflections (obs.) 11413 12214<br />

a<br />

R (F o) 0.068 0.052<br />

b<br />

R w(F o) 0.1401 0.1078<br />

diff. peak (eÅ 3 ) 5.248 2.985<br />

diff. hole (eÅ 3 ) -5.107 -3.762<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

(CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-AsW 9 O 33 )]·5H 2 O) 183 ∞ W: (relative intensities in parenthesis)<br />

-130.5(2), -135.2(1), -138.1(2), -142.5(2), -146.2(2) ppm; 119 Sn: -175.9(1), -200.0(2)<br />

ppm; 13 C: 8.5 ppm; 1 H: 0.7 ppm. NMR (D 2 O, 293 K) for (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-<br />

SbW 9 O 33 )]·5H 2 O) : 183 W: (relative intensities in parenthesis) -108.1(2), -117.4(1), -120.1(2),<br />

-127.1(2), -134.7(2) ppm; 119 Sn: -164.6(1), -205.2(2) ppm; 13 C: 8.7 ppm; 1 H: 0.7 ppm. All<br />

NMR spectra were recorded with freshly synthesized, concentrated solutions <strong>of</strong> CsNa-1<br />

<strong>and</strong> CsNa-2 on a JEOL Eclipse 400 instrument. The 183 W-NMR measurements were performed<br />

at 16.656 MHz in 10 mm tubes <strong>and</strong> the 119 Sn, 13 C, <strong>and</strong> 1 H spectra were recorded<br />

in 5 mm tubes at 149.081, 100.525 <strong>and</strong> 399.782 MHz, respectively. The chemical shifts<br />

for unbound dimethyltin dichloride at pH 3 are at -243.7 ppm ( 119 Sn), 12.2 ppm ( 13 C)<br />

<strong>and</strong> 0.84 ppm ( 1 H), respectively.<br />

29


Fig. 3.3: W 183 NMR spectra <strong>of</strong> (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 ( β-XW 9 O 33 )]·5H 2 O) ([X = As (top)CsNa-<br />

1, Sb (bottom)CsNa-2]<br />

3.1.2 Results <strong>and</strong> discussion<br />

The novel <strong>and</strong> isostructural polyoxoanion-based 2-D materials (CsNa 4 [Sn(CH 3 ) 2 } 3 O(H 2 O) 4<br />

(β-XW 9 O 33 )]·5H 2 O) (X = As, CsNa-1; Sb, CsNa-2), see Figures 3.5. The structure <strong>of</strong><br />

CsNa-1 <strong>and</strong> CsNa-2 is best described as a polymeric network composed <strong>of</strong> monopolyanionic<br />

building blocks ({Sn(CH 3 ) 2 (H 2 O) 2 } 3 (β-XW 9 O 33 )) (X = As, Sb) that are linked via<br />

Sn-O-(W’) bridges. This leads to a 2-D surface which is not planar, but could perhaps be<br />

described as a shelf with a zig-zag backbone where the individual levels are decorated by<br />

methyl groups. Clearly, the solid state structure is governed by the organic functionalities.<br />

Compounds CsNa-1 <strong>and</strong> CsNa-2 are isostructural, which means that the different sizes<br />

<strong>of</strong> the As <strong>and</strong> Sb heteroatoms with their associated lone pairs have no significant effect<br />

on the solid state structures. The three organo-tin groups attached to each monomeric<br />

unit <strong>of</strong> CsNa-1 <strong>and</strong> CsNa-2 contain tin centers that are octahedrally coordinated by<br />

two oxo groups, two methyl groups <strong>and</strong> two water molecules, respectively. Furthermore<br />

the two methyl groups on each tin atom are positioned trans to each other. However,<br />

only two organotin groups are structurally equivalent <strong>and</strong> different from the third. The<br />

molecular formulae <strong>and</strong> charges <strong>of</strong> CsNa-1 <strong>and</strong> CsNa-2 are supported by bond valence<br />

30


Fig. 3.4: Sn 119 NMR spectra <strong>of</strong> (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 ( β-XW 9 O 33 )]·5H 2 O) [X = As (top)CsNa-<br />

1, Sb(bottom)CsNa-2]<br />

calculations, which indicate that all terminal oxygen atoms bound to the tin atoms are<br />

diprotonated.[82] Close inspection <strong>of</strong> CsNa-1 <strong>and</strong> CsNa-2 indicates that the C-Sn-C<br />

bond angles are significantly smaller than ∼180 ◦ C indicating that the interior methyl<br />

groups experience still some degree <strong>of</strong> repulsion (see Figure 3.6). The C-Sn-C bond angles<br />

<strong>of</strong> the two equivalent organotin groups are very similar for 1a (149(1), 152(1)) <strong>and</strong><br />

CsNa-2 (150(1), 151(1) ◦ ). However, the C-Sn-C bond angle <strong>of</strong> the unique organotin<br />

unit is distinctly larger in polyanion CsNa-1 (166(1) ◦ ) than in CsNa-2 (162(1) ◦ ). This<br />

seems to reflect (a) the smaller size <strong>of</strong> the As III atom compared to Sb III , (b) the shorter<br />

As III -O bond lengths compared to Sb III <strong>and</strong> (c) the smaller size <strong>of</strong> the As III lone pair<br />

compared to Sb III . It is known that the alpha to beta isomerization <strong>of</strong> [α-AsW 9 O 33 ] 9−<br />

<strong>and</strong> [α-SbW 9 O 33 ] 9− is facilitated in acidic, aqueous medium [69, 83]. This is in complete<br />

agreement with the synthetic conditions for CsNa-1 <strong>and</strong> CsNa-2, which were isolated<br />

at pH 3. The (β-AsW 9 O 33 ) fragment has so far only been observed in two different polyoxoanion<br />

structures [33, 83]. The large tungstoarsenate(III) [As 6 W 65 O 217 (H 2 O) 7 ] 26− is the<br />

only example <strong>of</strong> a polyanion that contains both isomers, (α-AsW 9 O 33 ) <strong>and</strong> (β-AsW 9 O 33 ),<br />

31


Fig. 3.5: Left: combined polyhedral <strong>and</strong> ball/stick representation <strong>of</strong> the 2-D solid state structure <strong>of</strong><br />

(CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-XW 9 O 33 )]·5H 2 O) ∞ (X = As, CsNa-1; Sb, CsNa-2). The WO 6 octahedra<br />

are purple <strong>and</strong> the balls represent tin (green), arsenic/antimony (blue), oxygen (red) <strong>and</strong> carbon<br />

(yellow). Hydrogen atoms are omitted for clarity (left),Right: Side view <strong>of</strong> the 2-D solid state structure<br />

<strong>of</strong> (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-XW 9 O 33 )]·5H 2 O) ∞ (X = As, CsNa-1; Sb, CsNa-2).<br />

in the same structure [33]. The fact that no cations are involved in the exclusively covalent<br />

2-D network <strong>of</strong> CsNa-1 <strong>and</strong> CsNa-2 requires it to possess a negative charge. Indeed the<br />

appropriate number <strong>of</strong> cations (1 Cs, 4 Na) for each polyanionic unit was found by X-ray<br />

diffraction <strong>and</strong> elemental analysis. Therefore CsNa-1 <strong>and</strong> CsNa-2 are best described as<br />

polymeric polyanion salts. This view is further supported by our observation that CsNa-1<br />

<strong>and</strong> CsNa-2 are water-soluble upon heating. Clearly, in solution the polymeric structure<br />

<strong>of</strong> CsNa-1 <strong>and</strong> CsNa-2 has to decompose <strong>and</strong> we identified the likely sites: the very<br />

long Sn-O(W’) bonds in CsNa-1 (2.414 - 2.656 Å) <strong>and</strong> CsNa-2 (2.427 - 2.633 Å). The<br />

question arises in which form CsNa-1 <strong>and</strong> CsNa-2 are present in solution: oligomeric<br />

or monomeric. The latter option requires the existence <strong>of</strong> the hypothetical, monomeric<br />

polyanion [{Sn(CH 3 ) 2 (H 2 O) 2 } 3 (β-XW 9 O 33 )] 3− (X = As III 1, Sb III 2). Polyanion 1 consists<br />

<strong>of</strong> a (β-AsW 9 O 33 ) fragment which is stabilized by three dimethyltin fragments <strong>and</strong><br />

polyanion 2 represents its antimony derivative (see Figures 3.6).<br />

The three dimethyltin groups <strong>of</strong> 1 <strong>and</strong> 2 are grafted onto the polyanion via two Sn-<br />

32


Fig. 3.6: Left:Ball <strong>and</strong> stick representation <strong>of</strong> [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-XW 9 O 33 )] 3− (X = As,1; Sb,2.(A)<br />

The balls represent tungsten (black), tin (green), arsenic/antimony (blue), oxygen (red), carbon (yellow)<br />

<strong>and</strong> hydrogen (small black); Right:Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong> the<br />

[{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-XW 9 O 33 )] 3− (X = As, 1; Sb, 2).(B) The WO 6 octahedra are red <strong>and</strong> the color<br />

codes <strong>of</strong> balls are same as above<br />

O(W) bonds each on the side <strong>of</strong> the hetero atom lone pair. Tungsten-183 NMR resulted<br />

in five peaks (intensity ratios 2:1:2:2:2) for 1a at -130.5, -135.2, -138.1, -142.5 <strong>and</strong> -146.2<br />

ppm <strong>and</strong> for 1b at -108.1, -117.4, -120.1, -127.1 <strong>and</strong> -134.7 ppm (see Figure 3.3). The<br />

latter spectrum shows an additional, very small peak at -124.6 ppm for which we do not<br />

yet have a good explanation. We also performed 119 Sn-NMR <strong>and</strong> observed two peaks with<br />

an intensity ratio <strong>of</strong> 1:2 at -175.9, -200.0 ppm CsNa-1 <strong>and</strong> -164.6, -205.2 ppm CsNa-2<br />

(see Figure 3.4). Interestingly both signals for CsNa-1 are somewhat broader than those<br />

for CsNa-2. The 13 C-NMR measurements resulted in one peak for CsNa-1 at 8.5 ppm<br />

<strong>and</strong> for CsNa-2 at 8.7 ppm. As expected 1 H-NMR resulted also in one peak at 0.7<br />

ppm for CsNa-1 <strong>and</strong> CsNa-2. All <strong>of</strong> these results indicate that in solution a species<br />

with C s symmetry is present. This pro<strong>of</strong>s unequivocally the existence <strong>of</strong> the discrete,<br />

monomeric polyanions 1 <strong>and</strong> 2 in solution. Furthermore, the NMR spectra for CsNa-1<br />

<strong>and</strong> CsNa-2 do not change even after several months. Apparently the dimethyltin groups<br />

are tightly bound to the tungsten-oxo fragment. The monomeric polyoxoanions 1 <strong>and</strong> 2<br />

are synthesized by heating an aqueous solution containing dimethyltin dichloride <strong>and</strong><br />

Na 9 [α-XW 9 O 33 ] (X = As III , Sb III ) in the appropriate stoichiometric ratio (3:1) in aqueous,<br />

acidic medium. From this solution the polymeric materials CsNa-1 <strong>and</strong> CsNa-2 are<br />

33


formed upon crystallization <strong>and</strong> both products can be isolated in good yield. Interestingly<br />

formation <strong>of</strong> 1 <strong>and</strong> 2 is accompanied by the isomerization (α-XW 9 O 33 ) → (β-XW 9 O 33 )(X<br />

= As, Sb). Most likely steric interactions <strong>of</strong> the methyl groups do not allow formation <strong>of</strong><br />

the hypothetical trisubstituted alpha derivative [{Sn(CH 3 ) 2 (H 2 O) 2 } 3 (α-XW 9 O 33 )] 3− (X=<br />

As III , Sb III ). More precisely, the three ‘internal’ methyl groups <strong>of</strong> three different tin centers<br />

would come very close to each other if they were to be grafted on a (α-AsW 9 O 33 ) or<br />

(α-SbW 9 O 33 ) fragment. Furthermore the lone pair <strong>of</strong> electrons on the heteroatom may<br />

exhibit a repulsive effect on the internal methyl groups. Rotation <strong>of</strong> one W 3 O 13 by ∼60<br />

◦ C allows results in the beta-isomer <strong>and</strong> now all three dimethyltin units can be bound<br />

resulting in a stable structure. Almost certainly the terminal oxygens <strong>of</strong> all tin atoms are<br />

diprotonated. The presence <strong>of</strong> labile water lig<strong>and</strong>s explains the tendency <strong>of</strong> 1 <strong>and</strong> 2 to<br />

polymerize in the solid state.<br />

3.1.3 Conclusions<br />

The diorganotin fragments can be incorporated in polyoxotungstates. Polyanions 1 <strong>and</strong><br />

2 represent (a) novel examples <strong>of</strong> hybrid organic-inorganic polyoxoanions, (b) the first<br />

examples <strong>of</strong> diorganotin-substituted polyoxoanions, (c) the first monomeric organotin<br />

derivatives <strong>of</strong> lone pair containing polyoxoanions <strong>and</strong> (d) the first organotin derivatives <strong>of</strong><br />

[β-AsW 9 O 33 ] 9− <strong>and</strong> [β-SbW 9 O 33 ] 9− . The compounds presented here are rare examples <strong>of</strong><br />

discrete polyoxoanions which polymerize upon crystallization leading to a 2-D structure<br />

with inorganic <strong>and</strong> organic surface regions.<br />

34


3.2 The tetrakis-dimethyltin containing<br />

tungstophosphate ({Sn(CH 3 ) 2 } 4 {H 2 P 4 W 24 O 92 } 2 ) 28−<br />

evidence for lacunary Preyssler ion<br />

3.2.1 Experimental<br />

Synthesis<br />

K 17 Li 11 [{Sn(CH 3 ) 2 } 4 (H 2 P 4 W 24 O 92 ) 2 ]·51H 2 O (KLi-3). Polyanion 3 was synthesized by<br />

interaction <strong>of</strong> 0.072 g (0.30 mmol) (CH 3 ) 2 SnCl 2 with 0.728 g (0.10 mmol) K 16 Li 2 [H 6 P 4 W 24 O 94 ]<br />

in 20 mL LiOAc buffer at pH 4.1. The solution was heated to ∼50 ◦ C for 1h <strong>and</strong> filtered<br />

after it had cooled. Addition <strong>of</strong> 0.5 mL <strong>of</strong> 1.0 M KCl solution to the colorless filtrate <strong>and</strong><br />

slow evaporation at room temperature led to a white, crystalline product after about two<br />

weeks (yield 0.39 g, 54 %). Anal. calcd. (found) for 1a: K 4.7 (4.8), Li 0.5 (0.8), W 61.8<br />

(61.1), P 1.7 (1.5), Sn 3.3 (3.5), C 0.7 (0.8), H 0.9 (1.0). FTIR for compound KLi-3:<br />

1136, 1086, 1018, 981, 952, 928, 915, 812, 693, 573, 525, 462 cm −1 .<br />

Fig. 3.7: FTIR spectra <strong>of</strong> compound KLi-3(red) <strong>and</strong> K 16 Li 2 [H 6 P 4 W 24 O 94 ] (blue)<br />

Elemental analysis was performed by Kanti Labs Ltd. in Mississauga, Canada. The<br />

FTIR spectrum was recorded on a Nicolet Avatar FTIR spectrophotometer in a KBr<br />

pellet. All NMR spectra were recorded on a JEOL Eclipse 400 instrument at room<br />

temperature using D 2 O as a solvent.<br />

35


X-ray Crystallography<br />

A crystal <strong>of</strong> compound 2a was mounted on a glass fiber for indexing <strong>and</strong> intensity data<br />

collection at 163 K on a Bruker D8 SMART APEX CCD single-crystal diffractometer<br />

using Mo K α radiation [λ = 0.71073 Å]. Direct methods were used to solve the structure<br />

<strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining atoms were found from<br />

successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong> polarization corrections<br />

were applied <strong>and</strong> an absorption correction was performed using the SADABS program<br />

[81]. Crystallographic data are summarized in Table 3.2.<br />

Table 3.2: Crystal Data <strong>and</strong> Structure Refinement for compound KLi-3<br />

Emperical formula C 8 H 130 K 17 Li 11 O 235 P 8 Sn 4 W 48<br />

fw 14275.8<br />

space group (No.) P 4 2 /nmc (14)<br />

a (Å) 21.5112(17)<br />

b (Å) 21.5112(17)<br />

c (Å) 27.171(3)<br />

vol (Å 3 ) 12573(2)<br />

Z 2<br />

temp ( ◦ C) -120<br />

wavelength (Å) 0.71073<br />

d calcd (mg m −3 ) 3.72<br />

abs coeff. (mm −1 ) 22.69<br />

R [I > 2 σ(I)] a 0.045<br />

R w (all data) b 0.109<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

Electrochemistry : General Methods <strong>and</strong> Materials<br />

Pure water was used throughout. It was obtained by passing through a RiOs 8 unit<br />

followed by a Millipore-Q Academic purification set. All reagents were <strong>of</strong> high-purity grade<br />

<strong>and</strong> were used as purchased without further purification. The pH = 4 medium was made<br />

<strong>of</strong> 1 M CH 3 COOLi + CH 3 COOH. Electrochemical Experiments. The concentration <strong>of</strong><br />

polyanion 3 was 2 × 10 −4 M. The solutions were deaerated thoroughly for at least 30 min.<br />

with pure argon <strong>and</strong> kept under a positive pressure <strong>of</strong> this gas during the experiments.<br />

The source, mounting <strong>and</strong> polishing <strong>of</strong> the glassy carbon (GC, Tokai, Japan) electrodes<br />

36


has been described [84]. The glassy carbon samples had a diameter <strong>of</strong> 3 mm. The<br />

electrochemical set-up was an EG & G 273 A driven by a PC with the M270 s<strong>of</strong>tware.<br />

Potentials are quoted against a saturated calomel electrode (SCE). The counter electrode<br />

was a platinum gauze <strong>of</strong> large surface area. All experiments were performed at room<br />

temperature.<br />

3.2.2 Results <strong>and</strong> discussion<br />

Synthesis <strong>and</strong> Structure<br />

Reaction <strong>of</strong> (CH 3 ) 2 SnCl 2 with K 16 Li 2 [H 6 P 4 W 24 O 94 ] in aqueous, acidic(pH 4.1) medium<br />

at ∼50 ◦ C resulted in[{Sn(CH 3 ) 2 } 4 (H 2 P 4 W 24 O 92 ) 2 ] 28− 3. Single crystal X-ray analysis on<br />

K 17 Li 11 [{Sn(CH 3 ) 2 } 4 (H 2 P 4 W 24 O 92 ) 2 ]·51H 2 O KLi-3 revealed that the novel polyanion 3<br />

is composed <strong>of</strong> two (P 4 W 24 O 92 ) fragments that are linked by four equivalent diorganotin<br />

groups. The unprecedented assembly 3 has D 2d symmetry <strong>and</strong> contains an empty, hydrophobic<br />

pocket (diameter around 2 Å) in the center <strong>of</strong> the molecule (see Figure 3.8).<br />

The two (P 4 W 24 O 92 ) fragments <strong>of</strong> 3 are orthogonal to each other <strong>and</strong> are held together<br />

by four structurally equivalent dimethyltin groups. Interestingly, the two (P 2 W 12 O 48 )<br />

Fig. 3.8: Ball <strong>and</strong> stick (left), Polyhedron (right) representation <strong>of</strong> polyanion 3<br />

subunits <strong>of</strong> each (P 4 W 24 O 92 ) fragment are fused via four W-O-W’ bridges involving cap<br />

<strong>and</strong> belt tungsten centers. It remains to be seen if the ‘free’ [H 6 P 4 W 24 O 94 ] 18− precursor<br />

has the same connectivity, or only two W-O-W’ bridges involving the caps as originally<br />

suggested by Contant <strong>and</strong> Tézé [76]. We are currently trying to obtain single crystals<br />

<strong>of</strong> K 16 Li 2 [H 6 P 4 W 24 O 94 ] in order to resolve this open question. It is highly likely that<br />

37


the presence <strong>of</strong> the four (CH 3 ) 2 Sn 2+ groups causes the small angle ∼45 ◦ <strong>of</strong> the fused<br />

(P 2 W 12 O 48 ) groups. In agreement with our previously reported diorganotin-containing<br />

polyoxotungstates, the two methyl groups <strong>of</strong> each tin atom in 2 are in relative trans<br />

positions (C1-Sn-C2 =∼169.6(11) ◦ , see Figure 3.9. The Sn-C bond lengths in 2 are<br />

Fig. 3.9: Ball <strong>and</strong> stick representation <strong>of</strong> the asymmetric unit <strong>of</strong> polyanion 3<br />

2.12(3) <strong>and</strong> 2.19(3) Å, respectively, <strong>and</strong> the equatorial Sn-O bond lengths are 2.230(13)<br />

<strong>and</strong> 2.254(12) Å, respectively. The bond lengths <strong>of</strong> the tungstophosphate framework are<br />

not unusual. Bond valence sum calculations indicate that polyanion 3 contains only one<br />

type <strong>of</strong> protonated oxygen (O1A, s = 1.44) [85]. There is a total <strong>of</strong> eight O1A atoms in<br />

3 <strong>and</strong> they are all 2-oxo bridges (W1-O1A-Sn1) linking the four cap-tungstens <strong>of</strong> each<br />

(P 2 W 12 O 48 ) fragment with tin centers. The intermediate bond valence sum for O1A indicates<br />

that only 50% <strong>of</strong> all eight atoms are actually protonated. This probably means<br />

that at each cap <strong>of</strong> the two (P 2 W 12 O 48 ) units in 3 only one <strong>of</strong> the two W-O-Sn bridges<br />

is actually protonated. Interestingly, this is in complete agreement with the conclusions<br />

<strong>of</strong> Contant <strong>and</strong> Tézé in their studies <strong>of</strong> the [H 2 P 2 W 12 O 48 ] 12− <strong>and</strong> [H 2 P 4 W 24 O 94 ] 22−<br />

precursors.[76] All potassium ions could be identified crystallographically (K4 with half<br />

occupancy), but none <strong>of</strong> the lithium ions. Nevertheless, their presence <strong>and</strong> therefore the<br />

complete molecular formula <strong>of</strong> KLi-3 was determined by elemental analysis.<br />

38


Solution NMR<br />

Polyanion 3 is diamagnetic <strong>and</strong> contains four spin 1 nuclei 2<br />

(183 W, 119 Sn, 13 C, 1 H) <strong>and</strong><br />

therefore represents a good c<strong>and</strong>idate for solution NMR studies at room temperature.<br />

We also examined the solution properties <strong>of</strong> 3 by multinuclear NMR (D 2 O, ∼20 ◦ ). Our<br />

31 P NMR measurements resulted in two singlets (-7.1, -8.7 ppm) <strong>of</strong> equal intensity <strong>and</strong><br />

1 H NMR showed a singlet at 0.7 ppm. For 119 Sn <strong>and</strong> 13 C NMR (both 1 H decoupled)<br />

we observed singlets at -243.2 ppm <strong>and</strong> 8.6 ppm, respectively. Although we used the<br />

lithium salt <strong>of</strong> 3 we were unable to observe any signals in 183 W NMR, which indicates<br />

that the concentration <strong>of</strong> the polyanion was too small due to very poor solubility. Close<br />

inspection <strong>of</strong> the structure <strong>of</strong> each (P 4 W 24 O 92 ) fragment in 3 indicates that it may be<br />

considered as a lacunary derivative <strong>of</strong> the well-known Preyssler ion [NaP 5 W 30 O 110 ] 14−<br />

[86]. This plenary polyanion has approximate D 5h symmetry <strong>and</strong> consists <strong>of</strong> a cyclic<br />

assembly <strong>of</strong> five (PW 6 O 22 ) units. The (P 4 W 24 O 92 ) fragments in 3 may be considered as<br />

being composed <strong>of</strong> a cyclic assembly <strong>of</strong> four (PW 6 O 22 ) units with a linkage pattern <strong>of</strong> the<br />

individual units identical to the Preyssler ion. However, due to one missing (PW 6 O 22 )<br />

unit in each (P 4 W 24 O 92 ) fragment the latter is best described as a lacunary Preyssler ion.<br />

Electrochemistry<br />

The polyanion 3 was also studied by cyclic voltammetry in 1 M (CH 3 COOLi + CH 3 COOH)<br />

pH 4 buffer corresponding essentially to its synthesis medium. Figure 3.10A shows the<br />

features <strong>of</strong> interest in the overall CV <strong>of</strong> 3 in the above pH 4 medium. This CV is constituted<br />

by a broad first wave, followed by a large current intensity second wave which<br />

peaks close to the electrolyte discharge. Actually, the second multi-electron reduction<br />

process is a combination <strong>of</strong> two very closely-spaced waves. The decomposition products<br />

generated at this combined wave deposit on the electrode surface upon continuous cycling<br />

<strong>of</strong> the potential up to the cathodic limit shown in Figure 3.10A. The efficiency <strong>of</strong> the<br />

deposition process increases when the potential scan rate decreases. Specifically, a very<br />

faint reversibility can be detected for this combined second wave at a scan rate <strong>of</strong> 200 mV<br />

s −1 . Such reductive deposition phenomena at fairly negative potentials are known <strong>and</strong><br />

have been described for a large variety <strong>of</strong> heteropolyanions [87]. The complex phenom-<br />

39


Fig. 3.10: Cyclic voltammogram <strong>of</strong> 2 × 10 −4 M solution <strong>of</strong> 3 in a pH 4 medium (1 M CH 3 COOLi +<br />

CH 3 COOH). The scan rate was 10 mV.s −1 , the working electrode was glassy carbon <strong>and</strong> the reference<br />

electrode was SCE. (A) The whole voltammetric pattern(left), (B) The voltammetric pattern restricted<br />

to the first redox processes (right)<br />

ena associated with the second combined wave will not be considered further here. The<br />

behavior <strong>of</strong> the first wave is more characteristic <strong>of</strong> 3 <strong>and</strong> is shown in Figure 3.10B. In the<br />

following, the CV pattern is restricted to this process. The cathodic part features a broad<br />

electrochemical wave which remains composite if the potential scan rate is lower than<br />

the 200 mV s −1 explored here. In contrast, two well-separated oxidation processes are<br />

observed on potential reversal, at - 0.612 V <strong>and</strong> - 0.356 V vs SCE, respectively, thus confirming<br />

the composite nature <strong>of</strong> the cathodic scan. The detailed molecular structure <strong>of</strong> the<br />

polyanion precursor [H 2 P 4 W 24 O 94 ] 22− before its engagement in the self-assembly process<br />

resulting in formation <strong>of</strong> 3 remains unknown. Nevertheless, comparison <strong>of</strong> the first wave<br />

system <strong>of</strong> 3 with the cyclic voltammograms <strong>of</strong> [H 2 P 4 W 24 O 94 ] 22− <strong>and</strong> the wheel-shaped<br />

[H 7 P 8 W 48 O 184 ] 33− in the same medium is expected to be useful. This study might allow<br />

us to extract some distinct features <strong>of</strong> 3 which are due to the incorporated dimethyltin<br />

units. All three polyanions can be considered as lacunary species <strong>and</strong> thus, their electrochemistry<br />

should be carried out in a well-buffered electrolyte [88]. The comparisons<br />

between 3 <strong>and</strong> [H 2 P 4 W 24 O 94 ] 22− on the one h<strong>and</strong> <strong>and</strong> between 3 <strong>and</strong> [H 7 P 8 W 48 O 184 ] 33−<br />

on the other h<strong>and</strong>, are shown in Figures 3.11, respectively. The potential locations <strong>of</strong><br />

the reduction peaks <strong>and</strong> the number <strong>of</strong> electrons consumed by each wave are gathered in<br />

Table 3.3.<br />

The number <strong>of</strong> electrons for the waves <strong>of</strong> [H 7 P 8 W 48 O 184 ] 33− (three equal 8-electron<br />

40


Fig. 3.11: Comparison <strong>of</strong> the cyclic voltammograms <strong>of</strong> 2 × 10 −4 M solution <strong>of</strong> 3 in a pH 4 medium (1 M<br />

CH 3 COOLi + CH 3 COOH). The scan rate was 10 mV.s −1 , the working electrode was glassy carbon <strong>and</strong><br />

the reference electrode was SCE. (A) Comparison <strong>of</strong> the voltammetric patterns <strong>of</strong> 3 <strong>and</strong> [H 2 P 4 W 24 O 94 ] 22−<br />

(left)(B) Comparison <strong>of</strong> the voltammetric patterns <strong>of</strong> 3 <strong>and</strong> [H 7 P 8 W 48 O 184 ] 33 (right)<br />

Table 3.3: Reduction peak potentials measured from CVs <strong>and</strong> number (n 1 or n 2 ) <strong>of</strong> electrons corresponding<br />

to each wave <strong>of</strong> 3, [H 2 P 4 W 24 O 94 ] 22− <strong>and</strong> [H 7 P 8 W 48 O 184 ] 33− The scan rate was 10 mV s −1 , the<br />

working electrode was glassy carbon <strong>and</strong> the reference electrode was SCE<br />

Polyoxometalate n 1 - E pc1 / V vs SCE n 2 - E pc2 / V vs SCE<br />

3 16 0.755 - -<br />

[H 2 P 4 W 24 O 94 ] 22− 4 0.802 4 ∼1.110<br />

[H 7 P 8 W 48 O 184 ] 33− 8 0.610 8 0.728<br />

waves)[76, 89] <strong>and</strong> [H 2 P 4 W 24 O 94 ] 22− (three equal 4-electron waves)[76] were known from<br />

previous works. For the first wave <strong>of</strong> 3 the corresponding number <strong>of</strong> electrons was evaluated<br />

in two different ways. A first rough estimate was performed by determining the<br />

area delimited by the relevant voltammograms <strong>and</strong> comparing the charges corresponding<br />

to each wave. This method was used to compare 3 <strong>and</strong> [H 2 P 4 W 24 O 94 ] 22− , which shows<br />

suitable, well-separated waves. The charge ratio was 4.07 in favor <strong>of</strong> 3, thus indicating<br />

roughly 16-electrons per molecule for the first wave <strong>of</strong> this complex. As a second method,<br />

controlled potential coulometry was carried out at - 0.780 V vs SCE, but the number<br />

<strong>of</strong> electrons consumed per molecule exceeds 16, due to a non-identified catalytic <strong>and</strong>/or<br />

decomposition process. Such behavior was also observed previously during coulometric<br />

studies on [H 7 P 8 W 48 O 184 ] 33− , albeit to a smaller extent [89]. Finally, it can be concluded<br />

that the result <strong>of</strong> the second method supports the value <strong>of</strong> roughly 16-electrons per molecule<br />

<strong>of</strong> 3 obtained by the first method. Further coulometric study <strong>of</strong> the catalytic process<br />

41


as a function <strong>of</strong> electrolyte composition might help to accurately determine electron numbers,<br />

but is beyond the scope <strong>of</strong> the present work.<br />

Figures 3.11 <strong>and</strong> Table 3.3 indicate unambiguously that the three polyanion species 3,<br />

[H 2 P 4 W 24 O 94 ] 22− <strong>and</strong> [H 7 P 8 W 48 O 184 ] 33− have distinct electrochemical properties. During<br />

these studies, we also observed that the stability <strong>of</strong> [H 2 P 4 W 24 O 94 ] 22− in solution is<br />

enhanced by formation <strong>of</strong> 2. In fact, the combined results from present <strong>and</strong> previous<br />

[76, 89] work allow to establish the following order <strong>of</strong> stability: [H 7 P 8 W 48 O 184 ] 33− >3 ><br />

[H 2 P 4 W 24 O 94 ] 22− . In agreement with literature, the CV pattern <strong>of</strong> [H 2 P 4 W 24 O 94 ] 22− is<br />

constituted by three separated 4-electron waves, two <strong>of</strong> which are shown in Figure 3.11A<br />

[76]. In Figure 3.11A <strong>and</strong> Table 3.3, these waves are located at more negative potentials<br />

than the sole wave considered for 3. In contrast, Figure 3.11B <strong>and</strong> Table 3.3 show<br />

the first two 8-electron waves <strong>of</strong> [H 7 P 8 W 48 O 184 ] 33− (already described previously)[76, 89]<br />

to be located at more positive potentials than that <strong>of</strong> 3. In the present case, these results<br />

might be interpreted based on the assumption that 3 is composed <strong>of</strong> two identical,<br />

mostly non-interacting moieties, a feature that would support the existence <strong>of</strong> a composite<br />

reduction wave. Then, the following reduction trends should be expected for the tungstenoxo-frameworks<br />

<strong>of</strong> the three polyoxometalates discussed here: comparing a substituted<br />

polyoxometalate <strong>and</strong> its lacunary precursor, binding <strong>of</strong> one or more cationic moieties to<br />

the latter will diminish the overall negative charge <strong>of</strong> the product polyanion. As a consequence,<br />

the reduction potential <strong>of</strong> the tungsten-oxo-framework will be driven in the<br />

positive direction compared to that <strong>of</strong> the precursor lacunary complex. One assumption<br />

behind this conclusion is that no specific complicating influences appear in the overall<br />

reduction schemes <strong>of</strong> the two polyanions. In short, the cation-substituted polyanion will<br />

behave much like a “saturated” complex, with a reduction potential intermediate between<br />

those <strong>of</strong> the plenary <strong>and</strong> lacunary parents, 3 a feature well illustrated by the reduction<br />

patterns <strong>of</strong> Keggin <strong>and</strong> Dawson-type tungstostannates [90]. However, this conclusion must<br />

be applied with caution in the general case because such a reasoning would not take into<br />

account any specific modification, e.g. a change in acid-base properties <strong>of</strong> the polyanion<br />

as a result <strong>of</strong> cation incorporation. Detailed parameters <strong>of</strong> the cation like size, electronic<br />

configuration <strong>and</strong> coordination geometry including possible distortions (e.g. Jahn-Teller)<br />

42


may also contribute to the specificity [91]. Interestingly, [H 7 P 8 W 48 O 184 ] 33− appears to<br />

behave as if it were more saturated than 3. However, the comparison is not straightforward<br />

as the linkage pattern <strong>of</strong> the (P 2 W 12 O 48 ) subunits is different in [H 7 P 8 W 48 O 184 ] 33−<br />

<strong>and</strong> 3 <strong>and</strong> for free [H 2 P 4 W 24 O 94 ] 22− it is not yet known. As a consequence, the acidities<br />

<strong>of</strong> [H 7 P 8 W 48 O 184 ] 33− ,[76] [H 2 P 4 W 24 O 94 ] 22− <strong>and</strong> 3 might also be different from each other.<br />

Such an assumption is supported by the splitting <strong>of</strong> the CV pattern <strong>of</strong> [H 7 P 8 W 48 O 184 ] 33−<br />

in two 8-electron waves at pH 4, without any possibility to make them merge until pH =<br />

7.3. Even then the merging is incomplete, as a slight decrease <strong>of</strong> the potential scan rate<br />

favors splitting <strong>of</strong> the waves. Finally, it is worth noting that for 3 no wave was observed<br />

which could be associated with reduction <strong>of</strong> the Sn-centers. This result is reminiscent<br />

<strong>of</strong> that obtained by Chorghade <strong>and</strong> Pope for Keggin <strong>and</strong> Dawson-type tungstostannates<br />

[90].<br />

Electrochemistry studies was done by our collaborators Dr. B. Keita, <strong>and</strong> Pr<strong>of</strong>. L. Nadjo,<br />

Université Paris-Sud, Orsay Cedex, France.<br />

3.2.3 Conclusions<br />

The title polyanion 3 is <strong>of</strong> interest for several reasons, (i) it represents the first example<br />

<strong>of</strong> a polyanion synthesized from the [H 6 P 4 W 24 O 94 ] 18− precursor, (ii) it represents the first<br />

example <strong>of</strong> a polyoxoanion containing the (P 4 W 24 O 92 ) fragment, (iii) it reveals that the<br />

(P 4 W 24 O 92 ) fragment is a lacunary derivative <strong>of</strong> the Preyssler ion [NaP 5 W 30 O 110 ] 14− , (iv)<br />

it represents a novel polyanion architecture, (v) it has a central, hydrophobic pocket,<br />

<strong>and</strong> (vi) it adds a new member to our series <strong>of</strong> monomeric, trimeric <strong>and</strong> dodecameric<br />

dimethyl-tin containing polyoxotungstates. The structure <strong>of</strong> 3 allows for a multitude<br />

<strong>of</strong> studies including host/guest chemistry, catalysis <strong>and</strong> medicine. Some <strong>of</strong> this work is<br />

currently in progress <strong>and</strong> the results will be reported in due time. The first reduction wave<br />

in the CV pattern <strong>of</strong> 3 was studied in detail because it contains the main features useful for<br />

a characterization <strong>of</strong> this complex. This 16-electron, broad reduction wave is associated<br />

on potential reversal with two well-separated oxidation processes. As judged from the<br />

potential locations <strong>and</strong> current intensities, the voltammetric pattern is unambiguously<br />

different from those <strong>of</strong> [H 7 P 8 W 48 O 184 ] 33− <strong>and</strong> [H 2 P 4 W 24 O 94 ] 22− , even though no feature<br />

43


was observed that could be traced to the reduction <strong>of</strong> Sn-centers.<br />

3.3 The trimeric-dimethyltin containing bowl shaped<br />

tungstophosphate(V): Cs 12 Na 6 {Sn(CH 3 ) 2 (OH)} 3<br />

(Sn(CH 3 ) 2 ) 3 {A-PW 9 O 34 } 3·14H 2 O<br />

3.3.1 Experimental<br />

The title compound (CsNa-4) was synthesized by dissolving a 1.46 g (0.600 mmol) <strong>of</strong><br />

Na 9 [A-PW 9 O 34 ] in 20 mL sodium acetate buffer followed by addition <strong>of</strong> 0.435 g (1.98<br />

mmol) <strong>of</strong> (CH 3 ) 2 SnCl 2 . This solution (pH 4.8) was heated to ∼80 ◦ C for 1 h <strong>and</strong> then<br />

cooled to room temperature. After filtration, 0.5 mL <strong>of</strong> 0.1 M CsCl solution was added<br />

<strong>and</strong> then the solution was allowed to evaporate in an open vial at room temperature.<br />

A white crystalline product started to appear after a week or two. Evaporation was<br />

continued until the solvent approached the solid product which was suitable for single<br />

crystal XRD, FTIR spectroscopy (see Figure 3.12) <strong>and</strong> elemental analysis.<br />

Fig. 3.12: FTIR spectra <strong>of</strong> compound CsNa-4(red) <strong>and</strong> Na 9 [(A-α-PW 9 O 34 )](blue,)<br />

44


X-ray Crystallography<br />

Single crystal <strong>of</strong> CsNa-4 was mounted on a glass fiber for indexing <strong>and</strong> intensity data<br />

collection at 163 K on a Bruker D8 SMART APEX CCD single-crystal diffractometer<br />

using Mo K α radiation [λ = 0.71073 Å]. Direct methods were used to solve the structure<br />

<strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining atoms were found from<br />

successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong> polarization corrections<br />

were applied <strong>and</strong> an absorption correction was performed using the SADABS program<br />

[81]. Crystallographic data are summarized in Table 3.4<br />

Table 3.4: Crystal Data <strong>and</strong> Structure Refinement for compound CsNa-4<br />

Empirical formula Cs 36 C 36 O 357 P 9 Sn 18 W 81<br />

fw 28236.2<br />

space group R 3m (160)<br />

a (Å) 29.7445(7)<br />

c (Å) 15.5915(7)<br />

vol (Å 3 ) 11946.26(67)<br />

Z 3<br />

temp. ( ◦ C) -100<br />

wavelength (Å) 0.71073<br />

dcalc (Mg m −3 ) 3.912<br />

abs. coeff. (mm −1 ) 23.148<br />

R [I > 2 σ(I)] a 0.053<br />

R w (all data) b 0.062<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

Electrochemistry : General Methods <strong>and</strong> Materials<br />

Pure water was used throughout. It was obtained by passing through a RiOs 8 unit<br />

followed by a Millipore-Q Academic purification set. All reagents were <strong>of</strong> high-purity grade<br />

<strong>and</strong> were used as purchased without further purification. The pH = 4 medium was made<br />

<strong>of</strong> 1 M CH 3 COOLi + CH 3 COOH. Electrochemical Experiments. The concentration <strong>of</strong><br />

polyanion 3 was 2 × 10 −4 M. The solutions were deaerated thoroughly for at least 30 min.<br />

with pure argon <strong>and</strong> kept under a positive pressure <strong>of</strong> this gas during the experiments.<br />

The source, mounting <strong>and</strong> polishing <strong>of</strong> the glassy carbon (GC, Tokai, Japan) electrodes<br />

has been described [84]. The glassy carbon samples had a diameter <strong>of</strong> 3 mm. The<br />

45


electrochemical set-up was an EG & G 273 A driven by a PC with the M270 s<strong>of</strong>tware.<br />

Potentials are quoted against a saturated calomel electrode (SCE). The counter electrode<br />

was a platinum gauze <strong>of</strong> large surface area. All experiments were performed at room<br />

temperature.<br />

Electrochemistry studies was done by our collaborators Dr. B. Keita, <strong>and</strong> Pr<strong>of</strong>. L. Nadjo,<br />

(Université Paris-Sud, Orsay Cedex, France.<br />

3.3.2 Results <strong>and</strong> discussion<br />

Synthesis <strong>and</strong> Structure<br />

The discrete cyclic trimeric title polyanion [{Sn(CH 3 ) 2 (OH)} 3 (Sn(CH 3 ) 2 ) 3 {A-PW 9 O 34 } 3 ] 18−<br />

4 consists <strong>of</strong> three trilacunary A-α-[PW 9 O 34 ] 9− Keggin fragments linked to each other by<br />

three {Sn(CH 3 ) 2 } 2+ bridging groups <strong>and</strong> three {Sn(CH 3 ) 2 } 2+ are linked to each other<br />

forming a trimeric core resulting to a cyclic trimeric bowl type unprecedented structure<br />

with nominal C 3v symmetry. The novel dimethyl tin substituted, trimeric tungsto<br />

Fig. 3.13: Ball/stick (left) <strong>and</strong> Polyhedron (right) representation <strong>of</strong> polyanion 4<br />

phosphate(V) [{Sn(CH 3 ) 2 (OH)} 3 (Sn(CH 3 ) 2 ) 3 {A-PW 9 O 34 } 3 ] 18− 4 consists <strong>of</strong> three lacunary<br />

A-α-[PW 9 O 34 ] 9− Keggin fragments linked via three Sn(CH 3 ) 2+ 2 groups <strong>and</strong> three<br />

Sn(CH 3 ) 2+ 2 connected to each other by three (µ 2 -OH) bridges to form a cyclic core, lead-<br />

46


ing to a structure with nominal C 3v symmetry (see Figure 3.13). Alternatively 4 can<br />

be described as a trilacunary A-α-[PW 9 O 34 ] 9− fragment which has taken up three organotin<br />

units.The discrete cyclic trimeric title polyanion [{Sn(CH 3 ) 2 (OH)} 3 (Sn(CH 3 ) 2 ) 3 {A-<br />

PW 9 O 34 } 3 ] 18− 4 consist <strong>of</strong> three A-α-[PW 9 O 34 ] 9− linked to each other other by three outer<br />

Sn(CH 3 ) 2+ 2 bridging groups <strong>and</strong> three central Sn(CH 3 ) 2+ 2 connected by (µ 2 -OH) bridges<br />

forming a cyclic core which act as a cap to the polyanion resulting in a cyclic trimeric<br />

bowl unprecedented structure with nominal C 3v symmetry.<br />

The two terminal oxygen atoms from adjacent corner-shared WO 6 octahedra <strong>of</strong> each PW 9<br />

Keggin-type unit occupy equatorial positions <strong>of</strong> the tin coordination spheres. The “top”<br />

<strong>of</strong> the structure is capped by three Sn(CH 3 ) 2+ 2 connected to each other by three (µ 2 -<br />

OH) bridges which forms the cyclic fragment. This somewhat resembles to the structure<br />

[(UO 2 ) 3 (H 2 O) 5 As 3 W 29 O 124 ] 19− reported by Pope et al. [92]. However, this capping unit<br />

is analogous to W 3 O 13 reported by Pope et al. we have also observed the capping neutral<br />

fragment {As-AsW 9 } 3 in one <strong>of</strong> our dimethyltin substituted tetrameric species. The<br />

capping by the cyclic fragment <strong>of</strong> the plenary stucture (see Figure 3.13) forms a bowl<br />

shaped assembly. The trans consequence <strong>of</strong> the methyl groups makes the cavity <strong>of</strong> the<br />

bowl hydrophobic.<br />

Each outer dimethyltin group is coordinated through equatorial positions to two terminal<br />

oxygen atom <strong>of</strong> edge-shared WO 6 octahedra belonging to two adjacent Keggin units. The<br />

two methyl groups are trans to each other, which had been reported previously by Kortz<br />

et al. The three dimethyltin groups in the core are coordinated to each other as well as<br />

to two terminal oxygens <strong>of</strong> corner-shared WO 6 octahedra belonging to different (W 3 O 13 )<br />

triad. However, the coordination numbers (6) <strong>of</strong> all the tin centers in 4 are equivalent.<br />

The structural indifference between the outer <strong>and</strong> the core tin atoms is more pronounced<br />

due to significant difference in the C-Sn-C bond angle. Specifically, C-Sn1-C (175.530(2) ◦ )<br />

is significantly higher than the C-Sn2-C (153.670(1) ◦ ), most probably due to the oxygen<br />

bridging the inner dimetyltin core. (See Figure 3.14)<br />

Bond-valence-sum (BVS) calculations for 4 indicated that no oxygen <strong>of</strong> the three<br />

(A-α−PW 9 O 34 ) caps is protonated [85]. However, oxygen <strong>of</strong> the three (µ 2 -OH) bridges,<br />

which forms a cyclic core is protonated.<br />

47


Fig. 3.14: The central core <strong>of</strong> Sn(CH 3 ) 2 2+ connected to each other by three (µ 2 -OH) bridges<br />

Electrochemistry<br />

The polyanion 4 was also studied by cyclic voltammetry in 1 M (CH 3 COOLi + CH 3 COOH)<br />

pH 4 buffer corresponding essentially to its synthesis medium. Among the isomers <strong>of</strong> PW 9 ,<br />

A-α-PW 9 O 34 was used as the lacunary precursor for the synthesis <strong>of</strong> the trimer. X-ray<br />

crystallography indicates that the three such lig<strong>and</strong>s present in this trimer keep their<br />

original structure in the final complex. However, the problem <strong>of</strong> the behaviours <strong>of</strong> the<br />

trimer in solution must be considered. As a matter <strong>of</strong> fact, following the conclusions <strong>of</strong><br />

detailed studies published by Contant <strong>and</strong> Hervé [93], A-α-PW 9 O 34 alone is known to<br />

undergo, in solution, a series <strong>of</strong> transformations that generate α-PW11. Therefore, it was<br />

necessary to check whether the solid state structure <strong>of</strong> the trimer with three A-α-PW 9 O 34<br />

moieties was likely to be retained or not in solution. A pH = 4 medium was selected<br />

for this work. This study was facilitated by the following remark: among the isomers <strong>of</strong><br />

PW 9 fairly stable in solution, A-α-PW 9 O 34 turned out to show a voltammetric pattern<br />

different from those <strong>of</strong> the others <strong>and</strong> which can be considered as its fingerprint.<br />

Such a voltammogram is represented in Figure 3.15A <strong>and</strong> is characterized by a welldefined<br />

four-electron reduction wave, associated, on potential reversal, with a broad oxidation<br />

wave. The CV run for the trimer, in the same potential domain, shows the same<br />

shape as that <strong>of</strong> A-α-PW 9 O 34 . These CVs are clearly distinct from that featuring the<br />

redox behaviours <strong>of</strong> PW 11 which is comprised <strong>of</strong> two two-electron reversible waves [79, 94].<br />

48


Fig. 3.15: Cyclic voltammogram <strong>of</strong> 2 × 10 −4 M solution <strong>of</strong> 4 in a pH 4 medium (1 M CH 3 COOLi +<br />

CH 3 COOH). The scan rate was 10 mV.s −1 , the working electrode was glassy carbon <strong>and</strong> the reference<br />

electrode was SCE. (A) Superposition <strong>of</strong> the CVs restricted to the first redox pattern for trimer <strong>and</strong> A-α-<br />

PW 9 O 34 respectively.(left), (B) Complete CV for the trimer showing the presence <strong>of</strong> a second irreversible<br />

wave close to the electrolyte discharge. (right)<br />

The presence <strong>of</strong> a tiny reversible wave negative <strong>of</strong> the main reduction peak must be remarked,<br />

both in the CV <strong>of</strong> the trimer <strong>and</strong> that <strong>of</strong> A-α-PW 9 O 34 . The potential location <strong>of</strong><br />

this wave suggests the existence <strong>of</strong> trace amounts <strong>of</strong> PW 11 [94]. Furthermore, the current<br />

intensity <strong>of</strong> this new wave depends on the delay between the preparation <strong>of</strong> the solution<br />

<strong>of</strong> trimer or A-α-PW 9 O 34 <strong>and</strong> the recording <strong>of</strong> the CV. However, as an example, it must<br />

be noted that several hours are necessary before significant amounts <strong>of</strong> PW 11 could be<br />

observed in media <strong>of</strong> pH 4 to 5. The current intensity for the trimer solution is larger<br />

than that <strong>of</strong> A-α-PW 9 O 34 for identical concentrations <strong>of</strong> the complexes in solution. Even<br />

so, it reaches only 58.7% <strong>of</strong> the expected intensity if a complete decomposition <strong>of</strong> the<br />

trimer giving the A-α-PW 9 O 34 fragments free in solution were assumed. This difference<br />

in current intensities might be attributed to the difference in diffusion coefficients, that<br />

for the trimer being smaller than that for A-α-PW 9 O 34 . Comparison <strong>of</strong> other characteristics<br />

<strong>of</strong> the CVs <strong>of</strong> the trimer <strong>and</strong> <strong>of</strong> A-α-PW 9 O 34 support this hypothesis: the trimer<br />

is reduced at a more positive potential than A-α-PW 9 O 34 (the reduction peak potentials<br />

Ep are - 0.732 V <strong>and</strong> - 0.753 V respectively); <strong>and</strong> particularly, the overall electron transfer<br />

process associated with the redox couples, as evaluated from the anodic-to-cathodic<br />

peak potentials difference, are faster in the case <strong>of</strong> the trimer (△ Ep = 0.105 V) than for<br />

A-α-PW 9 O 34 (△ Ep = 0.160 V). This observation might be attributed to the presence<br />

<strong>of</strong> Sn-moieties attached to A-α-PW 9 O 34 fragments that modify the lacunary character<br />

49


<strong>of</strong> these fragments [88, 90]. For completeness, Figure 3.15B shows that, upon extending<br />

the potential domain in the negative direction, the CV <strong>of</strong> the trimer has a second wave<br />

peaking at - 0.840 V <strong>and</strong> which is irreversible. The corresponding wave for A-α-PW 9 O 34<br />

(not shown) appears at still more negative potential <strong>and</strong> is very close to the electrolyte<br />

limit. The reduction product(s) generated in the potential domain <strong>of</strong> this plurielectronic<br />

wave do not modify the electrode surface: as a matter <strong>of</strong> fact, it is observed that the<br />

characteristics <strong>of</strong> the first redox couple are maintained when a new voltammogram is<br />

run with this electrode. This behaviour is different from that observed with numerous<br />

plenary polyoxometalates <strong>of</strong> the Keggin <strong>and</strong> Dawson series: the reduction products generated<br />

on their last several-electron waves, modify irreversibly <strong>and</strong> persistently electrode<br />

surfaces [87, 95]. Finally, it must be pointed out that no reduction process is observed<br />

for the Sn-centers, a feature in agreement with Pope’s studies on Sn-substituted polyoxometalates<br />

[90] <strong>and</strong> our own recent work on the dimeric, tetrakis-dimethyltin assembly<br />

[{Sn(CH 3 ) 2 } 4 (H 2 P 4 W 24 O 92 ) 2 ] 28− .<br />

3.3.3 Conclusions<br />

Furthermore, 4 represents the member <strong>of</strong> a novel family <strong>of</strong> diorganotin-polyanion based<br />

cage complexes. The structure <strong>of</strong> 4 reveals that (a) the dimethyltin unit acts as a linker <strong>of</strong><br />

three (A-α-PW9O34) Keggin-type fragments, (b) the linkage is achieved via two, coplanar<br />

Sn-O(W) bonds with each <strong>of</strong> the Keggin fragments, (c) the methyl groups <strong>of</strong> the incorporated<br />

dimethyltin units are oriented trans to each other, (d) the incorporated dimethyltin<br />

units are oriented such that one <strong>of</strong> the two methyl group points inside the central polyanion<br />

cavity forming a bowl type assembly, (e) the tin atoms <strong>of</strong> the incorporated dimethyltin<br />

units have octahedral coordination geometries, (f) dimethyltin substituted polyanions<br />

tend to form cage-like structures, (g) the dimethyltin unit is an ideal, hydrolytically stable,<br />

sterically not too dem<strong>and</strong>ing building block/linker for the design <strong>of</strong> nanomolecular<br />

bowl type assembly, (h) the CV <strong>of</strong> the trimer has the same shape as that <strong>of</strong> its precursor<br />

lacunary species, A-A-α-PW 9 O 34 , thus indicating the presence <strong>of</strong> the latter without isomerization<br />

in the complex or free in solution.<br />

(Manuscript in preparation)<br />

50


3.4 The tetrameric, chiral tungstoarsenate(III),<br />

({Sn(CH 3 ) 2 (H 2 O)} 2 {Sn(CH 3 ) 2 }As 3 {α-AsW 9 O 33 } 4 ) 21−<br />

3.4.1 Experimental<br />

The tetrameric polyanion (KNH4-5) was synthesized by interaction <strong>of</strong> 0.15 g (0.66 mmol)<br />

(CH 3 ) 2 SnCl 2 with 1.58 g (0.30 mmol) K 14 [As 2 W 19 O 67 H 2 O] in 20 mL H 2 O at pH 4.1. The<br />

solution was heated to ∼80 ◦ C for 1 h <strong>and</strong> filtered after it had cooled. Addition <strong>of</strong> 0.5 mL<br />

<strong>of</strong> 1.0 M NH 4 Cl solution to the colorless filtrate <strong>and</strong> slow evaporation at room temperature<br />

led to 0.72 g (yield 67 %, based on As) <strong>of</strong> a white crystalline product after about one<br />

week.Anal. calcd. (found) for KNH4-5: K 2.6 (2.4), N 1.8 (1.9),W 61.7 (61.1), As 4.9<br />

(4.9), Sn 3.3 (3.1), C 0.7 (0.8), H 1.2 (1.0).<br />

Fig. 3.16: FTIR spectra <strong>of</strong> compound KNH4-5(red) <strong>and</strong> Na 9 [α-AsW 9 O 33 ](blue)<br />

X-ray Crystallography<br />

Single crystal X-ray analysis on K 7 (NH 4 ) 14 [(Sn(CH 3 ) 2 ) 3 (H 2 O) 2 As 3 (α-AsW 9 O 33 ) 4 ]·26H 2 O<br />

KNH4-5. A crystal <strong>of</strong> compound KNH4-5 was mounted on a glass fiber for indexing<br />

<strong>and</strong> intensity data collection at 163 K on a Bruker D8 SMART APEX CCD single-crystal<br />

51


diffractometer using Mo K α radiation [λ = 0.71073 Å]. Direct methods were used to<br />

solve the structure <strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining<br />

atoms were found from successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong><br />

polarization corrections were applied <strong>and</strong> an absorption correction was performed using<br />

the SADABS program [81]. Crystallographic data are summarized in Table 3.5<br />

Table 3.5: Crystal Data <strong>and</strong> Structure Refinement for compound KNH4-5<br />

Emperical formula As 7 C 6 H 130 K 7 N 14 O 160 Sn 3 W 36<br />

fw 10732.3<br />

space group (No.) P 2 1 /c (14)<br />

a (Å) 22.612(2)<br />

b (Å) 19.954(2)<br />

c (Å) 41.099(4)<br />

vol (Å 3 ) 18237(3)<br />

Z 4<br />

temp ( ◦ C) 27<br />

wavelength (Å) 0.71073<br />

d calcd (mg m −3 ) 3.72<br />

abs coeff. (mm −1 ) 24.53<br />

R [I > 2 σ(I)] a 0.170<br />

R w (all data) b 0.131<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

3.4.2 Results <strong>and</strong> discussion<br />

The title polyanion 5 is composed <strong>of</strong> four (B-α -AsW 9 O 33 ) fragments that are linked by<br />

three dimethyltin groups <strong>and</strong> three As(III) atoms resulting in an unprecedented, chiral<br />

cage-like assembly with C 1 symmetry (Figure 3.17, left).Polyanion 5 can also be described<br />

as a trimeric assembly [{Sn(CH 3 ) 2 (H 2 O)} 2 {Sn(CH 3 ) 2 }(α-AsW 9 O 33 ) 3 ] 21− (Figure<br />

3.17, right) which is capped by a fourth (α-AsW 9 O 33 ) Keggin fragment via three trigonal<br />

pyramidal As(III) linkers (Fig.3.18). Interestingly, the [As 3 (α-AsW 9 O 33 )] capping unit is<br />

formally neutral. The three As(III) linkers are coordinated by two oxygens <strong>of</strong> a Keggin<br />

unit from the triangular assembly <strong>and</strong> by one oxygen <strong>of</strong> the unique Keggin unit. As a result,<br />

only three <strong>of</strong> the six belt tungsten atoms <strong>of</strong> the unique (AsW 9 O 33 ) unit are involved<br />

in the bonding to adjacent Keggin building blocks in an alternating fashion.<br />

52


Fig. 3.17: Top view <strong>of</strong> [{Sn(CH 3 ) 2 (H 2 O)} 2 {Sn(CH 3 ) 2 }As 3 (α -AsW 9 O 33 ) 4 ] 21− (5).(left),the unique AsW 9<br />

fragment <strong>and</strong> its three associated As linkers are not shown for clarity.The octahedra represent WO 6 <strong>and</strong><br />

the balls are tin (green), arsenic (yellow), carbon (blue) <strong>and</strong> oxygen (red). Hydrogen atoms are omitted<br />

for clarity.(right)<br />

The orientation <strong>of</strong> this unique Keggin unit with respect to the triangular tri-Keggin<br />

fragment is the reason why 5 is chiral (Figure 3.17). The dimethyltin groups are coordinated<br />

to two oxygens <strong>of</strong> each <strong>of</strong> the adjacent Keggin units <strong>and</strong> two methyl groups<br />

which are in trans-positions, as had been observed previously for [{Sn(CH 3 ) 2 } 3 (H 2 O) 4 (β-<br />

XW 9 O 33 )] 3− (X = As III , Sb III ). However, the coordination number <strong>and</strong> geometry <strong>of</strong> the<br />

three tin centers in 5 is not equivalent. One <strong>of</strong> the tin atoms (Sn1) is six-coordinated (octahedral)<br />

whereas the other two tin atoms (Sn2, Sn3) are seven−coordinated (pentagonal<br />

bipyramidal) due to an additional terminal water lig<strong>and</strong> in the equatorial plane(Sn2−OH2<br />

2.285(19), Sn3−OH2, 2.49(4) Å). The Sn1−O(W) bonding can be described as two short<br />

bonds (2.091, 2.095(17) Å) <strong>and</strong> two very long bonds (2.400, 2.457(17) Å). On the other<br />

h<strong>and</strong>, the Sn2−O(W) bond lengths (2.300, 2.354, 2.372, 2.478(17) Å) <strong>and</strong> the Sn3−O(W)<br />

bond lengths (2.220, 2.282, 2.308, 2.462(17) Å) indicate three long bonds <strong>and</strong> one very<br />

long bond around the tin centers. The structural inequivalence between Sn1 on one side<br />

<strong>and</strong> Sn2 <strong>and</strong> Sn3 on the other side is further reflected by the C−Sn−C angles. Specifically,<br />

the C−Sn1−C angle (159.7(10) ◦ C) is significantly smaller than the C−Sn2−C (169.0(9)<br />

◦ C) <strong>and</strong> C−Sn3−C (172.4(11) ◦ C) angles, respectively. Polyanion 5 was synthesized in<br />

a simple one-pot procedure in aqueous, acidic medium by reaction <strong>of</strong> (CH 3 ) 2 SnCl 2 with<br />

K 14 [As 2 W 19 O 67 H 2 O]. We have also tried to prepare 5 by a more rational procedure (e.g.<br />

53


Fig. 3.18: Side view <strong>of</strong> [{Sn(CH 3 ) 2 (H 2 O)} 2 {Sn(CH 3 ) 2 }As 3 (α-AsW 9 O 33 ) 4 ] 21− The color code is the same<br />

as in above.<br />

3 (CH 3 ) 2 SnCl 2 + 4 Na 9 [AsW 9 O 33 ] + 1.5 As 2 O 3 ), but without success. Bond valence<br />

sum calculations indicate that the water molecules attached to Sn2 <strong>and</strong> Sn3 represent<br />

the only protonation sites in 5 [85]. All potassium ions in KNH 4 -5 could be identified<br />

crystallographically, but some <strong>of</strong> them are disordered resulting in partial (0.5) occupancy.<br />

They are located all around 5 <strong>and</strong> are coordinated to bridging <strong>and</strong> terminal oxygens <strong>of</strong><br />

the polyanion. We could not identify water molecules or cations in the central cavity<br />

<strong>of</strong> the cage-like structure. It appears that the hydrophobic nature <strong>of</strong> the pocket <strong>and</strong> its<br />

small size do not allow for inclusion <strong>of</strong> the available guests. It is difficult to determine<br />

the exact dimensions <strong>of</strong> the cavity due to the three bridging As centers <strong>and</strong> their lone<br />

pairs <strong>of</strong> electrons pointing inside the pocket. Nevertheless, we estimate the diameter <strong>of</strong><br />

the idealized spherical cavity to be around 3.5 Å. Interestingly, the three surface pores <strong>of</strong><br />

5 involving the unique Keggin unit are larger than the central cavity (around 7 Å).<br />

Solution NMR<br />

We also examined the solution properties <strong>of</strong> 5 by multinuclear NMR (D 2 O, ∼20 ◦ C).<br />

Our 119 Sn NMR measurements resulted in 2 singlets (−147.8, −158.4 ppm) with intensity<br />

54


atios 1:2, for 13 C NMR we observed a singlet at 7.4 ppm <strong>and</strong> for 1 H NMR we obtained<br />

a singlet at 0.9 ppm. This is in agreement with the solid state structure <strong>of</strong> 5, which<br />

indicates the presence <strong>of</strong> two magnetically inequivalent tin atoms. The peak at −158.4<br />

ppm is assigned to the seven−coordinated Sn2 <strong>and</strong> Sn3, whereas the signal at −147.8<br />

is due to the six−coordinated Sn1. It can be noticed that the magnetically inequivalent<br />

Sn2 <strong>and</strong> Sn3 cannot be distinguished by 119 Sn NMR spectroscopy. Apparently the chiral<br />

nature <strong>of</strong> 5, which is induced by a rotated binding mode <strong>of</strong> the unique (AsW 9 O 33 ) Keggin<br />

fragment, does not protrude all the way to the other end <strong>of</strong> the molecule. Therefore, 119 Sn<br />

NMR spectroscopy provides only information about the local coordination environment <strong>of</strong><br />

Sn2 <strong>and</strong> Sn3. The same conclusions can be drawn based on the 1 H <strong>and</strong> 13 C NMR results<br />

for 5, as all six methyl groups appear as magnetically equivalent. The actual symmetry<br />

<strong>of</strong> 5 in solution could be shown best by 183 W NMR, but unfortunately we were not able<br />

to obtain high quality spectra to date. This may be due to the very large number <strong>of</strong><br />

expected peaks (up to 36).<br />

3.4.3 Conclusions<br />

The title compound 5 represents the first polyanion containing dimethyltin <strong>and</strong> arsenic(III)<br />

linkers at the same time. Furthermore, 5 represents the first member <strong>of</strong> a novel family<br />

<strong>of</strong> diorganotin−polyanion based cage complexes. The structure <strong>of</strong> 5 reveals that (a) the<br />

dimethyltin unit acts as a linker <strong>of</strong> two (B-α-AsW 9 O 33 ) Keggin−type fragments, (b) the<br />

linkage is achieved via two, coplanar Sn−O(W) bonds with each <strong>of</strong> the Keggin fragments,<br />

(c) the methyl groups <strong>of</strong> the incorporated dimethyltin units are oriented trans<br />

to each other, (d) the incorporated dimethyltin units are oriented such that one <strong>of</strong> the<br />

two methyl group points inside the central polyanion cavity, (e) the tin atoms <strong>of</strong> the<br />

incorporated dimethyltin units can have octahedral as well as pentagonal−bipyramidal<br />

coordination geometries, (f) the diamagnetic nature <strong>of</strong> the dimethyltin unit allows for<br />

the use <strong>of</strong> multinuclear NMR as a structural characterization technique, (g) dimethyltin<br />

substituted polyanions tend to form cage-like structures, (h) the dimethyltin unit is an<br />

ideal, hydrolytically stable, sterically not too dem<strong>and</strong>ing building block/linker for the<br />

design <strong>of</strong> nanomolecular containers, (i) it is possible to prepare discrete, supramolecular<br />

55


host−guest systems with large cavities <strong>and</strong> a highly porous surface, <strong>and</strong> finally (j) we have<br />

discovered a novel class <strong>of</strong> polyanions which is <strong>of</strong> major interest for medicinal applications<br />

(e.g. antiviral) due to a unique combination <strong>of</strong> important properties (e.g. discrete,<br />

nanomolecular anion, stable at physiological pH, tightly bound organic moieties on surface,<br />

rational modification <strong>of</strong> steric <strong>and</strong> electrostatic surface properties). In summary,<br />

the large number <strong>of</strong> available Keggin− <strong>and</strong> Dawson−based lacunary polyanion precursors<br />

allows for an array <strong>of</strong> novel structural architectures to be envisioned. In addition, the<br />

possibilty <strong>of</strong> studying the mechanism <strong>of</strong> formation <strong>of</strong> such compounds accompanied by<br />

detailed investigations <strong>of</strong> applied (e.g. medicine) <strong>and</strong> academic (e.g. topology) properties<br />

indicate that 5, as well as the novel family <strong>of</strong> polyoxometalates it represents, will attract<br />

the attention <strong>of</strong> scientists from many different disciplines.<br />

3.5 The gigantic, ball-shaped heteropolytungstates<br />

({Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-XW 9 O 34 ) 12 ) 36−<br />

(X= P V , As V )<br />

3.5.1 Experimental<br />

(a)Preparation <strong>of</strong> Cs 14 Na 22 [{Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-PW 9 O 34 ) 12 ]·149H 2 O (Cs-<br />

6): A 1.46 g (0.600 mmol) sample <strong>of</strong> Na 9 [A-PW 9 O 34 ] [73] was added with stirring to<br />

a solution <strong>of</strong> 0.435 g (1.98 mmol) (CH 3 ) 2 SnCl 2 in 20 mL H 2 O. The pH was adjusted<br />

to 4 by addition <strong>of</strong> 4 M HCl. This solution was heated to ∼80 ◦ C for 1 hour <strong>and</strong><br />

then cooled to room temperature <strong>and</strong> filtered. Addition <strong>of</strong> 0.5 mL <strong>of</strong> 1.0 M CsCl<br />

solution to the colorless filtrate <strong>and</strong> slow evaporation at room temperature led to a<br />

white crystalline product after about one week. Yield: 1.3 g (69 %). FTIR spectroscopy:<br />

1084(s), 1070(s), 1019(m), 977(sh), 961(sh), 945(s), 928(s), 882(m), 833(s),<br />

774(vs), 719(vs), 668(m), 640(sh), 596(w), 574(w), 520(m), 495(sh), 478(w), 465(w)<br />

cm −1 . Anal. Calcd for Cs-6: Cs, 5.0; Na, 1.4; W, 52.8; Sn, 11.4; P, 1.0; C, 2.3; H,<br />

1.5. Found: Cs, 4.6; Na, 1.2; W, 53.6; Sn, 11.9; P, 1.2; C, 2.5; H, 1.2. (b)Preparation<br />

<strong>of</strong> Cs 14 Na 22 [Sn(CH 3 ) 2 (H 2 O) 24 Sn(CH 3 ) 212 (A-AsW 9 O 34 ) 12 ]·149H 2 O(Cs-7): The synthesis<br />

56


Fig. 3.19: FTIR spectra <strong>of</strong> compound Cs-6(red) <strong>and</strong> Na 9 [A-PW 9 O 34 ](blue)<br />

<strong>of</strong> this compound was analogous to Cs-6, but instead <strong>of</strong> Na 9 [A-PW 9 O 34 ] we used 1.59 g<br />

(0.600 mmol) <strong>of</strong> Na 8 H[A-AsW 9 O 34 ] (synthesized according to Bi et al., [72]<strong>and</strong> the pH<br />

was adjusted to 3. Yield: 1.2 g (63 %). FTIR spectroscopy: 1015(sh), 983(sh), 951(s),<br />

901(sh), 863(s), 840(sh), 773(s), 715(s), 658(s), 578(sh), 520(w), 484(w), 471(w), 411(m)<br />

cm −1 . Anal. Calcd for Cs-7: Cs, 4.9; Na, 1.3; W, 52.1; Sn, 11.2; As, 2.4; C, 2.3; H, 1.5.<br />

Found: Cs, 4.5; Na, 1.2; W, 53.1; Sn, 11.6; As, 2.6; C, 2.4; H, 1.6. Elemental analyses were<br />

Fig. 3.20: FTIR spectra <strong>of</strong> compound Cs-7(red)<strong>and</strong> Na 8 H[A-AsW 9 O 34 ](blue)<br />

performed by Kanti Labs Ltd. in Mississauga, Canada. Both compounds crystallized as<br />

mixed cesium-sodium salts <strong>and</strong> are in fact isomorphous. Due to the high symmetry <strong>of</strong><br />

the solid state arrangement (space group I m¯3).<br />

57


X-ray Crystallography<br />

Single crystal X-ray analysis on Cs 14 Na 22 [{Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-PW 9 O 34 ) 12 ]<br />

·149H 2 O Cs-6<strong>and</strong> Cs 14 Na 22 [Sn(CH 3 ) 2 (H 2 O) 24 {Sn(CH 3 ) 2 } 12 (A-AsW 9 O 34 ) 12 ]·149H 2 O Cs-<br />

7. The respective crystals were mounted on a glass fiber for indexing <strong>and</strong> intensity data<br />

collection at 173 K on a Bruker D8 SMART APEX CCD single-crystal diffractometer<br />

using Mo K α radiation (λ = 0.71073 Å). Direct methods were used to solve the structure<br />

<strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining atoms were found from<br />

successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong> polarization corrections<br />

were applied <strong>and</strong> an absorption correction was performed using the SADABS program<br />

[81]. Crystallographic data are summarized in Table 3.6.<br />

Table 3.6: Crystal Data <strong>and</strong> Structure Refinement for compounds Cs-6 <strong>and</strong> Cs-7<br />

crystal system Cubic Cubic<br />

Empirical formula C 72 H 562 Cs 14 Na 22 O 581 P 12 Sn 36 W 108 C 72 H 562 As 12 Cs 14 Na 22 O 581 Sn 36 W 108<br />

fw (g/mol) 37593.3 38120.7<br />

space group I m¯3 (204) I m¯3 (204)<br />

a Å 32.7441(4) 32.8121(4)<br />

volume (Å 3 ) 35107.44(74) 35326.62(75)<br />

Z 2 2<br />

temp. ( ◦ C) -100 -100<br />

wavelength (Å) 0.71073 0.71071<br />

dcalc (mg m −3 ) 3.488 3.516<br />

R [I > 2 σ(I)] a 0.0635 0.0564<br />

R w (all data) b 0.0635 0.0564<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

The asymmetric unit <strong>of</strong> Cs-6 <strong>and</strong> Cs-7 is very small <strong>and</strong> includes only 5 tungsten<br />

<strong>and</strong> two tin atoms (see Figure 3.21 ).<br />

The spherical structure <strong>of</strong> Cs-6 <strong>and</strong> Cs-7 is truly spectacular in terms <strong>of</strong> geometry<br />

<strong>and</strong> size (diameter <strong>of</strong> 30 Å) <strong>and</strong> is completely unprecedented in polyoxotungstate<br />

chemistry (see Figures 3.22). This supramolecular assembly is composed <strong>of</strong> 12 trilacunary<br />

[A-XW 9 O 34 ] 9− (X = P v , As v ) Keggin fragments which are linked by a total <strong>of</strong> 36<br />

dimethyltin groups (12 inner (CH 3 ) 2 Sn 2+ <strong>and</strong> 24 outer (CH 3 ) 2 (H 2 O)Sn 2+ groups) resulting<br />

in a polyanion with T h symmetry. Interestingly, Cs-6 <strong>and</strong> Cs-7 have almost 1000<br />

atoms <strong>and</strong> a molar mass <strong>of</strong> around 33000 g/mol. In addition, 14 cesium ions are closely<br />

58


Fig. 3.21: Ball <strong>and</strong> stick representation <strong>of</strong> the asymmetric unit <strong>of</strong> Cs-6 <strong>and</strong> Cs-7 showing the same<br />

labeling scheme as both compounds are isostructural<br />

Fig. 3.22: Left:Ball <strong>and</strong> stick representation <strong>of</strong> Cs-6 <strong>and</strong> Cs-7 including the 14 cesium counter ions. The<br />

color code is as follows: tungsten (black), tin (blue), phosphorus/arsenic (yellow), oxygen (red), carbon<br />

(green) <strong>and</strong> cesium (purple). No hydrogens are shown for clarity, Right:Polyhedral representation <strong>of</strong> Cs-6<br />

<strong>and</strong> Cs-7. The WO 6 octahedra are red <strong>and</strong> the XO 4 tetrahedra (X = P, As) are yellow. Otherwise, the<br />

labeling scheme is the same as in Figure 3.21A. No hydrogens <strong>and</strong> cesiums shown for clarity<br />

associated with Cs-6 <strong>and</strong> Cs-7 in the solid state. They are located in hydrophilic surface<br />

pockets <strong>of</strong> the spherical clusters, thereby stabilizing the assembly further (see Figure 3.22<br />

Left).<br />

59


3.5.2 Results <strong>and</strong> discussions<br />

Polyanions Cs-6 <strong>and</strong> Cs-7 represent the second largest, discrete polyoxotungstates ever<br />

reported [96]. Furthermore, the overall shape <strong>of</strong> Cs-6 <strong>and</strong> Cs-7 resembles Müllers<br />

Keplerates which are highly-symmetrical polyoxomolybdates [97]. This class <strong>of</strong> compounds<br />

is characterized by spherical clusters with icosahedral symmetry <strong>of</strong> the type<br />

(pentagon) 12 (linker) 30 where the centers <strong>of</strong> the 12 pentagons span an icosahedron <strong>and</strong><br />

the centers <strong>of</strong> the 30 linkers an icosidodecahedron. Close inspection <strong>of</strong> the structure <strong>of</strong><br />

Cs-6 <strong>and</strong> Cs-7 indicates that it is best described as a ‘pseudo Keplerate’. Although the<br />

12 hetero atoms (P in Cs-6 <strong>and</strong> As in Cs-7) span an almost perfect icosahedron (see<br />

Figure 3.23), the 12 inner tin atoms do not (see Figure 3.24A). Interestingly, also the 24<br />

Fig. 3.23: Representation <strong>of</strong> the icosahedron spanned by the hetero atoms <strong>of</strong> Cs-6(P) <strong>and</strong> Cs-7(As)<br />

outer Sn atoms form a highly symmetrical arrangement (see Figure 3.24B) <strong>and</strong> so do the<br />

14 cesium ions which span a hexa-capped cube (see Figure 3.25). Therefore polyanions<br />

6 <strong>and</strong> 7 exhibit some analogies with a Russian doll, but are in fact more complex as the<br />

shells have varying size, symmetry <strong>and</strong> chemical composition. The multi-shell nature <strong>of</strong> 6<br />

<strong>and</strong> 7 is nicely visible in Figure 3.26 (cesium ions not included for clarity). Bond valence<br />

sum calculations (BVS) <strong>of</strong> 6 <strong>and</strong> 7 indicate that the terminal oxygens attached to the 24<br />

outer tin atoms are actually water molecules. There are no other protonation sites on 6<br />

60


Fig. 3.24: Left:Representation <strong>of</strong> the polyhedron spanned by the 12 inner tin atoms <strong>of</strong> 6 <strong>and</strong> 7.Right:<br />

Representation <strong>of</strong> the polyhedron spanned by the 24 outer tin atoms <strong>of</strong> Cs-6 <strong>and</strong> Cs-7<br />

Fig. 3.25: Representation <strong>of</strong> the hexa-capped cube spanned by the 14 cesium ions <strong>of</strong> Cs-6 <strong>and</strong> Cs-7<br />

<strong>and</strong> 7 <strong>and</strong> therefore the charge must be -36 [85]. This is fully consistent with elemental<br />

analysis, which indicated the presence <strong>of</strong> 14 cesium <strong>and</strong> 22 sodium ions. The former could<br />

be identified by X-ray diffraction, but not the latter which is probably due to disorder.<br />

The central cavity <strong>of</strong> the ball-shaped 6 <strong>and</strong> 7 has a diameter <strong>of</strong> around 8 Å <strong>and</strong> it does not<br />

contain any water molecules or other ions. It can be noticed that the pocket is actually<br />

hydrophobic because it is lined by a total <strong>of</strong> 12 methyl groups. This probably explains<br />

why polyanions 6 <strong>and</strong> 7, which were synthesized in aqueous medium, do not contain any<br />

guests. Nevertheless, we believe that in principle small guest molecules with appropriate<br />

61


Fig. 3.26: Ball <strong>and</strong> stick representation <strong>of</strong> Cs-6 <strong>and</strong> Cs-7 highlighting the different polyhedral shells<br />

(12 inner Sn atoms: red, 12 hetero atoms: purple, 24 outer Sn atoms: yellow)Courtesy: Dr. H. Bögge,<br />

<strong>University</strong> <strong>of</strong> Bielefeld, Germany.<br />

size <strong>and</strong> polarity can be encapsulated during formation <strong>of</strong> 6 <strong>and</strong> 7. Furthermore, we identified<br />

hydrophobic channels (lined by methyl groups) passing through the entire structure<br />

<strong>of</strong> 6 <strong>and</strong> 7, which could allow for loading <strong>and</strong> discharging <strong>of</strong> guest molecules also after<br />

the polyanions have been formed. In addition the surface <strong>of</strong> 6 <strong>and</strong> 7 has a total <strong>of</strong> 14<br />

hydrophilic pockets, which are all occupied by cesium ions (see Figure 3.22).<br />

Solution NMR<br />

We also investigated the solution properties <strong>of</strong> 6 by multinuclear solution NMR ( 183 W,<br />

119 Sn, 31 P, 13 C, 1 H) at room temperature in D 2 O using a 400 MHz JEOL ECX instrument.<br />

The 183 W NMR spectrum <strong>of</strong> 6 exhibits two singlets at -134.0 <strong>and</strong> -157.9 ppm, respectively,<br />

with an intensity ratio 1:2. The 31 P NMR spectrum <strong>of</strong> 6 shows a singlet at -13.1 ppm.<br />

The 119 Sn NMR spectrum <strong>of</strong> 6 shows a peak at -170.8 ppm. The 13 C NMR spectrum<br />

exhibits two peaks at 22.1 <strong>and</strong> 8.4 ppm, respectively, with an intensity ratio 1:2 <strong>and</strong> 1 H<br />

NMR also shows two peaks at 1.9 <strong>and</strong> 0.8 ppm, respectively, with an intensity ratio 1:2. In<br />

summary, for solutions <strong>of</strong> 6 we identified 2 types <strong>of</strong> W, 1 type <strong>of</strong> P, <strong>and</strong> 2 types each <strong>of</strong> C<br />

<strong>and</strong> H by NMR. The 183 W NMR results indicate that all 12 Keggin units are magnetically<br />

equivalent <strong>and</strong> the 6 belt tungsten atoms (-157.9 ppm) can be distinguished from the 3<br />

cap tungstens (-134.0 ppm). As expected, all 12 P atoms are fully equivalent. We do not<br />

62


Fig. 3.27: 31 P(Left), <strong>and</strong> 183 W NMR spectra <strong>of</strong><br />

Cs 14 Na 22 {Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-PW 9 O 34 ) 12 ]·149H 2 O<br />

have a good explanation why solution NMR does not allow to distinguish the two types <strong>of</strong><br />

tin atoms which we identify based on the crystal structure (see Figures 3.26). The same<br />

is true for the number <strong>and</strong> relative intensities <strong>of</strong> the 13 C <strong>and</strong> 1 H peaks.<br />

3.5.3 STM studies <strong>of</strong> Cs-6<br />

The scanning tunneling microscope (STM) [98] has become an important tool to image<br />

surface with atomic resolution. There has been particular interest in recent years to obtain<br />

images <strong>of</strong> individual molecules adsorbed on surfaces, even if these substances are<br />

insulating in their bulk structure [99, 100]. Many insulating materials have been successfully<br />

imaged using STM by depositing these materials on conductive substrates [101].<br />

Particularly polyoxometalates (POMs) which form self-assembled monolayers as well as<br />

individual species in a periodic arrangement on different substrates were imaged by STM<br />

at small bias voltages [101–106]. This is surprising, since the gap energy between highest<br />

occupied molecular orbitals (HOMO) <strong>and</strong> lowest unoccupied molecular orbital (LUMO)<br />

<strong>of</strong> the isolated molecules is relatively large. One possible explanation <strong>of</strong> this type <strong>of</strong><br />

imaging could be the decrease <strong>of</strong> the HOMO-LUMO energy gap due to molecule-molecule<br />

or molecule-substrate interactions [107]. Although the STM is capable <strong>of</strong> atomic resolution<br />

when imaging solid surfaces, mapping <strong>of</strong> large, complex molecules with submolecular<br />

resolution is a difficult task. An STM image contains both geometric <strong>and</strong> electronic information<br />

about the sample in a complicated way [108–110]. Highly resolved STM images<br />

63


can distinguish between different molecules with very similar geometric structures. As,<br />

for example, we would consider the van der Waals surface <strong>of</strong> large <strong>and</strong> complex molecules,<br />

we would see only a featureless “blob“ <strong>of</strong> a large cluster <strong>of</strong> atoms. However, spectroscopic<br />

possibilities <strong>of</strong> STM allow us to probe electronic states <strong>of</strong> the molecules as a function <strong>of</strong><br />

energy [111, 112]. If there would be subunits <strong>of</strong> the molecule exhibiting a special type<br />

<strong>of</strong> chemical bonding, STM spectroscopy therefore allows filtering out special features <strong>of</strong><br />

the species if they arise at much different energies. The objective <strong>of</strong> this present work is<br />

to achieve topographic as well as spatially resolved electronic structural information <strong>of</strong><br />

ball-shaped heteropolytungstate complexes. Scanning tunneling microscopy (STM) <strong>and</strong><br />

scanning tunneling spectroscopy (STS) were performed on single molecules adsorbed onto<br />

HOPG. Several groups have been reported the formation <strong>of</strong> ordered <strong>and</strong> monolayer arrays<br />

<strong>of</strong> POMs on surfaces <strong>and</strong> have imaged these using STM [101–106, 113, 114]. However,<br />

in our case we concentrate on isolated, free st<strong>and</strong>ing single molecules in order to find out<br />

specific features <strong>of</strong> the electronics properties at the single- molecule level. Recently the<br />

investigations <strong>of</strong> specific functionalities <strong>of</strong> molecular nanostructures at surfaces were reviewed<br />

[115, 116]. STS measurements are very <strong>of</strong>ten carried out under ultra high vacuum<br />

conditions with low temperatures to increase signal-to-noise ratio [117, 118]. However,<br />

it has been shown that under certain conditions STM experiments are also capable to<br />

detect local electronic properties at room temperature [119–121]. A home-made scanning<br />

tunneling microscope was used for measurements. The microscope was equipped<br />

with a commercially available low current control system (RHK Technology). Samples<br />

<strong>of</strong> Cs-6 on HOPG were conveniently prepared by allowing 10 −9 M aqueous solutions <strong>of</strong><br />

pH 4 to 6 to evaporate under air. HOPG is one <strong>of</strong> the best studied substrate concerning<br />

its electronic properties both experimentally <strong>and</strong> theoretically [122]. Before adding the<br />

solution onto the substrate surface, we ensured that the tunneling tip had a sufficiently<br />

high resolution. We calibrated distances in the STM images by observing atomic spacing<br />

on highly oriented pyrolytic graphite (HOPG). Typically, for the STM measurements,<br />

tunneling currents between 5 <strong>and</strong> 200 pA were employed. The bias voltage was 50 mV<br />

to 500 mV. The scan frequency was varied between 2 <strong>and</strong> 5 Hz. Resolution was 256×256<br />

points for topography, <strong>and</strong> 128×128 in the STS measurements. STS studies have been<br />

64


performed in the current imaging tunneling spectroscopy mode (CITS) simultaneously<br />

with constant current image by the interrupted-feedback-loop technique. This was carried<br />

out by opening the feedback loop to fix the separation between the tip <strong>and</strong> sample,<br />

<strong>and</strong> ramping the bias voltage over the range <strong>of</strong> interest. The scan range <strong>of</strong> voltages was<br />

typically from -1 V to 0.1 V relative to the tip potential for 113 discrete voltage steps.<br />

Typically, tunneling resistances <strong>of</strong> the order <strong>of</strong> 5 G were set. We used mechanically cut<br />

Pt-Ir (90/10) tips from wires with a diameter <strong>of</strong> 0.25 mm.<br />

Results <strong>and</strong> discussion on STM studies <strong>of</strong> Cs-6<br />

The chemical structure <strong>of</strong> [{Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-PW 9 O 34 ) 12 ] 36− 6 complex<br />

is represented in Figure 3.22. Molecular ordering on a surface is controlled by a delicate<br />

balance between intermolecular forces <strong>and</strong> molecule-substrate interactions. As the bonding<br />

<strong>of</strong> the complex supramolecules to the substrate is weak, diffusion is possible along<br />

the surface. The mobility is reduced at defects, e.g. monoatomic graphite steps. Thus<br />

an aggregation <strong>of</strong> molecules is to be expected, particularly in such places. In very dilute<br />

conditions, the supramolecules get trapped along the defects <strong>of</strong> HOPG surface. High resolution<br />

STM images with increasing magnification <strong>of</strong> Cs-6 are depicted in Figure 3.28.<br />

The images reveal the well-ordered distribution <strong>and</strong> adsorption geometry <strong>of</strong> Cs-6 even at<br />

room temperature. The molecules are attached to the graphite defects in a linear fashion.<br />

Individual molecules were clearly distinguished <strong>and</strong> measured. The substrate was easily<br />

imaged simultaneously with the molecules. Figure 3.28 shows both, the substrate <strong>and</strong><br />

adsorbate, <strong>and</strong> can help to derive the exact adsorption geometry. The periodicity <strong>of</strong> the<br />

molecules is larger than the size <strong>of</strong> the individual molecules. This might have caused due to<br />

the strong repulsive forces between the large ion complexes. The apparent size <strong>of</strong> the molecules<br />

in the STM images is in accordance with the calculated molecular size from X-ray<br />

structure <strong>of</strong> 3 nm. The six-fold symmetry <strong>of</strong> the molecule is clearly discernible. In order<br />

to study the electronic properties <strong>of</strong> [{Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-PW 9 O 34 ) 12 ] 36−<br />

complex we applied CITS technique. CITS is current imaging tunneling spectroscopy<br />

<strong>and</strong> has been developed for scanning tunneling spectroscopy (STS) <strong>of</strong> a sample surface<br />

[123, 124]. CITS involves the measurement <strong>of</strong> the I-V characteristics at each pixel that the<br />

65


Fig. 3.28: STM pictures <strong>of</strong> {Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-PW 9 O 34 ) 12 ] 36−<br />

normal STM topography is taken. The tip to sample distance is defined by the topography<br />

parameters. One then has a normal STM image as well as current images, where the<br />

current images are obtained from the three-dimensional data structure <strong>of</strong> I (V, x, y). The<br />

current image thus represents a slice, at a given voltage, <strong>of</strong> the current as a function <strong>of</strong> the<br />

lateral x, y coordinates. Variations on this scheme can be performed whereby I-V characteristics<br />

are recorded at a subset <strong>of</strong> points in a grid or along a particular line in an image.<br />

The current contrast changes significantly when at certain bias voltages new molecular<br />

energy levels come into play thus enhancing the information obtained from topography<br />

alone. The use <strong>of</strong> current imaging, in particular with conductance measurements, allows<br />

energy-resolved spectroscopy to be performed with spatial emphasis. One can see the<br />

location <strong>of</strong> states as a function <strong>of</strong> energy <strong>and</strong> position [125]. CITS technique has been<br />

applied successfully to semiconductor materials but its application to organic molecules<br />

is rather difficult because <strong>of</strong> mobility or instability <strong>of</strong> the molecules, drift, etc [126]. With<br />

our home built low drift STM head we successfully applied this technique to the P-ball<br />

shaped complex, Figure 3.22 but we did not see any remarkable features in current images<br />

from voltages 0.1 to below -1.0 V. We assume that [A-PW 9 O 34 ] 9− fragments <strong>of</strong> the<br />

complex, Figure 3.28 could be far away from the Fermi level. That’s why these states do<br />

not play a significant role in the tunneling current up to the applied voltage [127].<br />

The STM studies was done by our collaborators Pr<strong>of</strong>. P. Müller <strong>and</strong> his coworkers Mr. M.<br />

S. Alam, V. Dremov, Physikalisches Institut III, (Universität Erlangen-Nürnberg, Germany.<br />

66


3.5.4 HPPS measurement<br />

The dynamic light scattering measurements was done on a freshly synthesized solution <strong>and</strong><br />

crystal redissolved <strong>of</strong> polyanion 6 <strong>and</strong> solution <strong>of</strong> polyanion 7 in collaboration with Pr<strong>of</strong>.<br />

M. Winterhalter <strong>and</strong> his group (IUB, Germany). The data obtained from high performance<br />

particle sizer(HPPS) provided by Malvern instruments, suggest that the polyanion<br />

6 were monodispersed. The size distribution by intensity for polyanion showed two kinds<br />

<strong>of</strong> particle, a size <strong>of</strong> diameter 148 nm <strong>and</strong> 1.1 nm but size distribution by number <strong>and</strong><br />

suggest that the particle <strong>of</strong> size 1.1 nm is more populated. (See Figure.3.29)<br />

The light scattering experiment for polyanion 7 suggest that the polyanion were monodis-<br />

Fig. 3.29: Size distribution by intensity <strong>and</strong> number <strong>of</strong> polyanion 6<br />

persed in solution <strong>and</strong> have identical size distribution by intensity <strong>and</strong> number which<br />

indicated that only one kind <strong>of</strong> particle size (1.2 nm), were seen to be populated. (See<br />

Figure.3.30) Crystal <strong>of</strong> polyanion Cs-6 were redissolved in same concentration <strong>of</strong> the<br />

Fig. 3.30: Size distribution by intensity <strong>and</strong> number <strong>of</strong> polyanion 7<br />

synthesized solution. The light scattering experiment suggest that the polyanion were<br />

monodispersed in solution <strong>and</strong> have identical size distribution by intensity <strong>and</strong> number<br />

which indicated that only one kind <strong>of</strong> particle size (126 nm), were seen to be populated.<br />

(See Figure.3.31) A detailed study on dynamic light scattering <strong>of</strong> polyanion 6 <strong>and</strong> polyan-<br />

67


Fig. 3.31: Size distribution by intensity <strong>and</strong> number <strong>of</strong> polyanion Cs-6<br />

ion 7 are on progress to underst<strong>and</strong> the dynamic behavior <strong>of</strong> the polyanion in solution<br />

<strong>and</strong> for the redissolved crystal. All the measurement were done in aqueous medium in a<br />

polystyrene cuvette.<br />

3.5.5 Conclusions<br />

In summary, we have synthesized the supramolecular, spherical polyoxotungstate assemblies<br />

Cs-6 <strong>and</strong> Cs-7 using a simple one-pot procedure in aqueous medium. The structures<br />

<strong>of</strong> these compounds are completely unprecedented <strong>and</strong> allow for a multitude <strong>of</strong> studies including<br />

host/guest chemistry, ion exchange, gas storage, catalysis <strong>and</strong> medicine. We have<br />

demonstrated that the dimethlytin group is a highly reactive electrophile which allows to<br />

link lacunary polyanion fragments in an unprecedented fashion. The resulting compounds<br />

are diamagnetic <strong>and</strong> therefore multinuclear solution NMR studies can be performed, which<br />

is <strong>of</strong> major importance for medicinal applications. The fact that all tungsten centers are<br />

in the fully oxidized +6 state also allows for unequivocal determination <strong>of</strong> the charges <strong>of</strong><br />

the product polyanions.<br />

68


Part-II<br />

Mono-Organo Tin POMs


3.6 The bis-phenyltin substituted, lone pair containing<br />

tungstoarsenate:<br />

({C 6 H 5 Sn} 2 As 2 W 19 O 67 (H 2 O)) 8−<br />

3.6.1 Experimental<br />

The precursor K 14 [As 2 W 19 O 67 (H 2 O)] was synthesized according to the published procedure<br />

<strong>of</strong> Kortz et al. <strong>and</strong> the purity was confirmed by infrared spectroscopy [33]. All other<br />

reagents were used as purchased without further purification. (NH 4 ) 7 Na[(C 6 H 5 Sn) 2 As 2 W 19<br />

O 67 (H 2 O)]·17.5H 2 O (NH4-8). The title compound was synthesized by dissolving 0.145<br />

mL (0.88 mmols) <strong>of</strong> C 6 H 5 SnCl 3 in 40 mL H 2 O followed by addition <strong>of</strong> 2.10 g (0.40 mmols)<br />

K 14 [As 2 W 19 O 67 (H 2 O)]. This solution (pH 1.6) was heated to 80 ∼ 80 ◦ C for 1 h <strong>and</strong> then<br />

cooled to room temperature. The solution was filtered <strong>and</strong> a few drops <strong>of</strong> 0.1 M NH 4 Cl<br />

<strong>and</strong> 0.1 M NaCl solution were added <strong>and</strong> then the solution was allowed to evaporate in<br />

an open vial at room temperature. A white crystalline product started to appear after a<br />

week. Evaporation was continued until the solvent approached the solid product (yield<br />

1.6 g, 72%). FTIR spectra for (NH 4 ) 7 Na[(C 6 H 5 Sn) 2 As 2 W 19 O 67 (H 2 O)]·17.5H 2 O: 964(m),<br />

905(m), 881(m), 860(m), 801(sh), 754(s), 736(s), 702(sh), 582(w), 518(sh), 486(w), 446(w)<br />

cm −1 . Anal. Calcd (Found) for (NH 4 ) 7 Na[(C 6 H 5 Sn) 2 As 2 W 19 O 67 (H 2 O)]·17.5H 2 O: N 1.8<br />

(1.6), Na 0.4 (0.2), Sn 4.2 (4.3), W 62.4 (61.6), As 2.7 (2.7), C 2.6 (2.7), H 1.6 (1.5). NMR<br />

spectroscopy <strong>of</strong> 8 at pH 1.6 (D 2 O, 293 K): 183 W: -110.5, -120.7, -150.0, -159.6, -170.5,<br />

-196.9 ppm (all singlets with intensities 4:2:4:4:4:1); 119 SnH: -432.5 ppm (singlet); 13 CH:<br />

129.7, 132.0, 133.6, 137.3 ppm (singlets); 1 H: 7.6, 8.1 ppm (multiplets). Monophenyltrichloride<br />

(C 6 H 5 SnCl 3 ) in H 2 O at pH 1.6, 119 SnH: -536.5 ppm (singlet); 13 CH: 128.9,<br />

130.4, 133.1, 134.3 ppm (singlets); 1H: 7.3, 7.5 ppm (multiplets). Elemental analysis<br />

was performed by Kanti Labs Ltd. in Mississauga, Canada. The FTIR spectrum was<br />

recorded on a Nicolet Avatar FTIR spectrophotometer in a KBr pellet. All NMR spectra<br />

were recorded on a JEOL Eclipse 400 instrument at room temperature using D 2 O as a<br />

solvent.<br />

70


Fig. 3.32: FTIR spectra <strong>of</strong> compound NH4-8(red) <strong>and</strong> K 14 [As 2 W 19 O 67 (H 2 O)](blue)<br />

X-ray Crystallography<br />

A crystal <strong>of</strong> compound NH4-8 was mounted on a glass fiber for indexing <strong>and</strong> intensity<br />

data collection at 163 K on a Bruker D8 SMART APEX CCD single-crystal diffractometer<br />

using Mo K radiation ( λ = 0.71073 Å). Direct methods were used to solve the structure<br />

<strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining atoms were found from<br />

successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong> polarization corrections<br />

were applied <strong>and</strong> an absorption correction was performed using the SADABS program<br />

[81]. Crystallographic data are summarized in Table 3.7<br />

3.6.2 Results <strong>and</strong> discussion<br />

The novel bis-phenyltin substituted, dimeric tungstoarsenate(III) [(C 6 H 5 Sn) 2 As 2 W 19 O 67<br />

(H 2 O)] 8− (6) consists <strong>of</strong> two lacunary B-α-[AsW 9 O 33 ] 9− Keggin fragments linked via<br />

two (C 6 H 5 Sn) 3+ groups <strong>and</strong> a WO(H 2 O) 4+ moiety leading to a structure with nominal<br />

C 2v symmetry (see Figure 3.33). Alternatively 8 can be described as a dilacunary<br />

[As 2 W 19 O 67 (H 2 O)] 14− fragment which has taken up two organotin units. It can be noticed<br />

that the tin atoms are situated well above the plane <strong>of</strong> the four equatorial 2-oxo<br />

lig<strong>and</strong>s. Therefore the tin atoms are displaced towards the terminal phenyl lig<strong>and</strong>, i.e.<br />

71


Table 3.7: Crystal Data <strong>and</strong> Structure Refinement for compound NH4-8<br />

Emperical formula As 2 C 12 H 75 N 7 NaO 85 .5Sn 2 W 19<br />

fw 5594.1<br />

space group (No.) P 2 1 /c (14)<br />

a (Å) 18.3127(17)<br />

b (Å) 24.403(2)<br />

c (Å) 22.965(2)<br />

vol (Å 3 ) 9854.0(16)<br />

Z 4<br />

temp ( ◦ C) -110<br />

wavelength (Å) 0.71073<br />

d calcd (mg m −3 ) 3.73<br />

abs coeff. (mm −1 ) 23.354<br />

R [I > 2 σ(I)] a 0.075<br />

R w (all data) b 0.146<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

Fig. 3.33: Combined polyhedron <strong>and</strong> ball/stick representations <strong>of</strong> [(C 6 H 5 Sn) 2 As 2 W 19 O 67<br />

(H 2 O)] 8− (left),Top view (right)<br />

towards the exterior <strong>of</strong> the equator <strong>of</strong> 8. This resembles the displacement <strong>of</strong> all tungsten<br />

sites <strong>of</strong> 8 towards the external, terminal oxo groups. On the other h<strong>and</strong>, the unique tungsten<br />

atom in the central belt <strong>of</strong> 8 is displaced towards the interior <strong>of</strong> the equator <strong>of</strong> 8. As<br />

a result <strong>of</strong> this symmetry breaking, polyanion 8 exhibits a surface with both positive <strong>and</strong><br />

negative curvature. Bond-valence-sum (BVS) calculations for 8 indicated that no oxygen<br />

72


<strong>of</strong> the two (AsW 9 O 33 ) caps is protonated [85]. However, the central tungsten atom has<br />

two trans related lig<strong>and</strong>s which are a water molecule <strong>and</strong> an oxo group. The latter is<br />

inside the central cavity <strong>of</strong> 8 whereas the former is on the outside. This is in complete<br />

agreement with the single-crystal XRD data <strong>of</strong> related structures that contain one or three<br />

central tungsten atoms linking two (α-AsW 9 O 33 ) fragments (e.g. [As 2 W 19 O 67 (H 2 O)] 14− ,<br />

[As 2 W 21 O 69 (H 2 O)] 6− ) [33]. The title polyanion 8 was synthesized rationally <strong>and</strong> in a<br />

one-pot reaction by interaction <strong>of</strong> C 6 H 5 SnCl 3 with K 14 [As 2 W 19 O 67 (H 2 O)] in aqueous,<br />

acidic medium (pH 2). However, we obtained 8 for the first time accidentally during<br />

our efforts to discover novel diphenyltin substituted species. Interaction <strong>of</strong> (C 6 H 5 ) 2 SnCl 2<br />

with Na 9 [α-AsW 9 O 33 ] <strong>and</strong> Na 2 WO 4 in a molar ratio <strong>of</strong> 3:1:3 in aqueous medium at pH<br />

2 resulted in 8. This means that the diphenyltin precursor underwent partial hydrolysis<br />

resulting in the loss <strong>of</strong> one phenyl group. Then we decided to reproduce this compound<br />

via a more rational synthetic procedure using the dilacunary polyoxotungstate precursor<br />

[As 2 W 19 O 67 (H 2 O)] 14− <strong>and</strong> (C 6 H 5 )SnCl 3 (see Experimental section). The tungstoarsenate<br />

[As 2 W 19 O 67 (H 2 O)] 14− was synthesized for the first time about 30 years ago by Tourné<br />

et al. <strong>and</strong> recently Kortz et al. confirmed the proposed structure by X-ray diffraction<br />

[33, 128, 129]. Interestingly, to date only a few polyoxoanions have been synthesized<br />

using [As 2 W 19 O 67 (H 2 O)] 14− as a precursor [33, 130, 131]. Polyanion 8 represents<br />

a novel member in the class <strong>of</strong> monoorganotin substituted polyoxotungstates in general<br />

<strong>and</strong> the subclass <strong>of</strong> tungstoarsenates(III) in particular. Pope et al. have synthesized<br />

<strong>and</strong> characterized several monomeric <strong>and</strong> dimeric polyoxotungstates substituted by<br />

monoorganotin functions [44–46, 132]. Amongst them the tetrakis(monophenyltin) substituted<br />

tungstoarsenate(III) [(C 6 H 5 Sn) 2 O 2 H(α-AsW 9 O 33 ) 2 ] 9− <strong>and</strong> the tris(monophenyltin)<br />

substituted tungstoantimonate(III) [(C 6 H 5 Sn) 3 Na 3 (H 2 O) 6 (α-SbW 9 O 33 ) 2 ] 6− are related to<br />

8 [47]. These species were synthesized by reaction <strong>of</strong> C 6 H 5 SnCl 3 with Na 9 [α-AsW 9 O 33 ]<br />

<strong>and</strong> Na 9 [α-SbW 9 O 33 ], respectively, in aqueous acidic medium (pH 2). The antimony<br />

derivative [(C 6 H 5 Sn) 3 Na 3 (H 2 O) 6 (α -SbW 9 O 33 ) 2 ] 6− <strong>and</strong> 8 exhibit both a dimeric s<strong>and</strong>wich<br />

type structure with square pyramidal organotin groups, whereas [(C 6 H 5 Sn) 2 O 2 H(α-<br />

AsW 9 O 33 ) 2 ] 9− contains four monoorganotin units which are all octahedral. The solid state<br />

structure <strong>of</strong> NH4-8 indicates that the title polyanion contains a sodium atom in the cen-<br />

73


tral belt in-between the two tin atoms (dNa···Sn = 3.52 - 3.55(1) Å, see Figure 3.33).<br />

The sodium ion is six-coordinated by four bridging oxo-groups <strong>of</strong> 8 <strong>and</strong> two terminal<br />

water molecules resulting in an octahedral coordination sphere <strong>and</strong> typical bond lengths<br />

(dNa-O = 2.29 - 2.46(2) Å). The presence <strong>of</strong> a sodium ion in the belt <strong>of</strong> 8 is not all that<br />

surprising, as crystal structures <strong>of</strong> several derivatives <strong>of</strong> 8 have also revealed the presence<br />

<strong>of</strong> sodium ions in analogous positions. In[(C 6 H 5 Sn) 3 Na 3 (H 2 O) 6 ( α-SbW 9 O 33 ) 2 ] 6− three<br />

sodium ions are located in the central belt in addition to the three [C 6 H 5 Sn] 3+ groups<br />

[47]. This is analogous to the di- <strong>and</strong> tri-transition metal substituted derivatives <strong>of</strong> this<br />

structural type, [M 2 (H 2 O) 2 WO(H 2 O)Na 3 (H 2 O) 6 ( α-AsW 9 O 33 ) 2 ] 7− (M = Co 2+ , Zn 2+ ) <strong>and</strong><br />

[M 3 (H 2 O) 3 Na 3 (H 2 O) 6 (α-XW 9 O 33 ) 2 ] 9− (X = As III , M = Mn 2+ , Co 2+ , Cu 2+ , Zn 2+ ; X =<br />

Sb III , M = Cu 2+ , Zn 2+ ) [39, 133]. The -8 charge <strong>of</strong> the polyanion is balanced by the unique<br />

sodium ion in the belt <strong>of</strong> 8 <strong>and</strong> seven ammonium ions which surround the polyanion. The<br />

exact positions <strong>of</strong> the ammonium ions could not be identified by X-ray diffraction as they<br />

could not be distinguished from water molecules. However, the result <strong>of</strong> elemental analysis<br />

is in complete agreement with the formula <strong>of</strong> NH4-8. The solid state arrangement <strong>of</strong><br />

8 deserves special attention (see Figure 3.34). The title polyanions exhibit 2-D packing in<br />

Fig. 3.34: Projection <strong>of</strong> the crystal packing on the bc plane showing the 2-D arrangement <strong>of</strong> compound 8<br />

the bc plane <strong>and</strong> additional intermolecular connectivities in the c direction via W-O-Na<br />

bonds lead to the formation <strong>of</strong> chains. Interestingly the phenyl rings <strong>of</strong> adjacent polyanions<br />

interact in an orthogonal fashion. Nevertheless, the respective Na-O-W bond lengths<br />

74


(Na1-O12T = 2.401(14) Å; W12-O12T = 1.698(13) Å) indicate that this intermolecular<br />

interaction is rather weak <strong>and</strong> mostly a result <strong>of</strong> crystal packing. It can be expected that<br />

redissolution <strong>of</strong> NH4-8 leads to a complete breakdown <strong>of</strong> the entire lattice, resulting in<br />

the presence <strong>of</strong> individual polyanions 8 in solution. In order to verify this assumption we<br />

decided to perform multinuclear NMR spectroscopy.<br />

Solution NMR<br />

Polyanion 8 is diamagnetic <strong>and</strong> contains four spin 1/2 nuclei ( 183 W, 119 Sn, 13 C, 1 H) <strong>and</strong><br />

therefore represents a good c<strong>and</strong>idate for solution NMR studies at room temperature.<br />

The 1 H <strong>and</strong> 13 C-NMR spectra <strong>of</strong> 8 are consistent with two equivalent phenyl groups.<br />

Especially 183 W-NMR is a very sensitive technique which allows to verify if the solid state<br />

structure <strong>of</strong> a polyoxotungstate is preserved in solution. However, the low natural abundance<br />

<strong>of</strong> the 183 W-nucleus requires preparation <strong>of</strong> very concentrated solutions. In order<br />

to accomplish this we synthesized 8 as described in the Experimental section but with<br />

a four times higher concentration. During the reaction solid LiClO 4 was added in order<br />

to prevent precipitation <strong>of</strong> 8. The solid KClO 4 was filtered <strong>of</strong>f before running the NMR<br />

measurements, which means that essentially all potassium ions were removed from the<br />

solution. Therefore the only remaining cations in the solution <strong>of</strong> 8 studied by NMR were<br />

NH 4+ <strong>and</strong> Li + . For 8 a six line pattern with relative intensities 4:4:4:4:2:1 is expected<br />

<strong>and</strong> indeed this is what we observed, confirming the C 2v symmetry <strong>of</strong> 8 (see Figure 3.33).<br />

This conclusion is based on the assumption that the two phenyl groups can rotate freely<br />

in solution, which is most likely the case. The smallest peak in the 183 W-NMR spectrum<br />

( -196.9 ppm) is somewhat hard to be identified, but we performed several experiments<br />

to confirm our assignment. The 119 Sn-NMR spectrum <strong>of</strong> 1 is expected to show a single<br />

peak, if the two Sn atoms are equivalent. Indeed a single resonance at -432.5 ppm is<br />

observed, but in addition we see a pair <strong>of</strong> fairly intense satellites which most likely result<br />

from coupling <strong>of</strong> a 119 Sn nucleus to an adjacent 183 W nucleus (see Figure 3.34). Pope et<br />

al. observed a singlet in 119 Sn-NMR for [(C 6 H 5 Sn) 3 Na 3 (H 2 O) 6 (α-SbW 9 O 33 ) 2 ] 6− <strong>of</strong> almost<br />

identical chemical shift (-417.8 ppm) to 8 <strong>and</strong> a pair <strong>of</strong> satellites with similar intensity.[47]<br />

In his polyanion the tin atoms are in a square-pyramidal coordination geometry, in com-<br />

75


Fig. 3.35: 183 W NMR spectra <strong>of</strong> compound 8<br />

Fig. 3.36: 119 Sn NMR spectra <strong>of</strong> compound 8<br />

plete analogy to 8. Although single-crystal X-ray diffraction <strong>of</strong> NH4-8 clearly indicated<br />

the presence <strong>of</strong> a sodium ion in the central belt <strong>of</strong> the title polyanion, it must be remembered<br />

that we synthesized 8 in the absence <strong>of</strong> sodium ions (see Experimental section).<br />

Furthermore we eliminated the potassium ions by precipitation as KClO 4 as described<br />

above. Sodium ions were only added after the synthesis <strong>of</strong> 8 in order to obtain better<br />

quality crystals. Based on our observation that in the solid state structure a sodium ion<br />

is located in the central belt <strong>of</strong> 8, we decided to investigate by 119 Sn-NMR if sodium ions<br />

also play an important role in solution. We discovered that in the absence <strong>of</strong> sodium ions<br />

only one signal is observed at -432.5 ppm together with a pair <strong>of</strong> satellites (see Figure<br />

3.36. Addition <strong>of</strong> solid NaCl to this solution or alternatively the presence <strong>of</strong> sodium ions<br />

already during the synthesis <strong>of</strong> 8 resulted in both cases in the appearance <strong>of</strong> a second<br />

76


119 Sn-NMR signal at -434.9 ppm with the expected pair <strong>of</strong> satellites (2JSn-W = 100 Hz).<br />

This peak increased with the concentration <strong>of</strong> sodium ions in solution, but never reached<br />

the same intensity as the -432.5 ppm peak. Our conclusion is that the peak at -434.9 ppm<br />

almost certainly represents 8 with a sodium ion in the vacancy. We could not identify any<br />

two-bond 119 Sn- 23 Na coupling in this peak pattern, most likely for the following reasons:<br />

(a) the 23 Na nucleus is quadrupolar (I = 3/2) <strong>and</strong> the satellite signal would be a quartet,<br />

(b) the Na-O bond lengths are rather long (2.44 - 2.46(2) Å) so that any coupling would<br />

be expected to be rather weak. On the other h<strong>and</strong> we believe that the peak at -432.5 ppm<br />

indicates a vacancy in the title polyanion 8 in-between the two tin atoms. The fact that<br />

the 119 Sn-NMR signal for the sodium-substituted species at -434.9 ppm is always smaller<br />

than the signal at -432.5 ppm (even for high Na + concentrations) indicates that the latter<br />

represents almost certainly a species with a ‘vacant’ belt. The two freely rotating phenyl<br />

groups probably exhibit a steric effect which makes incorporation <strong>of</strong> an alkali cation in the<br />

belt <strong>of</strong> 8 more difficult. We also performed 23 Na-NMR on solutions <strong>of</strong> 8 with added sodium<br />

chloride. For such solutions a single peak is observed at around -1.5 ppm, which is very<br />

similar to the chemical shift <strong>of</strong> our 1 M NaCl reference solution. No satellite peaks could<br />

be observed, but we noticed that the peak <strong>of</strong> the polyanion solution is significantly broader<br />

than that <strong>of</strong> the NaCl reference solution. This could be a result <strong>of</strong> exchange-broadening,<br />

which appears to be fast on the NMR timescale. The observed Sn-O-W coupling constant<br />

for 8 (2JSn-W = 96 Hz) is even larger than the largest values reported by Pope et al. for<br />

[(C 6 H 5 Sn) 3 P 2 W 15 O 59 ] 9− ( 2 JSn-W = 78 Hz), [(C 4 H 9 Sn) 3 P 2 W 15 O 59 ] 9− ( 2 JSn-W = 49 Hz),<br />

[(C 6 H 5 SnOH) 3 ( α-SiW 9 O 34 ) 2 ] 14− ( 2 JSn-W = 35 Hz) <strong>and</strong> [(C 6 H 5 SnOH) 3 (PW 9 O 34 ) 2 ] 12−<br />

( 2 JSn-W = 33 Hz). Pope et al. have identified that an increase in the Sn-O-W bond angle<br />

is accompanied by an increasing Sn-O-W coupling constant for [(C 6 H 5 Sn) 3 P 2 W 15 O 59 ] 9−<br />

( 2 JSn-W = 78 Hz, Sn-O-W = 147-158 ◦ ) <strong>and</strong> [(C 6 H 5 SnOH) 3 (PW 9 O 34 ) 2 ] 12− ( 2 JSn-W<br />

= 33 Hz, Sn-O-W = 139-142 ◦ ).[44] However, this bond angle argument alone does<br />

not explain the very large coupling constant observed for 8, as the Sn-O-W bond angle<br />

ranges from 135.0-139.6 ◦ . Most likely the coordination geometries <strong>of</strong> the tin atoms<br />

play also an important role. In all <strong>of</strong> Popes polyanions above (with the exception <strong>of</strong><br />

[(C 6 H 5 Sn) 3 Na 3 (H 2 O) 6 (α-SbW 9 O 33 ) 2 ]6−) the Sn atoms are six-coordinated (octahedral),<br />

77


whereas they are five-coordinated (square-pyramidal) in 8. Perhaps the overall charge <strong>of</strong><br />

the polyanion also affects the Sn-O-W coupling constant. It can be noticed that 8 has<br />

the smallest charge (-8) <strong>and</strong> the largest Sn-O-W coupling constant ( 2 JSn-W = 96 Hz).<br />

For Popes polyanions above the species with large negative charges have smaller coupling<br />

constants than those with smaller charges (e.g. [(C 6 H 5 Sn) 3 P 2 W 15 O 59 ] 9− , 2JSn-W = 78<br />

Hz vs. [(C 6 H 5 SnOH) 3 (α -SiW 9 O 34 ) 2 ] 14− , 2 JSn-W = 35 Hz). Nevertheless, NMR data <strong>of</strong><br />

more polyanions containing square pyramidal organotin groups are needed in order to be<br />

able to draw definitive conclusions. The 183 W-, 119 Sn-, 13 C- <strong>and</strong> 1 H-NMR spectra <strong>of</strong> 8 remain<br />

the same even if the same solution is measured again after several weeks, indicating<br />

that 8 does not undergo structural transformations in aqueous acidic medium.<br />

3.6.3 Conclusions<br />

We have synthesized the bis-phenyltin substituted, lone pair containing tungstoarsenate<br />

[(C 6 H 5 Sn) 2 As 2 W 19 O 67 (H 2 O)] 8− 8. Polyanion 6 was characterized by several solution<br />

(multinuclear NMR spectroscopy) <strong>and</strong> solid state (FTIR spectroscopy, elemental analysis,<br />

single-crystal XRD) techniques. This polyanion adds a new member to the class <strong>of</strong><br />

monophenyltin substituted polyoyotungstates. Our work reemphasizes (a) the facile synthesis<br />

<strong>of</strong> polyoxoanions by one-pot reactions, (b) the structural variety <strong>of</strong> polyoxoanion<br />

chemistry, (c) the stability <strong>of</strong> polyoxoanions in solution, (d) the complementary nature<br />

<strong>of</strong> XRD <strong>and</strong> multinuclear NMR, (e) easy incorporation <strong>of</strong> monoorganotin fragments in<br />

polyoxotungstate precursors, (f) the strong attachment <strong>of</strong> organotin fragments to polyoxoanions.<br />

Nevertheless the work <strong>of</strong> Pope <strong>and</strong> also our work shows that interaction <strong>of</strong><br />

monoorganotin groups with lacunary polyoxoanions almost inevitably leads to monomeric<br />

or dimeric products. Currently we investigate the reactivity <strong>of</strong> diorganotin groups with<br />

lacunary polyoxotungstates in order to synthesize discrete polyoxometalates with fundamentally<br />

novel architectures <strong>of</strong> large size.<br />

78


3.7 Conclusions <strong>and</strong> Outlook<br />

3.7.1 Conclusion<br />

The primary objective <strong>of</strong> this work was to synthesize novel discrete polyoxometalates<br />

(POMs) functionalized by diorganotin groups. Diorganotin electrophiles were reacted with<br />

a variety <strong>of</strong> lacunary POM precursors (lig<strong>and</strong>s) in aqueous media over a wide range <strong>of</strong> synthetic<br />

conditions. Novel discrete POMs species as monomers, dimers, trimers, tetramers<br />

<strong>and</strong> dodecamers were synthesized <strong>and</strong> structurally characterized by FTIR, single crystal<br />

X-ray diffraction, multinuclear NMR, electrochemistry (Dr. B. Keita, <strong>and</strong> Pr<strong>of</strong>. L. Nadjo,<br />

(Université Paris-Sud, Orsay Cedex, France), scanning tunneling microscopy ( Pr<strong>of</strong>. P.<br />

Müller <strong>and</strong> his coworkers Mr. M. S. Alam, V. Dremov, Physikalisches Institut III, (Universität<br />

Erlangen-Nürnberg, Germany). From our observations it can be concluded that:<br />

(i)The dimethyltin unit (CH 3 ) 2 Sn 2+ is a reactive electrophile <strong>and</strong> interacts readily with<br />

lacunary polyoxotungstates, such as the trilacunary [P V W 9 O 34 ] 9− <strong>and</strong> [As V W 9 O 34 ] 9−<br />

or the trilacunary, lone pair containing [As III W 9 O 33 ] 9− <strong>and</strong> [Sb III W 9 O 33 ] 9− . (ii) The<br />

dimethyltin unit (CH 3 ) 2 Sn 2+ is grafted exclusively at lacunary sites <strong>of</strong> polyanions, predominantly<br />

via two Sn-O(W) bonds. (iii) The methyl groups <strong>of</strong> the dimethyltin-containing<br />

POMs are trans to each other <strong>and</strong> therefore the (CH 3 ) 2 Sn 2+ group is an effective linker <strong>of</strong><br />

POM precursors, resulting in large, open cage type structures. (iv) In most dimethyltincontaining<br />

POMs the tin atom exihibits coordination number six (octahedral) but in the<br />

tetrameric polyanion 5 ([{Sn(CH 3 ) 2 (H 2 O)} 2 {Sn(CH 3 ) 2 }As 3 (α-AsW 9 O 33 ) 4 ] 12− ) coordination<br />

number seven (pentagonal bipyramidal) was observed. (v) The Sn-C bond lengths<br />

in in dimethyltin-containing POMs range from 1.9 - 2.1 Å. (vi) The C-Sn-C bond angles<br />

in dimethyltin-containing POMs range from 150 - 175 ◦ . (vii) The dimethyltin-containing<br />

POM structures are open <strong>and</strong> exhibit central cavities <strong>and</strong>/or surface pockets which may<br />

allow for host-guest chemistry.. (viii) The diamagnetic nature <strong>of</strong> dimethyltin-containing<br />

POMs allows for multinuclear solution NMR (e.g. 119 Sn, 183 W, 13 C, 1 H). (ix) In a collaboration<br />

with the group <strong>of</strong> Pr<strong>of</strong>. P. Müller (Physikalisches Institut III, Universität<br />

Erlangen-Nürnberg, Germany) the dodecameric 6 ([{Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-<br />

PW 9 O 34 ) 12 ] 36− ) was deposited on highly oriented pyrolitic graphite (HOPG) <strong>and</strong> then<br />

79


investigated by scanning tunneling microscopy (STM) Individual polyanions could be visualized.<br />

(x) In a collaboration with the group <strong>of</strong> Pr<strong>of</strong>. L. Nadjo <strong>and</strong> Dr. B. Keita (Université<br />

Paris-Sud, Orsay Cedex, France) detailed electrochemistry studies were carried<br />

out on our dimethyltin-containing POMs. This work usually involved cyclic voltammetry<br />

studies. (xi) Reaction <strong>of</strong> (C 6 H 5 ) 2 Sn 2+ with lacunary tungstoarsenates(III) in aqueous<br />

acidic medium <strong>and</strong> heating resulted in partial hydrolysis <strong>of</strong> the former. Loss <strong>of</strong> one phenyl<br />

group resulted in formation <strong>of</strong> monophenyltin-POMs.<br />

3.7.2 Outlook<br />

The reactivity <strong>of</strong> (C 6 H 5 ) 2 Sn 2+ can be further investigated by using other lacunary POMs<br />

including borotungstates (e.g.[B 3 W 39 O 132 ] 21− ), arsenotungstates (e.g. [As 2 W 19 O 67 (H 2 O)] 14− ,<br />

[As 2 W 20 O 68 (H 2 O)] 10− ), silicotungstates (e.g. [γ-SiW 10 O 36 ] 8− , [SiW 9 O 34 ] 10− ), germanotungstates<br />

(e.g. [γ-GeW 10 O 36 ] 8− , [GeW 9 O 34 ] 10− ) <strong>and</strong> perhaps also mixed polyoxotungstate<br />

lig<strong>and</strong>s. In is also <strong>of</strong> interest to study the reactivity <strong>of</strong> the diethyltin group (C 2 H 5 ) 2 Sn 2+<br />

with a variety <strong>of</strong> POM precursors. Such POM derivatives might allow for further derivatization<br />

<strong>of</strong> the alkyl chain leading to peptides, thiols etc. The interaction <strong>of</strong> these species<br />

with enzymes <strong>and</strong> other biomolecules can also be studied <strong>and</strong> might allow for a better<br />

underst<strong>and</strong>ing <strong>of</strong> the antitumor/-viral activities <strong>of</strong> this class <strong>of</strong> compounds. Moreover,<br />

such derivatized POMs may also be used for stabilizing nanoparticles. Finally, it can also<br />

be envisaged to extend the diorganotin-POM work to the lighter homologue diorganogermanium<br />

R 2 Ge 2+ 80


Part-III<br />

Titanium Containing POMs


3.8 The di-titanium substituted, lone pair containing<br />

tungstoarsenate:<br />

{(TiOH) 2 WO(H 2 O)(B-α-AsW 9 O 33 ) 2 } 8−<br />

3.8.1 Introduction<br />

Polyoxometalates (POMs) are an important class <strong>of</strong> inorganic compounds <strong>and</strong> they exhibit<br />

a diverse compositional range <strong>and</strong> significant structural versatility [1, 134]. POMs<br />

are usually composed <strong>of</strong> early transition metal MO 6 (M = W 6+ , Mo 6+ etc.) octahedra<br />

<strong>and</strong> main group XO 4 (X = P, Si etc.) tetrahedra. The most famous POMs are probably<br />

the Keggin (e.g. (PW 12 O 40 ) 3− ) <strong>and</strong> the Wells-Dawson (e.g. (P 2 W 18 O 62 ) 6− ) ions. Nevertheless,<br />

also lone pair containing main group elements (e.g. As III ) can act as hetero<br />

groups. There are numerous uses <strong>of</strong> POMs utilizing their specific molecular composition,<br />

size, shape, charge density, redox potentials, acidity <strong>and</strong> solubility characteristics, which<br />

span a wide range <strong>of</strong> applications including homogeneous <strong>and</strong> heterogeneous catalysts,<br />

electro-catalysts, coatings, medicinal agents, pigments, recording materials, toners, precursors<br />

for oxide films, sensors <strong>and</strong> nano/biotechnology [2–5, 135]. POMs have received<br />

particular attention in environmentally benign catalytic processes <strong>and</strong> in antiviral <strong>and</strong><br />

antitumoral chemotherapy [2–5, 135]. Nevertheless the structure/composition-activity<br />

relationship <strong>of</strong> POMs in these <strong>and</strong> other applications is not yet fully understood <strong>and</strong><br />

therefore the synthesis <strong>of</strong> new types <strong>of</strong> such polyanions remains an important research<br />

objective. Catalytic activity (e.g. activation <strong>of</strong> O 2 <strong>and</strong> H 2 O 2 ) combined with high thermal<br />

stability have led to industrial applications <strong>of</strong> these species, mostly as heterogeneous<br />

catalysts in the oxidation <strong>of</strong> organic substrates (e.g. Wacker process) [2–5, 135]. Titanium<br />

substituted polyoxoanions are <strong>of</strong> interest for structural reasons as well as for their<br />

universal catalytic properties. Polyoxometalates substituted with early transition metal<br />

d 0 ions such as vanadium (V) <strong>and</strong> niobium (V) in a fully oxidized state increases the<br />

basicity <strong>of</strong> the polyanions, which may enable the binding <strong>of</strong> organometallic fragments to<br />

specific sites <strong>of</strong> the anion, or even the functionalization <strong>of</strong> Ir nanoclusters to the polyanions,<br />

which is well documented [136–145]. Substitution <strong>of</strong> V V <strong>and</strong> Nb V by Ti IV ions<br />

82


should even lead to a more pronounced effect. Indeed it has been observed that titanium<br />

atoms incorporated in polyoxoanions are reactive sites with a strong tendency to form<br />

dimers through Ti-O-Ti bonds as seen in [(α-Ti 3 PW 9 O 38.5 ) 2 ] 12− , [(x-Ti 3 SiW 9 O 38.5 ) 2 ] 14−<br />

(x =α,β), [(α-Ti 3 GeW 9 O 38.5 ) 2 ] 14− , [(α-1,2-Ti 2 PW 10 O 39 ) 2 ] 10− , <strong>and</strong> [(β-Ti 2 SiW 10 O 39 ) 4 ] 24−<br />

[146–151]. Nevertheless, two monomeric species have also been structurally characterized<br />

[152–155]. Interestingly all these monomeric, dimeric <strong>and</strong> cyclic tetrameric species are<br />

polyoxotungstates based on the Keggin structure as a basic building framework. Some <strong>of</strong><br />

the above species have shown interesting catalytic (e.g. photocatalysis) as well as medicinal<br />

activity (e.g. antiviral) [156–162]. The search for novel transition-metal substituted<br />

polyanions is predominantly driven by their catalytic properties. Our group have been<br />

interested in Ti-substituted polyoxometalates for a few years mainly because <strong>of</strong> their attractive<br />

properties like catalysis <strong>and</strong> photocatalysis. Kortz et al. reported that interaction<br />

<strong>of</strong> titanium(IV) with the Wells-Dawson ion [P 2 W 15 O 56 ] 12− in aqueous solution can lead<br />

to a variety <strong>of</strong> products with unexpected, supramolecular structures [163]. Soon after<br />

similar work was also reported by Nomiya et al. but they were unsuccessful to obtain<br />

single crystal suitable for X-ray diffraction [164, 165]. Here we report on the structural<br />

characterization <strong>of</strong> titanium(IV)-substituted polyoxoanions <strong>of</strong> the lone pair containing<br />

dilacunary tungstoarsenate(III) [As 2 W 19 O 67 (H 2 O)] 14− .<br />

3.8.2 Experimental<br />

The precursor K 14 [As 2 W 19 O 67 (H 2 O)] was synthesized according to the published procedure<br />

<strong>of</strong> Kortz et al. <strong>and</strong> the purity was confirmed by infrared spectroscopy [33]. All other<br />

reagents were used as purchased without further purification. Cs 8 [(TiOH) 2 As 2 W 19 O 67 (H 2 O)]<br />

·10.5 H 2 O (Cs-9). The title compound was synthesized by dissolving 0.141 g (0.88 mmols)<br />

<strong>of</strong> TiO(SO 4 ) in 40 mL H 2 O, followed by addition <strong>of</strong> 2.10 g (0.40 mmols) K 14 [As 2 W 19 O 67 (H 2 O)].<br />

This solution (pH 2.0) was heated to ∼80 ◦ C for 1 h <strong>and</strong> then cooled to room temperature.<br />

The solution was filtered <strong>and</strong> a few drops <strong>of</strong> 0.1 M CsCl solution were added <strong>and</strong> then the<br />

solution was allowed to evaporate in an open vial at room temperature. A light yellow<br />

crystalline product started to appear after a week or two. Evaporation was continued<br />

until the solvent approached the solid product (yield 1.43 g, 63.8 %). FTIR spectra for<br />

83


Fig. 3.37: Ball/stick (left), polyhedral (right) representations <strong>of</strong> 9<br />

Cs 8 [(TiOH) 2 As 2 W 19 O 67 (H 2 O)]·10.5 H 2 O: 963(s), 898(s), 761(s), 712(s), 589(sh), 484(w),<br />

448(w) cm −1 . The FTIR spectrum was recorded on a Nicolet Avatar FTIR spectrophotometer<br />

in a KBr pellet.<br />

X-ray Crystallography<br />

A crystal <strong>of</strong> compound Cs-9 was mounted on a glass fiber for indexing <strong>and</strong> intensity data<br />

collection at 173 K on a Bruker D8 SMART APEX CCD single-crystal diffractometer<br />

using Mo K α radiation (λ = 0.71073 Å). Direct methods were used to solve the structure<br />

<strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining atoms were found from<br />

successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong> polarization corrections<br />

were applied <strong>and</strong> an absorption correction was performed using the SADABS program<br />

[81]. Crystallographic data are summarized in Table 3.8<br />

84


Table 3.8: Crystal Data <strong>and</strong> Structure Refinement for compound Cs-9<br />

Empirical formula As 2 Cl 2 Cs 8.5 O 78.5 Ti 2 W 19<br />

fw 6195.42<br />

space group P 2 1 /m (11)<br />

a (Å) 12.776(19)<br />

b (Å) 19.425(3)<br />

c (Å) 18.149(3)<br />

β ( ◦ ) 110.23(3)<br />

vol (Å 3 ) 4226(5)<br />

Z 2<br />

temp. ( ◦ C) -100<br />

wavelength (Å) 0.71073<br />

d calc (mg m −3 ) 4.868<br />

abs. coeff. (mm −1 ) 30.466<br />

R [I > 2 σ(I)] a 0.081<br />

R w (all data) b 0.092<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

3.8.3 Results <strong>and</strong> discussion<br />

Synthesis <strong>and</strong> Structure<br />

The novel di-titanium substituted, dilacunary tungstoarsenate(III) [(TiOH) 2 As 2 W 19 O 67 (H 2 O)] 8−<br />

9 consists <strong>of</strong> two lacunary B-α-[AsW 9 O 33 ] 9− Keggin fragments linked via two Ti(IV)<br />

atoms <strong>and</strong> a {WO(H 2 O)} 4+ moiety leading to a structure with nominal C 2 v symmetry<br />

(see Figure 3.37). Alternatively 9 can be described as a dilacunary [As 2 W 19 O 67 (H 2 O)] 14−<br />

fragment which has taken up two Ti(IV) units. It can be noticed that the titanium atoms<br />

are situated well above the plane <strong>of</strong> the four equatorial oxo lig<strong>and</strong>s. Therefore the titanium<br />

atoms are displaced towards the exterior <strong>of</strong> the equator <strong>of</strong> 9. This resembles the<br />

displacement <strong>of</strong> all tungsten sites <strong>of</strong> 9 towards the external, terminal oxo groups. On<br />

the other h<strong>and</strong>, the unique tungsten atom in the central belt <strong>of</strong> 9 is displaced towards<br />

the interior <strong>of</strong> the equator <strong>of</strong> 9. Bond-valence-sum (BVS) calculations for 9 indicated<br />

that no oxygen <strong>of</strong> the two (AsW 9 O 33 ) caps is protonated [85]. However, the central tungsten<br />

atom has two trans related lig<strong>and</strong>s which are a water molecule <strong>and</strong> an oxo group.<br />

The latter is inside the central cavity <strong>of</strong> 9 whereas the former is on the outside. This<br />

is in complete agreement with the single-crystal XRD data <strong>of</strong> related structures that<br />

contain one or three central tungsten atoms linking two ( α-AsW 9 O 33 ) fragments (e.g.<br />

85


[As 2 W 19 O 67 (H 2 O)] 14− , [As 2 W 21 O 69 (H 2 O)] 6− ) [33, 35, 129]. The title polyanion 9 was<br />

synthesized rationally <strong>and</strong> in a one-pot reaction by interaction <strong>of</strong> titanium oxosulphate<br />

TiO(SO 4 ) with K 14 [As 2 W 19 O 67 (H 2 O)] in aqueous, acidic medium (pH 2.0). However, we<br />

obtained 9 for the first time accidentally during our efforts to discover novel titanium<br />

substituted s<strong>and</strong>wich type species. Interaction <strong>of</strong> TiO(SO 4 ) with Na 9 [B-AsW 9 O 33 ] in<br />

a molar ratio <strong>of</strong> 4:1 in aqueous medium at pH 4.0 resulted in 9. Then we decided to<br />

reproduce this compound via a more rational synthetic procedure using the dilacunary<br />

polyoxotungstate precursor [As 2 W 19 O 67 (H 2 O)] 14− <strong>and</strong> TiO(SO 4 ) (see Experimental Section).<br />

The tungstoarsenate [As 2 W 19 O 67 (H 2 O)] 14− was synthesized for the first time about<br />

30 years ago by Tourné et al. <strong>and</strong> recently Kortz et al. confirmed the proposed structure<br />

by X-ray diffraction [33, 128, 129]. Interestingly, to date only a few polyoxoanions have<br />

been synthesized using [As 2 W 19 O 67 (H 2 O)] 14− as a precursor [33, 130, 131]. Polyanion<br />

9 represents a novel member in the class <strong>of</strong> titanium substituted polyoxotungstates in<br />

general <strong>and</strong> the subclass <strong>of</strong> tungstoarsenates(III) in particular. Recently, our group reacted<br />

(C 6 H 5 ) 2 SnCl 2 with [As III 2W 19 O 67 (H 2 O)] 14− in aqueous medium at pH 1.6, resulting<br />

in novel bis-phenyltin substituted species, [Na(H 2 O)(C 6 H 5 Sn) 2 As III 2W 19 O 67 (H 2 O)] 7−<br />

[166]. Most recently our group interacted Pd 2+ ions with the lacunary Keggin precursors<br />

A-α-[SiW 9 O 34 ] 10− in buffer solutions (pH 4.9) resulting in novel structurally characterized<br />

palladium substituted species. [Pd 2 WO(H 2 O)(A-α-SiW 9 O 34 ) 2 ] [167]. Therefore, the<br />

mechanism <strong>of</strong> formation <strong>of</strong> 9 involves insertion <strong>and</strong> isomerization (B-alpha-[AsW 9 O 33 ] 9−<br />

→ [As III 2W 19 O 67 (H 2 O)] 14− ). Most interestingly, the central As III atoms in<br />

[As III 2W 19 O 67 (H 2 O)] 14− have lone pairs <strong>of</strong> electrons but in the later case the central hetero<br />

atoms Si IV do not contain lone pairs <strong>of</strong> electrons. So, this tungsten-oxo fragment for<br />

silicotungstates [Si 2 W 19 O 69 (H 2 O)] 16− has never been observed before <strong>and</strong> therefore it represents<br />

the first 2:19 series <strong>of</strong> silicotungstate containing central hetero atoms without lone<br />

pairs <strong>of</strong> electrons. But such 2:19 series <strong>of</strong> phosphotungstates are known for two decades<br />

where the heteroatom do not contain lone pair <strong>of</strong> electrons.[80] Currently, our group is<br />

engaged in isolating the novel silicotungstate assembly [Si 2 W 19 O 69 (H 2 O)] 16− in a more<br />

rational way. The mechanism <strong>of</strong> formation <strong>of</strong> polyoxometalates is not well understood<br />

<strong>and</strong> commonly described as self-assembly. Therefore, the design <strong>of</strong> novel polyoxometa-<br />

86


lates remains a challenge for synthetic chemists. This is analogous to the di-transition<br />

metal substituted derivatives <strong>of</strong> this structural type, [M 2 (H 2 O) 2 WO(H 2 O)Na 3 (H 2 O) 6 (α<br />

-AsW 9 O 33 ) 2 ] 7− (M = Co 2+ , Zn 2+ ) [39]. The two titanium cations are bound to two “terminal”<br />

oxygen atoms <strong>of</strong> pairs <strong>of</strong> edge-shared WO 6 octahedra <strong>of</strong> each B-α-[AsW 9 O 33 ] 9− anion<br />

<strong>of</strong> the dimer. Both the titanium atoms display square pyramidal geometry. Bond-valencesum<br />

(BVS) calculations show that both titanium are coordinated axially to -OH lig<strong>and</strong>s.<br />

The Ti-O bond lengths (1.964-1.985 Å) <strong>and</strong> the O-Ti-O angles (84-85 ◦ , 159-159.5 ◦ ) are<br />

very close to that <strong>of</strong> nominal values. Ti1···Ti2, Ti1···W19 <strong>and</strong> Ti2···W19 distances in 9<br />

are 5.1, 4.6 <strong>and</strong> 4.6 Å, respectively. These values indicate that two Ti <strong>and</strong> a W in the<br />

central belt <strong>of</strong> polyanion 9 form an approximate isosceles triangle. The -8 charge <strong>of</strong> the<br />

title polyanion is balanced by eight cesium ions which surround the polyanion. The exact<br />

positions <strong>of</strong> all the cesium ions could not be identified by X-ray diffraction. However, the<br />

result <strong>of</strong> elemental analysis is in complete agreement with the formula <strong>of</strong> Cs-9.<br />

3.8.4 Conclusions<br />

We have synthesized the di-titanium substituted, lone pair containing tungstoarsenate<br />

[(TiOH) 2 As 2 W 19 O 67 (H 2 O)] 8− 9. Polyanion 9 was characterized in the solid state by<br />

different analytical techniques like FTIR spectroscopy, elemental analysis, single-crystal<br />

XRD. This polyanion adds a new member to the class <strong>of</strong> titanium substituted polyoyotungstates.<br />

Our work reemphasizes (a) the facile synthesis <strong>of</strong> polyoxoanions by one-pot<br />

reactions, (b) the structural variety <strong>of</strong> polyoxoanion chemistry, (c) the first example where<br />

the polyoxometalates dimerizes without formation <strong>of</strong> Ti-O-Ti bond in low pH. (d) adds<br />

to the subclass where Keggin fragment is used as building blocks (e) easy incorporation<br />

<strong>of</strong> titanium(IV) in lone pair containing tungstatoarsenate precursors. Currently we<br />

are investigating the electro- <strong>and</strong> photochemical,electrocatalytic, <strong>and</strong> oxidation catalysis<br />

properties <strong>of</strong> 9 <strong>and</strong> this work is under investigation.<br />

(Manuscript in preparation)<br />

87


3.9 The cyclic trimeric-titanium substituted,<br />

tungstophosphate: [{Ti 3 O 4 (A-α-PW 9 O 34 )} 3 (PO 4 )] 13−<br />

3.9.1 Experimental<br />

The precursor K 14 [P 2 W 19 O 69 (H 2 O)] was synthesized according to the published procedure<br />

<strong>of</strong> Tourné et al. <strong>and</strong> the purity was confirmed by infrared spectroscopy [80]. All<br />

other reagents were used as purchased without further purification. Rb 9 K 4 [{(Ti 3 O 4 )(A-α-<br />

PW 9 O 34 )} 3 (PO 4 )]·18 H 2 O (RbK-10). The title compound was synthesized by dissolving<br />

0.141 g (0.88 mmols) <strong>of</strong> TiO(SO 4 ) in 40 mL potassium acetate buffer followed by addition<br />

<strong>of</strong> 2.26 g (0.40 mmols) K 14 [P 2 W 19 O 69 (H 2 O)]. This solution (pH 4.8) was heated to ∼80<br />

◦ C for 1 h <strong>and</strong> then cooled to room temperature. The solution was filtered <strong>and</strong> a few<br />

drops <strong>of</strong> 0.1 M RbCl solution were added <strong>and</strong> then the solution was allowed to evaporate<br />

in an open vial at room temperature. A white crystalline product started to appear after<br />

a week or two. Evaporation was continued until the solvent approached the solid product<br />

(yield 1.43 g, 63.8 %). FTIR spectra for Rb 9 K 4 [{(Ti 3 O 4 )(A-α-PW 9 O 34 )} 3 (PO 4 )]·18 H 2 O<br />

: 1163(sh), 1064(s), 1030(w), 959(s), 883(w), 788(s), 686(s), 588(w), 511(w) cm −1 . The<br />

FTIR spectrum was recorded on a Nicolet Avatar FTIR spectrophotometer in a KBr<br />

pellet.<br />

Fig. 3.38: FTIR spectra <strong>of</strong> compound RbK-10(red) <strong>and</strong> K 14 [P 2 W 19 O 69 (H 2 O)](blue)<br />

88


X-ray Crystallography<br />

A crystal <strong>of</strong> compound RbK-10 was mounted on a glass fiber for indexing <strong>and</strong> intensity<br />

data collection at 173 K on a Bruker D8 SMART APEX CCD single-crystal diffractometer<br />

using Mo K α radiation (λ = 0.71073 Å). Direct methods were used to solve the structure<br />

<strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining atoms were found from<br />

successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong> polarization corrections<br />

were applied <strong>and</strong> an absorption correction was performed using the SADABS program<br />

[81]. Crystallographic data are summarized in Table 3.9<br />

Fig. 3.39: Ball/stick <strong>and</strong> polyhedral representation <strong>of</strong> polyanion 10, color codes: blue polyhedra (W),<br />

yellow balls (Ti), red balls (O) <strong>and</strong> pink polyhedra (PO4)<br />

Results <strong>and</strong> discussion<br />

The novel polyanion 10 is composed <strong>of</strong> three [A-α-PW 9 O 34 ] units <strong>and</strong> nine TiO 6 octahedra.<br />

The nine TiO 6 octahedra are connected to each other by Ti-O-Ti bridges. Three<br />

TiO 6 octahedra <strong>of</strong> each trilacunary [A-α-PW 9 O 34 ] unit fills the lacuna to form a complete<br />

Keggin cluster. Of the three TiO 6 octahedra two are connected to Ti-centers <strong>of</strong> adjacent<br />

polyanions by Ti-O-Ti bridges, the unique TiO 6 octahedra in each Keggin subunit is connected<br />

to a phosphate lig<strong>and</strong>, which act as a capping tris monodentae fragment, leading<br />

to a discrete cyclic trimeric cluster. The polyanion has a nominal symmetry <strong>of</strong> C 3v (See<br />

89


Table 3.9: Crystal Data <strong>and</strong> Structure Refinement for compound RbK-10<br />

Empirical formula K 9 O 136 P 4 Rb 4 Ti 9 W 27<br />

fw 8388.71<br />

space group R 3m (160)<br />

a (Å) 29.7444(7)<br />

c (Å) 13.6254(9)<br />

volume (Å 3 ) 10439.8(8)<br />

Z 3<br />

temp. ( ◦ C) -100<br />

wavelength (Å) 0.71073<br />

dcalc (mg m −3 ) 4.003<br />

abs. coeff. (mm −1 ) 24.508<br />

R [I > 2 σ(I)] a 0.063<br />

R w (all data) b 0.079<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

Figure. 3.36)<br />

Further studies are on progress.<br />

3.10 Structural control on the nanomolecular scale:<br />

Self- assembly <strong>of</strong> the polyoxotungstate wheel<br />

({β-Ti 2 SiW 10 O 39 } 4 ) 24−<br />

3.10.1 Introduction<br />

The catalytic activity (e.g. activation <strong>of</strong> O 2 <strong>and</strong> H 2 O 2 )combined with high thermal stability<br />

have led to industrial applications <strong>of</strong> these species, mostly as heterogeneous catalysts<br />

in the oxidation <strong>of</strong> organic substrates (e.g. Wacker process) [3–5]. The redox-activity <strong>of</strong><br />

titanium in the oxidation state +4 is well known <strong>and</strong> has led to numerous catalytic studies<br />

using TiO 2 as photocatalyst [168–170]. To date only a few titanium(IV)-substituted<br />

polyoxoanions have been synthesized <strong>and</strong> most <strong>of</strong> them are <strong>of</strong> the Keggin-type. Structural<br />

characterization in the solid state indicates a strong tendency towards dimer formation<br />

via Ti-O-Ti bonds as seen in [(α-Ti 3 PW 9 O 38.5 ) 2 ] 12− , [(x-Ti 3 SiW 9 O 38.5 ) 2 ] 14− (x = α, β),<br />

[(α-Ti 3 GeW 9 O 38.5 ) 2 ] 14− <strong>and</strong> [(α-1,2-Ti 2 PW 10 O 39 ) 2 ] 10− [146, 148–151]. Nevertheless two<br />

90


monomeric species have also been structurally characterized [152–155]. Some <strong>of</strong> the above<br />

species have shown interesting catalytic (e.g. photocatalysis) as well as medicinal activity<br />

(e.g. antiviral) [156–158, 162]. Recently some dimeric <strong>and</strong> tetrameric Ti-substituted<br />

polyoxotungstates based on the Wells-Dawson fragment have been structurally characterized<br />

by Kortz et al. [163]. Soon afterwards Nomiya et al. reported on similar tetrameric<br />

species [164, 165]. Our group has been working extensively on the interaction <strong>of</strong> the<br />

metastable, dilacunary tungstosilicate [γ-SiW 10 O 36 ] 8− with low-valent, first row transition<br />

metals [171–173]. Now we decided to investigate the system Ti 4+ /[γ- SiW 10 O 36 ] 8− in<br />

some detail.<br />

3.10.2 Experimental<br />

Preparation <strong>of</strong> K 24 [β-Ti 2 SiW 10 O 394 ]·50H 2 O (K-11): A 2.23 g (0.75 mmol) sample <strong>of</strong><br />

K 8 [γ-SiW 10 O 36 ] was added with stirring to a solution <strong>of</strong> 0.26 g (1.65 mmol) TiO(SO 4 ) (E.<br />

Merck) in 40 mL H 2 O. The pH was adjusted to 2 by addition <strong>of</strong> 4 M HCl. This solution<br />

was heated to ∼80 ◦ C for 1 hour <strong>and</strong> then cooled to room temperature <strong>and</strong> filtered. Slow<br />

evaporation at room temperature resulted in a white, crystalline product after 2-3 weeks<br />

that was filtered <strong>of</strong>f <strong>and</strong> air-dried. Yield: 1.39 g (61%). FTIR spectroscopy: 1000(w),<br />

966(m), 913(s), 803(s), 657(m), 541(w), 516(w), 487(w), 467(w) cm −1 . Anal. Calcd for<br />

K 24 -7: K, 7.7; W, 60.4; Ti, 3.1; Si, 0.9. Found: K, 7.3; W, 61.6; Ti, 2.8; Si, 1.2. Elemental<br />

analysis was performed by Kanti Labs Ltd. in Mississauga, Canada.<br />

Interaction <strong>of</strong> solid TiO(SO 4 ) with [γ-SiW 10 O 36 ] 8− in the ratio 2:1 in aqueous acidic<br />

medium (pH 2) resulted in the novel, tetrameric [β-Ti 2 SiW 10 O 394 ] 24− (7), see Figures<br />

3.41.<br />

X-ray Crystallography<br />

A crystal <strong>of</strong> compound K-11 was mounted on a glass fiber for indexing <strong>and</strong> intensity data<br />

collection at 173 K on a Bruker D8 SMART APEX CCD single-crystal diffractometer<br />

using Mo K α radiation (λ = 0.71073 Å). Direct methods were used to solve the structure<br />

<strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining atoms were found from<br />

successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong> polarization corrections<br />

91


Fig. 3.40: FTIR spectra <strong>of</strong> compound K-11(red) <strong>and</strong> K 8 [γ-SiW 10 O 36 ](blue)<br />

Fig. 3.41: Polyhedral (left) <strong>and</strong> ball <strong>and</strong> stick (right) representations <strong>of</strong>({β-Ti 2 SiW 10 O 39 } 4 ) 24− .<br />

were applied <strong>and</strong> an absorption correction was performed using the SADABS program<br />

[81]. Crystallographic data are summarized in Table 3.10<br />

92


Table 3.10: Crystal Data <strong>and</strong> Structure Refinement for compound K-11<br />

Emperical formula H 100 K 24 O 206 Si 4 Ti 8 W 40<br />

fw 12184.9<br />

space group (No.) P 2 1 /n (14)<br />

a (Å) 12.5188(13)<br />

b (Å) 18.864(2)<br />

c (Å) 41.075(4)<br />

vol (Å 3 ) 9618.1(18)<br />

Z 2<br />

temp ( ◦ C) -100<br />

wavelength (Å) 0.71073<br />

d calcd (mg m −3 ) 4.20<br />

abs coeff. (mm −1 ) 24.631<br />

R [I > 2 σ(I)] a 0.104<br />

R w (all data) b 0.201<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

3.10.3 Results <strong>and</strong> discussion<br />

Polyanion 11 represents the first cyclic, tetrameric polyoxotungstate <strong>and</strong> it is also the<br />

largest Ti-substituted tungstosilicate known to date. Bond valence sum calculations<br />

(BVS) indicate that 11 is not protonated <strong>and</strong> therefore its charge must be 24 [85]. This<br />

is fully consistent with elemental analysis, which indicated the presence <strong>of</strong> 24 potassium<br />

ions. Due to disorder we could identify only 17 potassium counterions by X-ray diffraction.<br />

The central cavity <strong>of</strong> the wheel-shaped 11 is occupied by a potassium cation <strong>and</strong><br />

an additional two K + ions act as ‘wheel caps. The remaining potassium ions are wrapped<br />

all around 11 being coordinated to bridging <strong>and</strong> terminal oxo-groups <strong>of</strong> 11 as well water<br />

molecules <strong>of</strong> hydration. Polyanion 11 is composed <strong>of</strong> four (β-Ti 2 SiW 10 O 39 ) Keggin fragments<br />

that are linked via Ti-O-Ti bridges leading to a cyclic assembly. The only symmetry<br />

element <strong>of</strong> 11 is an inversion center, resulting in the point group C i . Close inspection <strong>of</strong><br />

11 indicates that the four Keggin fragments are <strong>of</strong> the beta-type, which is rare <strong>and</strong> the<br />

first compound containing such a dilacunary tungstosilicate fragment was the dimeric,<br />

nickel-substituted polyanion [β- Ni 2 SiW 10 O 36 (OH) 2 (H 2 O) 2 ] 12− [173]. Therefore 11 constitutes<br />

only the second example composed <strong>of</strong> a beta-decatungstosilicate Keggin fragment<br />

<strong>and</strong> it represents the first tetrameric derivative. Interestingly the two titanium atoms <strong>of</strong><br />

each beta- Keggin fragment are not located in the same M 3 O 13 triad. In fact they are<br />

93


Fig. 3.42: Side-view <strong>of</strong> compound 11 including the potassium ions (purple) inside the central cavity (K8)<br />

<strong>and</strong> above <strong>and</strong> below the cavity (K9, <strong>and</strong> K9’).<br />

separated by four bonds (Ti2-O-W3-O-Ti1 or Ti2-O-W8-O-Ti1, see Figure 3.41 (right))<br />

<strong>and</strong> one <strong>of</strong> the two Ti atoms is located in the rotated triad (Ti2 <strong>and</strong> Ti4, see Figure 3.41).<br />

Based on the IUPAC nomenclature the two titanium atoms are located in positions 1 <strong>and</strong><br />

10 (the latter being in the rotated triad) [2, 174]. In 11 each <strong>of</strong> the four Keggin fragments<br />

has a mirror plane <strong>and</strong> both titanium atoms are located in this plane <strong>of</strong> symmetry (see<br />

Figures 3.41). For comparison, the nickel(II) centers in [β- Ni 2 SiW 10 O 36 (OH) 2 (H 2 O) 2 ] 12−<br />

are in the 4,10 positions [173]. The structure <strong>of</strong> polyanion 11 can be viewed as a larger<br />

titanium-derivative <strong>of</strong> the trimeric, cyclic manganese(II)-substituted tungstosilicate [(β 2 -<br />

SiW 11 MnO 38 OH) 3 ] 15− . This product was also synthesized from [γ-SiW 10 O 36 ] 8− ) <strong>and</strong> it<br />

also contains beta-Keggin fragments [171]. However, the important difference to 11 is<br />

that its basic building block is a monosubstituted Keggin unit <strong>and</strong> as a result the Keggin-<br />

Keggin connectivities are accomplished via Mn-O-W bonds. It can be noticed that 11<br />

is not perfectly cyclic, but somewhat ellipsoidal (see Figures 3.41 (right)). This reflects<br />

the inequivalence <strong>of</strong> the two Ti-centers within each Keggin fragment. In 11, the short<br />

axis (distance between O3Ti <strong>and</strong> O3Ti’, see Figure 3.41(right)) is 10.8 Å, whereas the<br />

long axis is 13.0 Å (distance between O1TT <strong>and</strong> O1TT’). The reason for this is the fact<br />

that 1 contains two pairs <strong>of</strong> inequivalent Ti-O-Ti bridges: the type which involves two<br />

Ti centers in the rotated triad (Ti2-O1TT-Ti4, Ti2’-O1TT’- Ti4’, see Figure 3.41(right))<br />

<strong>and</strong> the type which involves two Ti centers that are not in the rotated triad (Ti3-O3Ti-<br />

Ti3’,Ti1-O3Ti-Ti1). The two types <strong>of</strong> Ti-O-Ti bond angles are only marginally different:<br />

94


152.8(12) ◦ (Ti3-O3Ti-Ti3) vs. 153.9(12) ◦ (Ti2-O1TT-Ti4) in the solid state. All <strong>of</strong> the<br />

above indicates that 11 is best described as a dimer <strong>of</strong> dimers, which sheds some light<br />

Fig. 3.43: Dimer <strong>of</strong> Dimer <strong>of</strong> β(1,10)-Ti 2 (OH) 2 SiW 10 O 38 ] 6−<br />

on its mechanism <strong>of</strong> formation. Synthesis <strong>of</strong> 11 is accomplished by reaction <strong>of</strong> TiO(SO 4 )<br />

with [γ-SiW 10 O 36 ] 8− , which means that the mechanism <strong>of</strong> formation <strong>of</strong> 11 must occur<br />

in the following sequence: (a) metal insertion, (b) rotational isomerization (γ-Keggin ··<br />

β-Keggin), (c) dimer formation <strong>and</strong> (d) ring closure. Most likely insertion <strong>of</strong> two Ti 4+<br />

ions into [γ-SiW 10 O 36 ] 8− leads to the monomeric species [γ-Ti 2 (OH) 2 SiW 10 O 38 ] 6− , which<br />

isomerizes to [β(1,10)-Ti 2 (OH) 2 SiW 10 O 38 ] 6− . This species is also unstable <strong>and</strong> dimerizes<br />

resulting in [β(1,10)- Ti 2 (OH)SiW 10 O 38 ] 2 (O) 12− , which is equivalent to the asymmetric<br />

unit <strong>of</strong> 11. In all structurally characterized Ti-substituted polyoxometalates the titanium<br />

atoms do not contain terminal bonds. Therefore it is not surprising that [β(1,10)-<br />

Ti 2 (OH)SiW 10 O 38 ] 2 (O) 12− is a highly reactive, dimeric species which reacts with other,<br />

identical dimers leading to the formation <strong>of</strong> 11. This “dimer <strong>of</strong> dimer mechanism” excludes<br />

the possibility <strong>of</strong> a gradual growth by individual Keggin units <strong>and</strong> it also implies<br />

that a trimeric intermediate is not present during formation <strong>of</strong> 11. Of course it is virtually<br />

impossible to pro<strong>of</strong> the above hypothesis <strong>and</strong> the identity <strong>of</strong> the proposed intermediates<br />

experimentally, as formation <strong>of</strong> polyoxoanions occurs via rapid self-assembly as soon as<br />

the proper reaction conditions are present. We also investigated the solution properties <strong>of</strong><br />

95


11 by 183 W-NMR (at 16.66 MHz on a JEOL Eclipse 400 instrument at room temperature<br />

using D 2 O as a solvent). We identified 10 major peaks <strong>of</strong> about equal intensity at 111.4,<br />

-118.0, - 122.8, -125.7, -141.4, -156.2, -157.8, -170.6, -173.8, - 221.0 ppm. This fits with the<br />

solid state structure <strong>of</strong> 11, because exactly this peak pattern is expected. Polyanion 11 has<br />

a center <strong>of</strong> inversion <strong>and</strong> the asymmetric unit (see Figure 3.41(right)) consists <strong>of</strong> 20 tungsten<br />

atoms. Each <strong>of</strong> the two Keggin fragments has a mirror plane so that 10 magnetically<br />

inequivalent tungsten pairs are present per asymmetric unit (W1/W2, W3/W8, W4/W7,<br />

W5/W6, W9/W10, W11/W12, W13/W18, W14/W17, W15/W16, W19/W20, see Figure<br />

3.41(right)). For the cyclic polyanion 11 this means that each <strong>of</strong> the 10 groups contains<br />

4 magnetically equivalent tungsten atoms. The 183 W-NMR spectrum also contains some<br />

additional peaks <strong>of</strong> much smaller intensity which we cannot assign yet. Nevertheless, any<br />

solid which precipitated or crystallized from the very concentrated 183 W-NMR solutions<br />

resulted in an FTIR spectrum identical to that <strong>of</strong> 11.In summary, we have synthesized<br />

a unique tetrameric <strong>and</strong> cyclic silicotungstate assembly using mild <strong>and</strong> one-pot reaction<br />

conditions.<br />

3.10.4 Conclusions<br />

Polyanion 11 was fully characterized in solution <strong>and</strong> in the solid state by several analytical<br />

techniques. Formation <strong>of</strong> 11 indicates that (a) very large <strong>and</strong> cyclic Ti(IV)-substituted silicotungstates<br />

can be formed, (b) it might be possible to construct even larger wheel-shaped<br />

polyoxotungstates, (c) it might be possible to construct inorganic nanotubes by linking or<br />

orientating individual polyoxoanion wheels appropriately, (d) the gigantic, mixed-valence<br />

polyoxomolybdate wheels have some smaller, fully oxidized polyoxotungstate analogues,<br />

(e) the structural variety <strong>of</strong> supra- <strong>and</strong> supersupramolecular polyoxotungstates is just<br />

beginning to be explored <strong>and</strong> finally that (f) undoubtedly no other class <strong>of</strong> inorganic<br />

compounds besides polyoxometalates allows for preparation <strong>of</strong> discrete, nanomolecular<br />

objects <strong>of</strong> similar size,structure <strong>and</strong> function.<br />

96


Part-IV<br />

Cadmium <strong>and</strong> Indium Containing<br />

POMs


3.11 Structure <strong>and</strong> solution properties <strong>of</strong> the<br />

cadmium(II)-substituted tungstoarsenate:<br />

(Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ) 12−<br />

3.11.1 Introduction<br />

The transition metal substituted polyoxometalates (TMSPs), known to date are the class<br />

<strong>of</strong> s<strong>and</strong>wich-type species, probably the largest subfamily [175]. In 1973 Weakley et al. described<br />

the first s<strong>and</strong>wich-type polyoxoanion, [Co 4 (H 2 O) 2 (B-α-PW 9 O 34 ) 2 ] 10− [176]. Since<br />

then, several other derivatives <strong>of</strong> this structural type have been reported. To date, the<br />

class <strong>of</strong> Weakley-type s<strong>and</strong>wich species consists <strong>of</strong> the Keggin derivatives [M 4 (H 2 O) 2 (Bα-XW<br />

9 O 34 ) 2 ] n− (n = 12, X = Ge IV , Si IV , M = Mn 2+ , Cu 2+ , Zn 2+ , Cd 2+ ; n = 10, X =<br />

P V , As V , M = Mn 2+ , Co 2+ , Ni 2+ , Cu 2+ , Zn 2+ , Cd 2+ ; n = 6, X = P V , As V , M = Fe 3+ ;<br />

n = 16, X = M = Cu 2+ ; n = 10, X = M = Fe 3+ ) <strong>and</strong> the Wells-Dawson derivatives<br />

[M 4 (X 2 W 15 O 56 ) 2 ] n− (X = P V , As V , n = 16, M = Mn 2+ , Co 2+ , Ni 2+ , Cu 2+ , Zn 2+ , Cd 2+ ,<br />

n = 12, M = Fe 3+ ) [172, 176–195]. The second class <strong>of</strong> s<strong>and</strong>wich-type POMs is based on<br />

two lone-pair containing, α-Keggin fragments, e.g. [α-As III W 9 O 33 ] 9− . The first member<br />

<strong>of</strong> this class ([Cu 3 (H 2 O) 2 (α-AsW 9 O 33 ) 2 ] 12− ) was reported by Hervé et al. in 1982 [196].<br />

Since then a number <strong>of</strong> isostructural derivatives has been characterized: [M 3 (H 2 O) 3 (α-<br />

XW 9 O 33 ) 2 ] n− (n = 12, X = As III , Sb III , M = Mn 2+ , Co 2+ , Ni 2+ , Cu 2+ , Zn 2+ ; n = 10, X<br />

= Se IV , Te IV , M = Cu 2+ ) <strong>and</strong> [(VO) 3 (α-XW 9 O 33 ) 2 ] n− (n = 12, X = As III , Sb III , Bi III ;<br />

n = 11, X = As III ) [197–203]. Kortz et al. reported on the tri-palladium(II) substituted<br />

derivative [Pd 3 (α-SbW 9 O 33 ) 2 ] 12− .[204] The third class <strong>of</strong> s<strong>and</strong>wich-type POMs is based on<br />

two lone-pair containing, β-Keggin fragments, e.g. [β-Sb III W 9 O 33 ] 9− . The first members<br />

<strong>of</strong> this class, ([M 2 (H 2 O) 6 (WO 2 ) 2 (β-SbW 9 O 33 ) 2 ] (14−2n)− (M n+ = Fe 3+ , Co 2+ , Mn 2+ , Ni 2+ ),<br />

were reported by Krebs et al. in 1997 [69]. Since then some more isostructural derivatives<br />

have been characterized: ([M 2 (H 2 O) 6 (WO 2 ) 2 (β-BiW 9 O 33 ) 2 ] (14−2n)− (M n+ = Fe 3+ , Co 2+ ,<br />

Ni 2+ , Cu 2+ , Zn 2+ ), [(VO(H 2 O) 2 ) 2 (WO 2 ) 2 (β-BiW 9 O 33 ) 2 ] 10− , [Sn 1.5 (WO 2 (OH)) 0.5 (WO 2 ) 2<br />

(β-XW 9 O 33 ) 2 ] 10.5− (X = Sb III , Bi III ), [M 3 (H 2 O) 8 (WO 2 )(β-TeW 9 O 33 ) 2 ] 8− (M = Ni 2+ ,<br />

Co 2+ ), [(Zn(H 2 O) 3 ) 2 (WO 2 ) 1.5 (Zn(H 2 O) 2 ) 0.5 (β-TeW 9 O 33 ) 2 ] 8− , [(VO(H 2 O) 2 ) 1.5 (WO(H 2 O) 2 ) 0.5<br />

98


(WO 2 ) 0.5 (VO(H 2 O)) 1.5 (β-TeW 9 O 33 ) 2 ] 7− <strong>and</strong> [M 4 (H 2 O) 10 (β-XW 9 O 33 ) 2 ] n− (n = 6, X =<br />

As III <strong>and</strong> Sb III , M = Fe III <strong>and</strong> Cr III ; n = 4, X = Se IV , Te IV , M = Fe III <strong>and</strong> Cr III ; n<br />

= 8, X = Se IV , Te IV , M = Mn II , Co II , Ni II , Zn II , Cd II <strong>and</strong> Hg II ) [83, 203, 205–207].<br />

The fourth class <strong>of</strong> s<strong>and</strong>wich-type POMs is based on two A-α-Keggin fragments, e.g.<br />

[A-α-PW 9 O 34 ] 9− . The first member <strong>of</strong> this class, [Co 3 ](H 2 O) 3 (A-α-PW 9 O 34 ) 2 ] 12− , was<br />

reported by Knoth et al. in 1985 [208]. Since then the following isostructural derivatives<br />

have been identified: [M 3 (A-XW 9 O 34 ) 2 ] n− (n = 14, X = Si IV , M = Sn 2+ , Co 2+ ; n =<br />

12, X = P V , M = Mn 2+ , Ni 2+ , Cu 2+ , Zn 2+ , Pd 2+ , Sn 2+ ; n = 9, X = P V , M = Fe 3+ ),<br />

[(CeO) 3 (H 2 O) 2 (A-PW 9 O 34 ) 2 ] 12− <strong>and</strong> [(ZrOH) 3 (A-SiW 9 O 34 ) 2 ] 11− [209–212]. Very recently<br />

Kortz et al. reported on the di-palladium(II) substituted derivative [Pd 2 WO(H 2 O)(A-α-<br />

SiW 9 O 34 ) 2 ] 12− .[167] Surprisingly little work on cadmium-substituted polyoxotungstates<br />

has been reported to date, especially regarding structurally characterized species. Nevertheless,<br />

the diamagnetic nature <strong>of</strong> Cd-containing POMs has prompted some NMR studies.<br />

Twenty years ago Contant reported for the first time on cadmium-containing Keggin <strong>and</strong><br />

Wells-Dawson species [213–215]. In 1995 Kirby <strong>and</strong> Baker reported on a NMR study<br />

<strong>of</strong> the first s<strong>and</strong>wich-type polyoxometalate including Cd 2+ ions [216]. More recently, Bi<br />

et al. synthesized Cd-substituted, s<strong>and</strong>wich type polyanions based on the tungstoarsenate<br />

fragments [As 2 W 15 O 56 ] 12− <strong>and</strong> [AsW 9 O 34 ] 9− , respectively.[191, 192] Very recently,<br />

Alizadeh et al. synthesized <strong>and</strong> structurally characterized the tetra-cadmium containing<br />

[Cd 4 (H 2 O) 2 (PW 9 O 34 ) 2 ] 10− [217]. Here we report on interaction <strong>of</strong> Cd 2+ ions with the<br />

trilacunary Keggin derivative [A-HAsW 9 O 34 ] 8− in aqueous solution.<br />

3.11.2 Experimental<br />

Synthesis <strong>of</strong> lacunary polyanion precursor Na 8 [A-HAsW 9 O 34 ]·11H 2 O : A 30 g (91 mmol)<br />

sample <strong>of</strong> Na 2 WO 4·2H 2 O <strong>and</strong> 1.18 g (5.1 mmol) As 2 O 5 were dissolved in 40 mL <strong>of</strong> water.<br />

Subsequently, the pH was adjusted to 8.4 using glacial acetic acid. After a few<br />

seconds the solution became cloudy <strong>and</strong> after about 2 min. a white precipitate started<br />

to form. The solution was stirred for another 30 min. <strong>and</strong> then the solid was isolated<br />

on a sintered funnel. The product was washed three times with ethanol <strong>and</strong> air-dried<br />

with aspiration. FTIR spectroscopy: 937(m), 884(m), 851(m), 765(s), 672(m), 518(w),<br />

99


472(w), 444(w) cm −1 . This procedure is different from that reported previously by one <strong>of</strong><br />

us, which involves using H 3 AsO 4 instead <strong>of</strong> As 2 O 5 .[72] Synthesis <strong>of</strong> Cs 4 K 3 Na 5 [Cd 4 Cl 2 (Bα-AsW<br />

9 O 34 ) 2 ]·20H 2 O (CsKNa-12) A 0.33 g (1.66 mmol) sample <strong>of</strong> CdCl 2·H 2 O was<br />

dissolved with stirring in 20 mL <strong>of</strong> 0.5 M NaAc buffer (pH 4.8). Then 2.00 g (0.75 mmol)<br />

<strong>of</strong> Na 8 [A-HAsW 9 O 34 ]·11H 2 O was added. The solution was heated to ∼80 ◦ C for about 1<br />

h <strong>and</strong> filtered after it had cooled. Then 0.5 mL <strong>of</strong> 1.0 M CsCl <strong>and</strong> 0.5 mL <strong>of</strong> 1.0 M KCl<br />

solutions were added to the filtrate. Slow evaporation at room temperature led to 1.8<br />

g (yield 78%) <strong>of</strong> a white crystalline product after about one week. FTIR spectroscopy:<br />

951(s), 882(s), 835(s), 784(s), 746 (sh), 724 (s), 480 (w), 442 (w) cm −1 . Anal. Calcd<br />

(Found) for 1a: Cs 8.6 (8.3), K 1.9 (2.1), Na 1.9 (1.9), W 53.4 (53.9), Cd 7.3 (7.1), As<br />

2.4 (2.4), Cl 1.2(1.3). Polyanion 13 was prepared by an analogous procedure as 12, but<br />

Fig. 3.44: FTIR spectra <strong>of</strong> compound (CsKNa-12)(red) <strong>and</strong> Na 8 [A-HAsW 9 O 34 ]·11H 2 O (blue)<br />

Cd(NO 3 ) 2·4H 2 O was used instead <strong>of</strong> CdCl 2·H 2 O to ascertain the absence <strong>of</strong> chloride ions.<br />

X-ray Crystallography<br />

A colorless crystal with dimensions 0.26 × 0.12 × 0.08 mm was selected <strong>and</strong> mounted on<br />

a glass fiber for indexing <strong>and</strong> intensity data collection at 167 K on a Bruker D8 SMART<br />

APEX CCD single-crystal diffractometer using Mo Kα radiation (λ = 0.71073 Å). Direct<br />

methods were used to solve the structure <strong>and</strong> to locate the heavy atoms (SHELXS97).<br />

Then the remaining atoms were found from successive difference maps (SHELXL97).<br />

100


Fig. 3.45: FTIR spectra <strong>of</strong> compound (CsNa-13)(red) <strong>and</strong> Na 8 [A-HAsW 9 O 34 ]·11H 2 O (blue)<br />

Routine Lorentz <strong>and</strong> polarization corrections were applied <strong>and</strong> an absorption correction<br />

was performed using the SADABS program [81]. Crystallographic data are summarized<br />

in Table 3.11.<br />

Table 3.11: Crystal Data <strong>and</strong> Structure Refinement for compound CsKNa-12<br />

Emperical formula Cs 4 K 3 Na 5 [Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ]·20H 2 O<br />

fw 6192.0<br />

space group (No.) P 2 1 /n (14)<br />

a (Å) 13.1402(12)<br />

b (Å) 19.0642(17)<br />

c (Å) 17.5666(15)<br />

vol (Å 3 ) 4400.5(7)<br />

Z 2<br />

temp ( ◦ C) -106<br />

wavelength (Å) 0.71073<br />

d calcd (mg m −3 ) 4.608<br />

abs coeff. (mm −1 ) 27.069<br />

R [I > 2 σ(I)] a 0.046<br />

R w (all data) b 0.107<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

Solution NMR spectroscopy: 183 W <strong>and</strong> 111 Cd NMR<br />

All NMR spectra were recorded on a 400 MHz JEOL ECX instrument at room temperature<br />

using D 2 O as a solvent. The 183 W-NMR measurements were performed at 16.656<br />

101


MHz in 10 mm tubes <strong>and</strong> the 111 Cd NMR measurements were performed at 84.757 MHz<br />

in 5 mm tubes. As references (external st<strong>and</strong>ards) we used aqueous solutions <strong>of</strong> 1 M<br />

Na 2 WO 4 <strong>and</strong> 0.1 M Cd(ClO 4 ) 2 , respectively. 183 W NMR is a very powerful solution technique<br />

in polyanion chemistry to obtain structural information. The fact that polyanions<br />

12 <strong>and</strong> 13 are diamagnetic makes them ideal c<strong>and</strong>idates for NMR studies. In addition<br />

both species contain cadmium ions <strong>and</strong> therefore we decided to perform a detailed 183 W<br />

<strong>and</strong> 111 Cd NMR study on 12 <strong>and</strong> 13. If the dimeric, s<strong>and</strong>wich-type polyanion structure<br />

with C 2h symmetry is preserved in solution, then five peaks (intensity ratio 2:2:2:2:1) are<br />

expected in 183 NMR <strong>and</strong> two peaks (1:1) are expected for 111 Cd NMR. In fact, this is<br />

exactly what we observed. The room-temperature 183 W-NMR spectrum <strong>of</strong> 12 exhibits<br />

five peaks at -75.9, -92.0, -94.2, -109.8, -121.8 ppm with intensity ratios 1:2:2:2:2 (see<br />

Figure 3.46, top). For 13 we obtained an almost identical spectrum with five peaks at<br />

-75.9, -92.6, -94.3, -107.9, -119.9 ppm <strong>and</strong> the anticipated intensity ratios (see Figure 3.46,<br />

bottom). As expected, a change <strong>of</strong> the terminal lig<strong>and</strong> on the external cadmium centers<br />

in 12 <strong>and</strong> 13 does not affect the chemical shifts <strong>of</strong> the tungsten centers significantly. In<br />

fact, only four <strong>of</strong> the six belt tungsten atoms in each Keggin unit should be affected,<br />

as they are only 3 bonds from the terminal chloro lig<strong>and</strong>s in 12 or aqua lig<strong>and</strong>s in 13<br />

(see Figure 3.45). Therefore we suggest that the two upfield signals in the 183 W NMR<br />

spectra <strong>of</strong> 12 <strong>and</strong> 13 (see Figure 3.46) correspond most likely to those tungsten atoms in<br />

question. All other tungsten centers are further away from the terminal cadmium lig<strong>and</strong>s<br />

in 12 <strong>and</strong> 13. Of course, the unique cap tungsten atom corresponds to the smallest, most<br />

downfield signal. The 183 W NMR spectra <strong>of</strong> 12 <strong>and</strong> 13 correspond also very nicely to our<br />

results on the isostructural germanium derivatives [Cd 4 (H 2 O) 2 (B-α-GeW 9 O 34 ) 2 ] 12− <strong>and</strong><br />

[Zn 4 (H 2 O) 2 (B-α-GeW 9 O 34 ) 2 ] 12− .[195] The 111 Cd NMR spectrum <strong>of</strong> 12 shows two singlet<br />

peaks at 55.3 <strong>and</strong> 14.7 ppm, respectively (see Figure 3.47, top). The upfield signal is very<br />

broad which is probably due to the quadrupolar moment <strong>of</strong> the chloro lig<strong>and</strong>. Therefore,<br />

we can assign this signal to the external cadmium ions in 12 whereas the other, sharp<br />

signal corresponds to the internal cadmium centers. The 111 Cd NMR spectrum <strong>of</strong> 13<br />

exhibits two sharp peaks at 57.0 <strong>and</strong> 11.1 ppm, respectively with equal intensities (see<br />

Figure 3.47, bottom). This is fully consistent with water molecules as terminal lig<strong>and</strong>s<br />

102


Fig. 3.46: 183 W NMR spectra <strong>of</strong> [Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ] 12− (12), top, <strong>and</strong> [Cd 4 (H 2 O) 2 (B-α-<br />

AsW 9 O 34 ) 2 ] 10− (13)<br />

on the external cadmium ions. In analogy to 12 we assign the upfield signal at 11.1<br />

ppm to the external cadmium ions <strong>and</strong> the downfield signal at 57.0 ppm to the internal<br />

cadmium ions <strong>of</strong> 13. These results are also in good agreement with Alizadeh et al., who<br />

performed 113 Cd NMR studies on [Cd 4 (H 2 O) 2 (B-α-PW 9 O 34 ) 2 ] 10− . They observed a very<br />

similar chemical shift for the internal cadmium ions (52.76 ppm), but the signal for the<br />

outer cadmium centers is significantly more downfield (26.25 ppm) [217]. This difference<br />

in 111 Cd NMR behavior must be due to the different hetero atoms X (X = As v , P v ),<br />

as all other parameters are the same for both polyanions (e.g. structure, charge). Our<br />

work shows that this ‘hetero atom effect’ is more pronounced for the external than the<br />

internal cadmium ions. The nuclear shielding <strong>of</strong> the outer cadmium centers is a much<br />

better sensor for the type <strong>of</strong> hetero atom present than that <strong>of</strong> the internal cadmiums, in<br />

spite <strong>of</strong> the fact that the Cd-O bond lengths <strong>and</strong> Cd-O-X angles are very similar in both<br />

103


Fig. 3.47: 111 Cd NMR spectra <strong>of</strong> [Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ] 12− 12, top, <strong>and</strong> [Cd 4 (H 2 O) 2 (B-α-<br />

AsW 9 O 34 ) 2 ] 10− 13, bottom.<br />

cases.<br />

3.11.3 Results <strong>and</strong> discussion<br />

We have synthesized the dimeric tungstoarsenate(V) [Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ] 12− (1) as<br />

a mixed cesium-potassium-sodium salt. The title polyanion consists <strong>of</strong> two trilacunary<br />

[B-α-AsW 9 O 34 ] 9− Keggin units linked via a rhomblike Cd 4 O 14 Cl 2 group resulting in a<br />

s<strong>and</strong>wich-type structure with idealized C 2h symmetry (see Figure 3.48). Polyanion 12 was<br />

synthesized by interaction <strong>of</strong> Cd 2+ ions with the trilacunary precursor [A-HAsW 9 O 34 ] 8−<br />

in aqueous acidic medium upon heating to ∼80 ◦ C. This indicates that the Keggin precursor<br />

has undergone isomerization from A-AsW9 → B-AsW9 during the course <strong>of</strong> the<br />

reaction. However, this transformation is well known <strong>and</strong> has been observed in solution<br />

<strong>and</strong> in the solid state [192]. Bond valence sum (BVS) calculations confirm that there are<br />

no protonation sites on 12, indicating that the charge <strong>of</strong> the title polyanion must be 12<br />

[85]. Crystallographically we could identify only 4 Cs + , 3 K + <strong>and</strong> 3 Na + counterions,<br />

104


Fig. 3.48: Combined polyhedral <strong>and</strong> ball <strong>and</strong> stick representation <strong>of</strong> [Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ] 12−<br />

but elemental analysis indicated that indeed 5 Na + ions were present in 12. Therefore<br />

the charge <strong>of</strong> 12 for 12 is fully balanced in the solid state by an appropriate number <strong>of</strong><br />

counter ions. Disorder <strong>of</strong> some alkali ions (in particular sodium) is a common problem<br />

in polyanion chemistry. The structural type <strong>of</strong> 12 had first been described in 1973 by<br />

Weakley et al. for [Co 4 (H 2 O) 2 (PW 9 O 34 ) 2 ] 10− [10a]. To date this Keggin-based structure<br />

is known for most first-row transition metals (including mixed-metal derivatives)<br />

<strong>and</strong> it has also been possible to substitute the tetrahedral phosphorus(V) heteroatom by<br />

arsenic(V), silicon(IV), germanium(IV), iron(III), <strong>and</strong> copper(II) [172, 176–195]. Therefore,<br />

this structural type represents one <strong>of</strong> the largest TMSP families. The Weakley-type<br />

structures have in common that the two external transition metal ions in the central rhombus<br />

have one terminal lig<strong>and</strong> each, whereas the two internal metal ions do not have any<br />

terminal lig<strong>and</strong>s. In all <strong>of</strong> the above known derivatives with the Weakley structure, the<br />

terminal lig<strong>and</strong> on the external transition metal ions is a water molecule. However, in 12<br />

the two external cadmium ions have a chloro lig<strong>and</strong> instead <strong>of</strong> a water molecule. Clearly,<br />

the chloride ions originate from the cadmium source (CdCl 2·H 2 O) which we used for the<br />

synthesis <strong>of</strong> 12. Nevertheless, it must be remembered that synthesis <strong>of</strong> 12 is carried out<br />

105


in aqueous solution, so that the number <strong>of</strong> water molecules dominates by far the chloride<br />

ions. Therefore, the external cadmium ions incorporated in 12 have a strong preference<br />

for chloro rather than aqua lig<strong>and</strong>s. Recently one <strong>of</strong> us reported on the aqua derivative<br />

<strong>of</strong> 12 which was formulated as [Cd 4 (H 2 O) 2 (B-α-AsW 9 O 34 ) 2 ] 10− based on FTIR spectroscopy<br />

[72]. Very recently Alizadeh et al. reported on [Cd 4 (H 2 O) 2 (B-α-PW 9 O 34 ) 2 ] 10−<br />

which represents the phosphorus derivative <strong>of</strong> 12 [217]. They prepared this polyanion by<br />

reaction <strong>of</strong> Cd(NO 3 ) 2·4H 2 O with Na 8 [A-α-HPW 9 O 34 ]·24H 2 O in aqueous acidic medium<br />

(pH 6). These authors characterized their product in solution by 31 P <strong>and</strong> 113 Cd NMR<br />

spectroscopy. Very recently our group has synthesized the germanium(IV) derivative<br />

[Cd 4 (H 2 O) 2 (B-α-GeW 9 O 34 ) 2 ] 12− by reaction <strong>of</strong> CdCl 2 , GeO 2 , <strong>and</strong> Na 2 WO 4·2H 2 O in 0.5<br />

M sodium acetate buffer (pH 4.8) [195]. Interestingly, the external cadmium ions in the<br />

product polyanion contain aqua rather than chloro lig<strong>and</strong>s. It is possible that the higher<br />

affinity <strong>of</strong> Cd 2+ for chloro ions is overcompensated by charge density considerations <strong>of</strong><br />

the germanium(IV) based polyanion. Specifically, the hypothetical dichloro derivative<br />

[Cd 4 Cl 2 (B-α-GeW 9 O 34 ) 2 ] 14− would have a charge <strong>of</strong> 14-, which is apparently less stable<br />

than the diaqua derivative [Cd 4 (H 2 O) 2 (B-α-GeW 9 O 34 ) 2 ] 12− with a charge <strong>of</strong> 12-. Interestingly,<br />

our [Cd 4 (H 2 O) 2 (B-α-GeW 9 O 34 ) 2 ] 12− <strong>and</strong> the [Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ] 12− reported<br />

here have both the same charge <strong>of</strong> 12-. It is <strong>of</strong> interest to evaluate in detail the bond<br />

lengths <strong>and</strong> angles <strong>of</strong> the cadmium centers in 12 <strong>and</strong> to compare these with the isostructural<br />

derivatives [Cd 4 (H 2 O) 2 (B-α-PW 9 O 34 ) 2 ] 10− <strong>and</strong> [Cd 4 (H 2 O) 2 (B-α-GeW 9 O 34 ) 2 ] 12− . In<br />

general, as Cd 2+ is a d 10 ion no Jahn-Teller distortion is expected. In fact, the two inner<br />

Cd centers <strong>of</strong> 12 exhibit a fairly regular coordination sphere <strong>and</strong> the Cd(1)-O distances<br />

range from 2.19-2.34(1) Å. On the other h<strong>and</strong>, the coordination sphere <strong>of</strong> the external<br />

Cd(2) shows a Jahn-Teller distortion, which is most likely due to the presence <strong>of</strong><br />

the terminal chloro lig<strong>and</strong>. The equatorial Cd(2)-O distances range from 2.24-2.29(1) Å,<br />

whereas the axial Cd(2)-O <strong>and</strong> Cd(2)-Cl distances are 2.47(1) <strong>and</strong> 2.475(4) Å, respectively.<br />

The bond angles in the central Cd 4 Cl 2 O 14 fragment can be classified according<br />

to the type <strong>of</strong> bridging lig<strong>and</strong>. The Cd-O-Cd angles involving the µ 4 -oxo lig<strong>and</strong> O1As<br />

are: Cd1-O1As-Cd1’ 101.2(4) ◦ , Cd1-O1As-Cd2 93.3(3) ◦ <strong>and</strong> Cd1-O1As-Cd2 94.6(3) ◦ ,<br />

respectively. On the other h<strong>and</strong>, the angles around the µ 3 -oxo lig<strong>and</strong>s O8C2 <strong>and</strong> O9Cd<br />

106


are: Cd1-O8C2-Cd2 98.9(4) ◦ <strong>and</strong> Cd1-O9Cd-Cd2 100.4(4) ◦ . The Cd···Cd separations<br />

in 1 are as follows: Cd1···Cd2 3.533(2) Å, Cd1···Cd2’ 3.494(2) Å <strong>and</strong> Cd1···Cd1’ 3.614(2)<br />

Å, respectively. Comparison <strong>of</strong> these parameters for 12 with those <strong>of</strong> the structurally<br />

closely related [Cd 4 (H 2 O) 2 (B-α-GeW 9 O 34 ) 2 ] 12− [195] <strong>and</strong> [Cd 4 (H 2 O) 2 (B-α-PW 9 O 34 ) 2 ] 10−<br />

[217] allows to draw the following conclusions: the terminal chloro lig<strong>and</strong>s in 12 result in<br />

a significantly longer terminal cadmium bond (Cd-Cl 2.475 (5) Å) compared to those in<br />

the two aqua compounds (2.260 (14) <strong>and</strong> 2.309(8) Å). This is the reason for the more<br />

pronounced Jahn-Teller distortion <strong>of</strong> the external Cd centers in 12, which in turn results<br />

in larger Cd1···Cd2 distances (e.g. 3.494(2) Å, 3.533(2) Å in 1 vs. 3.357(2) Å, 3.404(2) Å<br />

in [Cd 4 (H 2 O) 2 (B-α-GeW 9 O 34 ) 2 ] 12− ). We were also able to synthesize the aqua derivative<br />

<strong>of</strong> 12 with the proposed formula [Cd 4 (H 2 O) 2 (B-α-AsW 9 O 34 ) 2 ] 10− 13 as based on FTIR<br />

<strong>and</strong> multinuclear NMR spectroscopies.<br />

3.11.4 Conclusions<br />

The dimeric, cadmium-substituted tungstoarsenate [Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ] 12− Polyanion<br />

13 was prepared by an analogous procedure as 12, but Cd(NO 3 ) 2·4H 2 O was used<br />

instead <strong>of</strong> CdCl 2·H 2 O to ascertain the absence <strong>of</strong> chloride ions. Polyanion 13 has been<br />

synthesized by reaction <strong>of</strong> Cd 2+ ions with [A-HAsW 9 O 34 ] 8− in aqueous, acidic medium.<br />

has a s<strong>and</strong>wich-type structure based on two [B-α-AsW 9 O 34 ] 9− Keggin moieties encapsulating<br />

four cadmium centers in a rhomblike (Cd 4 O 14 Cl 2 ) group. Therefore, 12 represents<br />

only the second structurally characterized, Cd-containing polyanion <strong>and</strong> it is the first<br />

structurally characterized, Cd-containing tungstoarsenate(V). An interesting aspect <strong>of</strong><br />

12 is the fact that the external cadmium centers have terminal chloro lig<strong>and</strong>s. Nevertheless,<br />

we were also able to synthesize the aqua derivative [Cd 4 (H 2 O) 2 (B-α-AsW 9 O 34 ) 2 ] 10−<br />

13. Both species were investigated by 183 W <strong>and</strong> 111 Cd NMR in solution <strong>and</strong> we have<br />

shown that the former is an ideal method to establish the purity <strong>and</strong> to confirm the overall<br />

polyanion structure <strong>of</strong> 12 <strong>and</strong> 13. On the other h<strong>and</strong>, the latter is an ideal technique<br />

to study the local coordination environment <strong>of</strong> the cadmium ions <strong>and</strong> it also allows to<br />

clearly distinguish 12 <strong>and</strong> 13 due to peak broadening caused by the quadrupolar chloride<br />

ions. Furthermore, this effect allows to assign the two Cd-NMR peaks to the two types<br />

107


<strong>of</strong> cadmium ions in 12 <strong>and</strong> 13 <strong>and</strong> it also allows to indirectly assign some <strong>of</strong> the 183 W<br />

NMR peaks <strong>of</strong> the title polyanions. In conclusion, synthesis <strong>of</strong> diamagnetic derivatives<br />

(e.g. Cd 2+ , Zn 2+ ) <strong>of</strong> TMSPs is very attractive, as it allows to perform multinuclear NMR<br />

studies in solution involving addenda atoms (e.g. 183 W 111 Cd). Although single crystal<br />

X-ray analysis is the most powerful technique to establish novel polyanion structures,<br />

multinuclear NMR is a very important, complementary tool to establish the structural<br />

integrity <strong>of</strong> polyanions in solution. Considering that polyanions have a multitude <strong>of</strong> potential<br />

applications in solution (e.g. medicine, catalysis) such diamagnetic complexes can<br />

be considered as model compounds to help underst<strong>and</strong> structure-activity relationships <strong>of</strong><br />

entire families <strong>of</strong> polyanions.<br />

3.12 Some indium(III)-substituted polyoxotungstates<br />

<strong>of</strong> the Keggin <strong>and</strong> Dawson type<br />

3.12.1 Introduction<br />

The synthesis <strong>of</strong> polyoxometalates (POMs) is mostly straightforward, once the proper reactions<br />

conditions have been identified. However, the mechanism <strong>of</strong> formation <strong>of</strong> POMs is<br />

not yet well understood <strong>and</strong> commonly described as self-assembly. Therefore, the design<br />

<strong>of</strong> novel POMs remains a challenge for synthetic chemists. The most rational synthesis<br />

procedure <strong>of</strong> POMs involves the use <strong>of</strong> lacunary precursors. Lacunary POMs are<br />

usually synthesized from complete precursor ions by loss <strong>of</strong> one or more MO 6 octahedra.<br />

Reaction <strong>of</strong> a stable, lacunary polyoxometalate with transition metal ions usually<br />

leads to a product with the heteropolyanion framework unchanged. This approach usually<br />

results in monomeric or dimeric polyoxoanions with expected structures, e.g. [A-α-<br />

SiW 9 O 34 ] 10− + 3Cu 2+ −→ [Cu 3 (H 2 O) 3 (A-α-SiW 9 O 37 ] 10− <strong>and</strong> 2[B-α -PW 9 O 34 ] 9− + 4Co 2+<br />

−→[Co 4 (H 2 O) 2 (B-α-PW 9 O 34 ) 2 ] 10− .[176, 218]<br />

S<strong>and</strong>wich-type POMs based on two [B-α-XW 9 O 34 ] n− (X = P V , As V , Si IV , Ge IV ) or<br />

[X 2 W 15 O 56 ] 12− (X = P V , As V ) fragments <strong>and</strong> four transition metal centers constitute a<br />

well-known class <strong>of</strong> compounds [195]. To date a large number <strong>of</strong> derivatives has been<br />

reported for both the Keggin <strong>and</strong> Wells-Dawson type structures. It has become apparent<br />

108


that this dimeric, s<strong>and</strong>wich-type structure (Weakley-type) allows for incorporation <strong>of</strong> essentially<br />

all first-row <strong>and</strong> some second row transition metal ions [195].<br />

Recently several examples <strong>of</strong> Weakley-type s<strong>and</strong>wich POMs with less than four transition<br />

metals have been reported [219]. Hill et al. described di- <strong>and</strong> tri-iron-substituted<br />

polyanions <strong>of</strong> the Wells-Dawson-type <strong>and</strong> interestingly the vacancies were occupied by<br />

sodium ions (e.g. [Fe 2 (NaOH 2 ) 2 (P 2 W 15 O 56 ) 2 ] 16− , [Fe 2 (FeOH 2 )(NaOH 2 )(P 2 W 15 O 56 ) 2 ] 14− ).<br />

However, it has been possible to substitute these sodium ions by first row-transition<br />

metal ions leading to mixed-metal s<strong>and</strong>wich-type polyanions [219]. The first example <strong>of</strong><br />

a tri-substituted, s<strong>and</strong>wich-type polyoxoanion based on a Keggin fragment is also known<br />

([Ni 3 Na(H 2 O) 2 (B-α-PW 9 O 34 ) 2 ] 11− [220].<br />

Lone pair containing polyoxotungstates (e.g. [Cs 2 Na(H 2 O) 10 Pd 3 (α-Sb III W 9 O 33 ) 2 ] 9− ,<br />

[Cu 4 K 2 (H 2 O) 8 ](α-As III W 9 O 33 ) 2 ) 8− ), [As III 6W 65 O 217 (H 2 O) 7 ] 26− ) are a well-established subclass<br />

<strong>of</strong> POMs [33, 202, 204]. The presence <strong>of</strong> a lone pair on the hetero element does not<br />

allow the closed Keggin unit to form. The dimeric, s<strong>and</strong>wich-type POMs based on two<br />

lone pair containing, β-Keggin fragments (e.g. [β -As III W 9 O 33 ] 9− , [β-Sb III W 9 O 33 ] 9− )<br />

represent also a well-known class <strong>of</strong> s<strong>and</strong>wich-type POMs. The first members <strong>of</strong> this<br />

class, ([M 2 (H 2 O) 6 (WO 2 ) 2 (β-SbW 9 O 33 ) 2 ] (14−2n)− (M n+ = Fe 3+ , Co 2+ , Mn 2+ , Ni 2+ ), were<br />

reported by Krebs et al. in 1997 [69]. Since then some more isostructural derivatives<br />

have been characterized: ([M 2 (H 2 O) 6 (WO 2 ) 2 (β-BiW 9 O 33 ) 2 ] (14−2n)− (M n+ = Fe 3+ , Co 2+ ,<br />

Ni 2+ , Cu 2+ , Zn 2+ ), [(VO(H 2 O) 2 ) 2 (WO 2 ) 2 (β-BiW 9 O 33 ) 2 ] 10− , [Sn 1.5 (WO 2 (OH)) 0.5 (WO 2 ) 2 (<br />

β-XW 9 O 33 ) 2 ] 10.5− (X=Sb III ,Bi III ), [M 3 (H 2 O) 8 (WO 2 )(β-TeW 9 O 33 ) 2 ] 8− (M=Ni 2+ ,Co 2+ ),<br />

[(Zn(H 2 O) 3 ) 2 (WO 2 ) 1.5 (Zn(H 2 O) 2 ) 0.5 (β-TeW 9 O 33 ) 2 ] 8− , [(VO(H 2 O) 2 ) 1.5 (WO(H 2 O) 2 ) 0.5 (WO 2 )<br />

0.5(VO(H 2 O)) 1.5 (β-TeW 9 O 33 ) 2 ] 7− <strong>and</strong> [M 4 (H 2 O) 10 (β-XW 9 O 33 ) 2 ] n− (n = 6, X = As III <strong>and</strong><br />

Sb III , M = Fe III <strong>and</strong> Cr III ; n = 4, X = Se IV , Te IV , M = Fe III <strong>and</strong> Cr III ; n = 8, X =<br />

Se IV , Te IV , M = Mn II , Co II , Ni II , Zn II , Cd II <strong>and</strong> Hg II [83, 203, 205–207].<br />

It becomes apparent that most <strong>of</strong> the known s<strong>and</strong>wich-type POMs contain divalent, paramagnetic<br />

transition metal ions. The motivation for synthesizing such compounds in the<br />

first place is usually centered around perceived applications in catalysis <strong>and</strong> medicine.<br />

Therefore, it is highly important to also investigate the solution properties <strong>of</strong> such POMs<br />

in addition to the solid state. The most sensitive <strong>and</strong> elegant tool for such studies is un-<br />

109


doubtedly 183 W-NMR, but the paramagnetic nature <strong>of</strong> most s<strong>and</strong>wich-type POMs complicates<br />

matters. Therefore, it is <strong>of</strong> interest to prepare isostructural <strong>and</strong> diamagnetic<br />

analogs <strong>of</strong> these compounds.<br />

In several cases it was possible to substitute the divalent, paramagnetic transition metal<br />

ions by diamagnetic Zn 2+ <strong>and</strong> then 183 W-NMR was used to pro<strong>of</strong> the structural integrity<br />

<strong>of</strong> the POMs in solution (e.g. [M 3 (H 2 O) 3 (α-XW 9 O 33 ) 2 ] 12− (M = Cu 2+ , Zn 2+ ; X = As III ,<br />

Sb III ) [39, 195]. Obviously, substitution <strong>of</strong> Fe(III)-containing POMs by Zn 2+ results in<br />

derivatives with a different charge (e.g. [Fe 4 (H 2 O) 10 (β-SeW 9 O 33 ) 2 ] 4− vs [Zn 4 (H 2 O) 10 (β-<br />

SeW 9 O 33 ) 2 ] 8− ). In order to solve this problem, we decided to prepare In(III)-substituted<br />

POMs with structures for which Fe(III)-analogs exist.<br />

Interestingly, very little work on indium-substituted polyoxometalates has been reported<br />

to date [221–223]. In 1995, Liu et al. described interaction <strong>of</strong> indium(III) with the trilacunary<br />

Keggin species α,β-[XW 9 O 34 ] 10− (X = Si, Ge) <strong>and</strong> based on 183 W-NMR they proposed<br />

trisubstituted, monomeric products [221]. A few years later, Wasfi et al. synthesized<br />

an indium(III) substituted heteropolyfluorotungstate <strong>and</strong> they proposed a Dawson-like<br />

structure [222]. Very recently, Krebs et al. described the first structurally characterized,<br />

indium(III) substituted polyanions [223]. Here we report on tri- <strong>and</strong> tetra-indium(III)<br />

substituted polyoxotungstates based on Keggin <strong>and</strong> Wells-Dawson fragments.<br />

3.12.2 Experimental<br />

The lacunary precursors Na 8 H[B-α-PW 9 O 34 ], Na 12 [P 2 W 15 O 56 ], Na 9 [α-AsW 9 O 33 ] <strong>and</strong> Na 9 [α-<br />

SbW 9 O 33 ] were synthesized according to published procedures <strong>and</strong> their purity was confirmed<br />

by infrared spectroscopy [69, 73, 75, 128]. All other reagents were used as purchased<br />

without further purification.<br />

D,L-(NH 4 ) 11 [In 3 Cl 2 (B-α-PW 9 O 34 ) 2 ]· 16H 2 O (NH4-14). A 0.39 g (1.76 mmols) sample <strong>of</strong><br />

InCl 3 was dissolved in 40 mL H 2 O followed by addition <strong>of</strong> 1.94 g (0.80 mmol) Na 8 H[B-α-<br />

PW 9 O 34 ]. The pH <strong>of</strong> this solution was adjusted to 2 by addition <strong>of</strong> 4 M HCl <strong>and</strong> then<br />

it was heated to ∼80 ◦ C for 1 h. After cooling to room temperature the solution was<br />

filtered. Then it was layered with 0.1 M NH 4 Cl <strong>and</strong> allowed to evaporate in an open<br />

beaker at room temperature. After 1-2 days a white crystalline product started to ap-<br />

110


pear. Evaporation was allowed to continue until the solvent level had approached the<br />

solid product, which was filtered <strong>of</strong>f <strong>and</strong> air-dried (1.6 g, yield 75 %). FTIR spectroscopy:<br />

1079(m), 1069(m), 1054(sh), 995(sh), 983(s), 963(sh), 885(m), 797(s), 728(sh), 681(sh),<br />

595(w), 522(w) cm −1 . Elemental analysis calcd. (found): N 2.9 (2.7), In 6.4 (6.5), P 1.2<br />

(1.1), W 61.7 (62.2), Cl 1.3 (1.0) %.<br />

31 P-NMR (D 2 O, 293 K): -11.4 ppm; 183 W-NMR (D 2 O, 293 K): (relative intensities<br />

in parenthesis) -98.3(1), -104.2(2), -128.6(2), -129.1(1), -133.1(2), -186.6(1)ppm. D,L-<br />

(NH 4 ) 9 Na 8 [In 3 Cl 2 (P 2 W 15 O 56 ) 2 ]·39H 2 O (NH4Na-15). A 0.39 g (1.76 mmols) sample <strong>of</strong><br />

InCl 3 was dissolved in 40 mL H 2 O followed by addition <strong>of</strong> 3.52 g (0.80 mmol)<br />

Na 12 [P 2 W 15 O 56 ]·24H 2 O. The pH <strong>of</strong> this solution was adjusted to 2 by addition <strong>of</strong> 4 M<br />

HCl <strong>and</strong> then it was heated to ∼80 ◦ C for 1 h. After cooling to room temperature the<br />

solution was filtered. Then it was layered with 0.1 M NH 4 Cl <strong>and</strong> allowed to evaporate in<br />

an open beaker at room temperature. After 1-2 days a white crystalline product started<br />

appear. Evaporation was allowed to continue until the solvent level had approached the<br />

solid product, which was filtered <strong>of</strong>f <strong>and</strong> air-dried (2.5 g, yield 69 %). FTIR spectroscopy:<br />

1092(s), 1059(m), 1017(w), 944(s), 917(s), 882(m), 831(sh), 810(sh), 764(s), 721(sh),<br />

641(sh), 598(sh), 524(w) cm −1 . Elemental analysis calcd. (found): Na 2.1 (2.1), N 1.4<br />

(1.2), In 3.9 (4.0), P 1.4 (1.3), W 62.0 (61.3), Cl 0.8 (0.9) %. 31 P-NMR (D 2 O, 293 K):<br />

(relative intensities in parenthesis) -7.1(1), -13.3(1) ppm.<br />

RbNa 3 [In 4 (H 2 O) 10 (β -AsW 9 O 32 OH) 2 ]·36H 2 O (RbNa-16). A 0.39 g (1.76 mmols) sample<br />

<strong>of</strong> InCl 3 was dissolved in 40 mL H 2 O followed by addition <strong>of</strong> 2.00 g (0.80 mmol) Na 9 [α-<br />

AsW 9 O 33 ]. The pH <strong>of</strong> this solution was adjusted to 2 by addition <strong>of</strong> 4 M HCl <strong>and</strong><br />

then it was heated to ∼80 ◦ C for 1 h. After cooling to room temperature the solution<br />

was filtered. Then it was layered with 0.1 M RbNO 3 <strong>and</strong> allowed to evaporate in an<br />

open beaker at room temperature. After 1-2 days a white crystalline product started<br />

to appear. Evaporation was allowed to continue until the solvent level had approached<br />

the solid product, which was filtered <strong>of</strong>f <strong>and</strong> air-dried. A total <strong>of</strong> 1.7 g (yield 72%) <strong>of</strong><br />

crystalline product was obtained. FTIR spectroscopy: 955(m), 881(sh), 823(s), 794(s),<br />

701(m), 614(sh), 510(sh), 478(w), 430(w) cm −1 . Elemental analysis calcd. (found): Rb<br />

1.4 (1.2), Na 1.2 (1.1), In 7.7 (7.5), As 2.5 (2.5), W 55.5 (55.0) %. 183 W-NMR (D 2 O, 293<br />

111


K): (relative intensities in parenthesis) -102.3(1), -109.6(2), -153.7(2), -171.3(2), -199.2(2)<br />

ppm.<br />

K 4 Na 2 [In 4 (H 2 O) 10 (β -SbW 9 O 33 ) 2 ]·30H 2 O (KNa-17). The synthesis was identical to that<br />

<strong>of</strong> (KNa-17), with the exception that 2.00 g (0.80 mmol) Na 9 [α-SbW 9 O 33 ] was used<br />

instead <strong>of</strong> Na 9 [α-AsW 9 O 33 ]. In addition, the solution was layered with 0.1 M KCl instead<br />

<strong>of</strong> RbNO 3 . In this case a total <strong>of</strong> 2.0 g (yield 82%) <strong>of</strong> crystalline product was obtained.<br />

FTIR spectroscopy: 949(m), 890(sh), 809(s), 693(m), 604(sh), 507(w), 467(w), 418(w)<br />

cm −1 . Elemental analysis calcd. (found): K 2.6 (2.8), Na 0.8 (0.5), In 7.7 (7.6), Sb<br />

4.1 (4.0), W 55.2 (55.9) %. We have also isolated the selenium(IV) <strong>and</strong> tellurium(IV)<br />

containing derivatives K 4 [In 4 (H 2 O) 10 (β-SeW 9 O 33 ) 2 ]·28H 2 O (K-18) <strong>and</strong> K 4 [In 4 (H 2 O) 10 (β<br />

-TeW 9 O 33 ) 2 ]·29H 2 O (K-19), respectively, as based on FTIR spectroscopy <strong>and</strong> elemental<br />

analysis. FTIR spectra for K-18: 954(m), 890(sh), 819(s), 765(s), 653(m), 504(w), 447(w)<br />

cm −1 . FTIR for K-19: 971(m), 889(m), 784(s), 749(m), 612(sh), 509(w), 423(w) cm −1 .<br />

Elemental analysis for (K-18) calcd.(found): K 2.7 (2.5), In 7.9 (7.6), Se 2.7 (2.8), W 56.8<br />

(57.5) %. Elemental analysis for K-19 calcd. (found): K 2.6 (2.7),In 7.7 (7.5), Te 4.3 (4.1),<br />

W 55.7 (56.2) %. Elemental analyses were performed by Kanti Labs Ltd. in Mississauga,<br />

Canada. FTIR spectra were recorded on a Nicolet Avatar FTIR spectrophotometer in a<br />

KBr pellet. All NMR spectra were recorded at room temperature with freshly synthesized<br />

solutions <strong>of</strong> 14-17 on a 400 MHz JEOL ECX instrument.<br />

X-ray Crystallography<br />

Single crystals <strong>of</strong> compounds NH4-14 <strong>and</strong> KNa-17 were mounted on a glass fiber for<br />

indexing <strong>and</strong> intensity data collection at 173 K for compound NH4-14 <strong>and</strong> KNa-17 <strong>and</strong><br />

200 K for compound NH4Na-15 <strong>and</strong> RbNa-16, respectively, on a Bruker D8 SMART<br />

APEX CCD single-crystal diffractometer using Mo K α radiation ( λ = 0.71073 Å). Direct<br />

methods were used to solve the structures <strong>and</strong> to locate the heavy atoms (SHELXS97).<br />

Then the remaining atoms were found from successive difference maps (SHELXL97).<br />

Routine Lorentz <strong>and</strong> polarization corrections were applied <strong>and</strong> an absorption correction<br />

was performed using the SADABS program [81]. Crystallographic data are summarized<br />

in Table 3.12.<br />

112


Table 3.12: Crystal Data <strong>and</strong> Structure Refinement for compounds NH4-14, NH4Na-15, RbNa-16<br />

<strong>and</strong> KNa-17<br />

Compounds NH4-14 NH4Na-15 RbNa-16 KNa-17<br />

fw 5361.5 8895.91 5959.8 5991.3<br />

space group P 2 1 /n (14) P ¯1 (2) P ¯1 (2) P ¯1 (2)<br />

a (Å) 17.5090(10) 13.0643(5) 12.8142(5) 12.2306(9)<br />

b (Å) 12.7361(7) 14.8901(6) 12.8672(5) 12.7622(10)<br />

c (Å) 18.7955(11) 19.8603(8) 16.1794(7) 16.1639(12)<br />

α ( ◦ ) 92.2920(10) 91.1370(10) 73.9890(10)<br />

β( ◦ ) 107.3030(10) 90.8680(10) 105.9450(10) 76.5550(10)<br />

γ ( ◦ ) 100.5630(10) 104.0980(10) 86.2130(10)<br />

vol. (Å 3 ) 4001.6(4) 3793.9(3) 2477.02(17) 2358.7(3)<br />

Z 2 1 1 1<br />

temp. ( ◦ C) -100 -73.3 -73.3 -100<br />

wavelength (Å) 0.71073 0.71073 0.71073 0.71073<br />

dcalc (Mg m −3 ) 4.404 3.851 3.918 4.167<br />

abs. coeff. (mm −1 ) 26.837 23.301 22.995 23.674<br />

R [I > 2 σ(I)] a 0.04 0.054 0.047 0.061<br />

R w (all data) b 0.075 0.133 0.119 0.113<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

3.12.3 Results <strong>and</strong> discussion<br />

The dimeric, chiral polyoxoanion [In 3 Cl 2 (B-α-PW 9 O 34 ) 2 ] 11− 14 is composed <strong>of</strong> two (Bα-PW<br />

9 O 34 ) units linked via three In(III) ions (see Figure 3.49). Polyanion 14 belongs<br />

to the well-known class <strong>of</strong> s<strong>and</strong>wich-type structures (Weakley type) <strong>and</strong> it represents the<br />

first indium-derivative [195]. Interestingly, only three indium centers are incorporated in<br />

14 <strong>and</strong> it is very unusual that one <strong>of</strong> the inner positions is vacant. Both outer positions<br />

are occupied by indium ions with a terminal chloro lig<strong>and</strong> each. Therefore polyanion 14<br />

exhibits C 2 point group symmetry (as opposed to C 2h for the tetrasubstituted transition<br />

metal analogs, e.g. [Mn 4 (H 2 O) 2 (B-α-PW 9 O 34 ) 2 ] 10− ) indicating that it is chiral. The only<br />

other trisubstituted analog <strong>of</strong> 14 known to date is [Ni 3 Na(H 2 O) 2 (B-α-PW 9 O 34 ) 2 ] 11− , but<br />

in this case the ‘vacant’ site is in an outer position <strong>and</strong> it is occupied by a sodium ion<br />

in the solid state [220]. Bond lengths <strong>and</strong> angles associated with the central indium-oxo<br />

fragment <strong>of</strong> 14 are shown in Table 3.13.<br />

The structure <strong>of</strong> 14 <strong>and</strong> its chiral nature somewhat resemble Tourné’s polyanion<br />

[WM 3 (H 2 O) 2 (B-α-XW 9 O 34 ) 2 ] 12− (X = M = Zn 2+ , Co 2+ ) [224]. Due to the fact that most<br />

polyoxoanions are not inherently chiral, Tourné’s species <strong>and</strong> especially its noble metal<br />

113


Fig. 3.49: Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong> [In 3 Cl 2 (PW 9 O 34 ) 2 ] 11− 14The red octahedra<br />

represent WO 6 , the blue tetrahedra represent PO 4 <strong>and</strong> the balls represent indium (green) <strong>and</strong><br />

chlorine (yellow).<br />

containing derivatives (e.g. [WZnRu III 2(H 2 O) 2 (B-α-ZnW 9 O 34 ) 2 ] 10− ) have been studied<br />

intensely for their potentially asymmetric catalytic activity [225]. It remains to be seen if<br />

chiral, transition metal substituted derivatives <strong>of</strong> 14 can be prepared. Polyanion 14 was<br />

synthesized by interaction <strong>of</strong> In 3+ ions with the trilacunary precursor [B-α-PW 9 O 34 ] 9− in<br />

aqueous, acidic medium. We discovered that the trisubstituted 14 is obtained even in the<br />

presence <strong>of</strong> excess In 3+ ions. Additional studies are needed in order to explain why the<br />

tri- rather than the tetrasubstituted product is formed <strong>and</strong> also, why the vacancy is in an<br />

interior site.<br />

Solution NMR<br />

As polyanion 14 is diamagnetic we performed 183 W <strong>and</strong> 31 P NMR studies in solution.<br />

The 31 P NMR spectrum <strong>of</strong> 14 exhibits one signal at -11.4 ppm <strong>and</strong> the 183 W NMR spectrum<br />

shows 6 peaks at -98.3(1), -104.2(2), -128.6(2), -129.1(1), -133.1(2) <strong>and</strong> -186.6(1)<br />

ppm, respectively (see Figure 3.50). Initially this six-line pattern with intensity ratios<br />

2:2:2:1:1:1 may seem unexpected, but it can be rationalized as follows. The hypotheti-<br />

114


Table 3.13: Selected bond lengths <strong>and</strong> angles in polyanions 14 <strong>and</strong> 15<br />

14 15<br />

In1-O4I1 2.061(8) In1-O10I 2.087(9)<br />

In1-O7A 2.082(8) In1-O13A 2.115(9)<br />

In1-O5IN 2.133(7) In1-O15I 2.136(9)<br />

In1-O6I2 2.146(7) In1-O14A 2.165(9)<br />

In1-O4P 2.320(8) In1-O4P2 2.248(8)<br />

In1-O4P 2.323(7) In1-O4P2 2.274(8)<br />

In2-O8A 2.104(7) In2-O11I 2.120(9)<br />

In2-O9A 2.118(7) In2-O14A 2.126(8)<br />

In2-O6I2 2.139(7) In2-O12I 2.126(9)<br />

In2-O5IN 2.147(7) In2-O15I 2.134(8)<br />

In2-O4P 2.287(7) In2-O4P2 2.224(9)<br />

In2-Cl1 2.418(3) In2-Cl1 2.394(6)<br />

O4I1-In1-O7A 95.7(3) O10I-In1-O13A 94.5(4)<br />

O4I1-In1-O5IN 101.7(3) O10I-In1-O15I 98.6(3)<br />

O7A-In1-O5IN 87.6(3) O13A-In1-O15I 87.3(3)<br />

O4I1-In1-O6I2 86.1(3) O10I-In1-O14A 87.6(3)<br />

O7A-In1-O6I2 100.2(3) O13A-In1-O14A 99.7(3)<br />

O5IN-In1-O6I2 168.5(3) O15I-In1-O14A 170.3(3)<br />

O4I1-In1-O4P 175.8(3) O10I-In1-O4P2 175.2(3)<br />

O7A-In1-O4P 88.1(3) O13A-In1-O4P2 90.3(3)<br />

O5IN-In1-O4P 80.3(3) O15I-In1-O4P2 81.9(3)<br />

O6I2-In1-O4P 91.4(3) O14A-In1-O4P2 91.3(3)<br />

O4I1-In1-O4P 88.1(3) O10I-In1-O4P2 88.9(3)<br />

O7A-In1-O4P 176.2(3) O13A-In1-O4P2 176.5(3)<br />

O5IN-In1-O4P 91.9(3) O15I-In1-O4P2 91.3(3)<br />

O6I2-In1-O4P 79.7(3) O14A-In1-O4P2 81.4(3)<br />

O4P-In1-O4P 88.1(3) O4P2-In1-O4P2 86.3(3)<br />

O8A-In2-O9A 94.4(3) O11I-In2-O14A 89.8(3)<br />

O8A-In2-O6I2 165.3(3) O11I-In2-O12I 91.6(3)<br />

O9A-In2-O6I2 88.7(3) O14A-In2-O12I 169.7(3)<br />

O8A-In2-O5IN 88.7(3) O11I-In2-O15I 169.5(3)<br />

O9A-In2-O5IN 167.4(3) O14A-In2-O15I 87.2(3)<br />

O6I2 In2 O5IN 85.4(3) O12I-In2-O15I 89.6(3)<br />

O8A In2 O4P 85.1(3) O11I-In2-O4P2 87.1(3)<br />

O9A In2 O4P 87.3(3) O14A-In2-O4P2 83.4(3)<br />

O6I2 In2 O4P 80.7(3) O12I-In2-O4P2 86.5(3)<br />

O5IN In2 O4P 80.8(3) O15I-In2-O4P2 82.6(3)<br />

O8A In2 Cl1 92.4(2) O11I-In2-Cl1 89.0(3)<br />

O9A In2 Cl1 89.30(19) O14A-In2-Cl1 98.7(3)<br />

O6I2 In2 Cl1 101.96(19) O12I-In2-Cl1 91.5(3)<br />

O5IN In2 Cl1 102.80(19) O15I-In2-Cl1 101.4(3)<br />

O4P In2 Cl1 175.6(2) O4P2-In2-Cl1 175.5(3)<br />

cal, tetrasubstitued derivative [In 4 Cl 2 4(B-α-PW 9 O 34 ) 2 ] 8− has C 2h symmetry <strong>and</strong> would<br />

therefore be expected to show five lines in 183 W-NMR with intensity ratios 2:2:2:2:1. Indeed,<br />

we have obtained exactly this pattern for the tetrasubstituted, isostructural Zn(II)<br />

<strong>and</strong> Cd(II) derivatives [M 4 (H 2 O) 2 (B-α-GeW 9 O 34 ) 2 ] 12− (M = Zn 2+ , Cd 2+ ).[195] Removal<br />

<strong>of</strong> one <strong>of</strong> the inner indium atoms from [In 4 Cl 2 (B-α-PW 9 O 34 ) 2 ] 8− results in 14 with C 2<br />

point group symmetry. Consequently, 9 peaks are expected in 183 W NMR. However, close<br />

inspection <strong>of</strong> the structure <strong>of</strong> 14 (see Figure 3.50) indicates that the three tungsten atoms<br />

in the Keggin cap <strong>and</strong> four <strong>of</strong> the six tungsten atoms in the Keggin belt are not likely<br />

to be affected significantly by removal <strong>of</strong> an inner indium atom as they do not share oxo<br />

lig<strong>and</strong>s with that indium atom. The inner indium atoms are coordinated to only three<br />

115


Fig. 3.50: 183 W NMR spectrum <strong>of</strong> [In 3 Cl 2 (B-α-PW 9 O 34 ) 2 ] 11−<br />

oxo groups <strong>of</strong> each Keggin half-unit, involving two belt tungsten atoms (W6, W7) <strong>and</strong><br />

the phosphorus hetero atom. The six-line 183 W NMR spectrum <strong>of</strong> 14 with intensity ratios<br />

2:2:2:1:1:1 indicates that only one <strong>of</strong> these two belt tungsten atoms is significantly<br />

affected by removal <strong>of</strong> an inner indium atom. Clearly, this must be W7 which experiences<br />

a major change in its coordination sphere as a result <strong>of</strong> In atom removal. Before, it has<br />

only one terminal oxo lig<strong>and</strong> <strong>and</strong> after it has two terminal, cis-related oxo lig<strong>and</strong>s. The<br />

dimeric polyanion [In 3 Cl 2 (P 2 W 15 O 56 ) 2 ] 17− 15 represents the Wells-Dawson analog <strong>of</strong> 10<br />

(see Figure 3.51).<br />

Fig. 3.51: Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong> [In 3 Cl 2 (P 2 W 15 O 56 ) 2 ] 17− 15. The color<br />

code is the same as in Figure 3.40<br />

116


Polyanion 15 contains three indium centers s<strong>and</strong>wiched in between two trilacunary<br />

(P 2 W 15 O 56 ) units <strong>and</strong> one <strong>of</strong> the inner positions is vacant. In analogy to 15 the terminal<br />

lig<strong>and</strong>s <strong>of</strong> the outer indium atoms in 15 are also chloride ions. As a result, polyanion<br />

15 exhibits also C 2 point group symmetry <strong>and</strong> is therefore chiral. The bond lengths<br />

<strong>and</strong> angles associated with the central indium-oxo fragment <strong>of</strong> 15 are shown in Table<br />

3.13. Some other trisubstituted analogs <strong>of</strong> 15 exist, but in all cases the ‘vacant’<br />

site is in an outer position <strong>and</strong> it is occupied by a sodium ion in the solid state (e.g.<br />

[Fe 2 (FeOH 2 )(NaOH 2 )(P 2 W 15 O 56 ) 2 ] 14− ) [219]. Polyanion 15 was synthesized by interaction<br />

<strong>of</strong> In 3+ ions with the trilacunary precursor [P 2 W 15 O 56 ] 12− in aqueous, acidic medium.<br />

We discovered that the trisubstituted 15 is formed even in the presence <strong>of</strong> excess In 3+<br />

ions. As polyanion 15 is diamagnetic we performed 183 W <strong>and</strong> 31 P NMR studies in solution.<br />

The latter exhibits two signals <strong>of</strong> equal intensity at -7.1 <strong>and</strong> -13.3 ppm, respectively,<br />

which is in agreement with the solid state structure. The two Wells-Dawson fragments<br />

in 15 are equivalent <strong>and</strong> they contain two P atoms each. The signal at -7.1 ppm can be<br />

assigned to the P-atom closer to the indium-core, whereas the signal at -13.3 ppm corresponds<br />

to the more distant P-atom. The 183 W NMR spectrum <strong>of</strong> 15 is expected to show<br />

around 9-10 peaks (based on the arguments used above for 14) <strong>and</strong> we see this number<br />

<strong>of</strong> signals between -80 <strong>and</strong> -240 ppm. However, the poor signal-to-noise ratio <strong>of</strong> our spectrum<br />

did not allow us to identify all peaks unequivocally. Although polyanions 14 <strong>and</strong><br />

15 are chiral they have crystallized as a racemic mixture. Our crystallographic studies<br />

have revealed that both inner indium positions are occupied, but only with occupancy<br />

factors <strong>of</strong> 0.5 each. This means that individual molecules <strong>of</strong> 14 <strong>and</strong> 15 contain only one<br />

indium atom in one <strong>of</strong> the two possible inner positions. As the synthesis procedure <strong>of</strong> 14<br />

<strong>and</strong> 15 is not stereoselective, both enantiomers are formed in equal amounts. Apparently<br />

the crystallization process is also not stereoselective, which is not a surprise as the chiral<br />

site in 14 <strong>and</strong> 15 is somewhat hidden by the two Keggin caps <strong>and</strong> as a result is not<br />

expected to influence polyanion packing significantly. Therefore, both enantiomers orient<br />

r<strong>and</strong>omly within the solid state lattices <strong>of</strong> NH4-14 <strong>and</strong> NH4Na-15 leading to the crystallographic<br />

observations described above.The indium-substituted, lone pair containing<br />

polyoxoanions [In 4 (H 2 O) 10 (β -AsW 9 O 32 OH) 2 ] 4− (16) <strong>and</strong> [In 4 (H 2 O) 10 (β-SbW 9 O 33 ) 2 ] 6−<br />

117


(17) are isostructural. They consist <strong>of</strong> two β-XW 9 (X = As III , Sb III ) Keggin moieties<br />

linked by four In 3+ ions resulting in a structure with idealized C 2h symmetry (see Figure<br />

3.52).<br />

Fig. 3.52: Combined polyhedral/ball <strong>and</strong> stick representation <strong>of</strong><br />

[In 4 (H 2 O) 10 (β-AsW 9 O 32 OH) 2 ] 4− (16) <strong>and</strong> [In 4 (H 2 O) 10 (β-SbW 9 O 33 ) 2 ] 6− (17). The WO 6 octahedra are<br />

shown in red <strong>and</strong> the balls represent indium (green), arsenic/antimony (blue) <strong>and</strong> water molecules (red).<br />

The four In 3+ ions consist <strong>of</strong> two inequivalent pairs, the inner two In 3+ ions have<br />

two terminal H 2 O lig<strong>and</strong>s <strong>and</strong> the outer two In 3+ ions have three terminal H 2 O lig<strong>and</strong>s.<br />

Polyanions 16 <strong>and</strong> 17 were synthesized in aqueous acidic medium (pH 2) from interaction<br />

<strong>of</strong> In 3+ ions with the lacunary Keggin precursors [α-AsW 9 O 33 ] 9− <strong>and</strong> [ α-SbW 9 O 33 ] 9− ,<br />

respectively. Therefore the mechanism <strong>of</strong> formation <strong>of</strong> 16 <strong>and</strong> 17 involves insertion,<br />

isomerization (α → β) <strong>and</strong> dimerization. The bond lengths <strong>and</strong> angles associated with<br />

the central indium-oxo fragments <strong>of</strong> 16 <strong>and</strong> 17 are shown in Table 3.14<br />

183 W NMR on freshly synthesized solutions <strong>of</strong> RbNa-16 resulted in five peaks (relative<br />

intensities in parenthesis) at -102.3(1), -109.6(2), -153.7(2), -171.3(2) <strong>and</strong> -199.2(2) ppm,<br />

respectively (see Figure 3.53).<br />

This result indicates that in solution a species with C 2h symmetry is present, which<br />

is in complete agreement with the solid state structure <strong>of</strong> RbNa-16. Furthermore, the<br />

NMR spectrum does not change even after several weeks, indicating the high stability<br />

118


Table 3.14: Selected bond lengths <strong>and</strong> angles in polyanions 16 <strong>and</strong> 17<br />

16 17<br />

In(1)-O(3I1) 2.069(8) In(1)-O(3IN) 2.083(11)<br />

In(1)-O(7I1) 2.073(7) In(1)-O(6IN) 2.109(10)<br />

In(1)-O(6T)#1 2.107(7) In(1)-O(9B)#2 2.120(10)<br />

In(1)-O(6I1) 2.109(7) In(1)-O(1I1) 2.146(10)<br />

In(1)-O(2WI) 2.192(8) In(1)-O(3I1) 2.164(12)<br />

In(1)-O(1WI) 2.206(8) In(1)-O(2I1) 2.193(12)<br />

In(2)-O(4I2) 2.084(7) In(2)-O(1A)#2 2.059(10)<br />

In(2)-O(5I2) 2.112(7) In(2)-O(5IN) 2.072(11)<br />

In(2)-O(4WI) 2.153(8) In(2)-O(9IN) 2.110(11)<br />

In(2)-O(5WI) 2.156(8) In(2)-O(9A)#2 2.117(10)<br />

In(2)-O(6I2) 2.160(8) In(2)-O(2I2) 2.170(12)<br />

In(2)-O(3WI) 2.206(9) In(2)-O(1I2) ) 2.174(11)<br />

As(1)-O(3AS) 1.794(7) Sb(1)-O(2SB) 1.989(9)<br />

As(1)-O(1AS) 1.810(7) Sb(1)-O(1SB) 2.014(10)<br />

As(1)-O(2AS) 1.814(7) Sb(1)-O(3SB) 2.014(10)<br />

O(3I1)-In(1)-O(7I1) 173.8(3) O(3IN)-In(1)-O(6IN) 86.1(4)<br />

O(3I1)-In(1)-O(6T)#1 89.4(3) O(3IN)-In(1)-O(9B)#2 99.4(4)<br />

O(7I1)-In(1)-O(6T)#1 95.0(3) O(6IN)-In(1)-O(9B)#2 95.6(4)<br />

O(3I1)-In(1)-O(6I1) 92.3(3) O(3IN)-In(1)-O(1I1) 95.3(4)<br />

O(7I1)-In(1)-O(6I1) 91.9(3) O(6IN)-In(1)-O(1I1) 177.1(4)<br />

O(6T)#1-In(1)-O(6I1) 92.4(3) O(9B)#2-In(1)-O(1I1) 86.7(4)<br />

O(3I1)-In(1)-O(2WI) 84.7(3) O(3IN)-In(1)-O(3I1) 174.4(4)<br />

O(7I1)-In(1)-O(2WI) 90.4(3) O(6IN)-In(1)-O(3I1) 93.7(4)<br />

O(6T)#1-In(1)-O(2WI) 172.0(3) O(9B)#2-In(1)-O(3I1) 86.2(4)<br />

O(6I1)-In(1)-O(2WI) 93.2(3) O(1I1)-In(1)-O(3I1) 84.6(4)<br />

O(3I1)-In(1)-O(1WI) 90.4(3) O(3IN)-In(1)-O(2I1) 91.9(4)<br />

O(7I1)-In(1)-O(1WI) 85.5(3) O(6IN)-In(1)-O(2I1) 93.2(4)<br />

O(6T)#1-In(1)-O(1WI) 86.8(3) O(9B)#2-In(1)-O(2I1) 166.1(4)<br />

O(6I1)-In(1)-O(1WI) 177.2(3) O(1I1)-In(1)-O(2I1) 84.2(4)<br />

O(2WI)-In(1)-O(1WI) 87.9(3) O(3I1)-In(1)-O(2I1) 82.5(5)<br />

O(4I2)-In(2)-O(5I2) 90.9(3) O(1A)#2-In(2)-O(5IN) 177.4(4)<br />

O(4I2)-In(2)-O(4WI) 91.6(3) O(1A)#2-In(2)-O(9IN) 92.2(4)<br />

O(5I2)-In(2)-O(4WI) 174.6(3) O(5IN)-In(2)-O(9IN) 90.0(4)<br />

O(4I2)-In(2)-O(5WI) 175.2(3) O(1A)#2-In(2)-O(9A)#2 89.6(4)<br />

O(5I2)-In(2)-O(5WI) 90.0(3) O(5IN)-In(2)-O(9A)#2 89.0(4)<br />

O(4WI)-In(2)-O(5WI) 87.2(3) O(9IN)-In(2)-O(9A)#2 88.8(4)<br />

O(4I2)-In(2)-O(6I2) 96.0(3) O(1A)#2-In(2)-O(2I2) 94.3(4)<br />

O(5I2)-In(2)-O(6I2) 98.8(3) O(5IN)-In(2)-O(2I2) 83.5(4)<br />

O(4WI)-In(2)-O(6I2) 85.7(3) O(9IN)-In(2)-O(2I2) 173.5(4)<br />

O(5WI)-In(2)-O(6I2) 88.5(3) O(9A)#2-In(2)-O(2I2) 90.3(4)<br />

O(4I2)-In(2)-O(3WI) 93.3(3) O(1A)#2-In(2)-O(1I2) 88.1(4)<br />

O(5I2)-In(2)-O(3WI) 90.9(3) O(5IN)-In(2)-O(1I2) 93.2(4)<br />

O(4WI)-In(2)-O(3WI) 84.1(3) O(9IN)-In(2)-O(1I2) 94.4(4)<br />

O(5WI)-In(2)-O(3WI) 82.0(3) O(9A)#2-In(2)-O(1I2) 176.2(4)<br />

O(6I2)-In(2)-O(3WI) 166.4(3) O(2I2)-In(2)-O(1I2) 86.8(4)<br />

O(3AS)-As(1)-O(1AS) 98.9(3) O(2SB)-Sb(1)-O(1SB) 90.7(4)<br />

O(3AS)-As(1)-O(2AS) 96.8(3) O(2SB)-Sb(1)-O(3SB) 93.1(4)<br />

O(1AS)-As(1)-O(2AS) 98.1(3) O(1SB)-Sb(1)-O(3SB) 93.3(4)<br />

<strong>of</strong> RbNa-16 in aqueous solution. The 183 W NMR behavior <strong>of</strong> the antimony derivative<br />

KNa-17 is apparently more complex <strong>and</strong> requires additional studies. Our preliminary<br />

results on freshly synthesized solutions <strong>of</strong> KNa-17 resulted in six peaks at -80.9, -99.5,<br />

-117.5, -133.9, -160.6 <strong>and</strong> -176.7 ppm. Very recently Krebs et al. reported on the solid<br />

state structures <strong>of</strong> the closely related compounds Na 5n H 2n [(In(H 2 O) 2 ) 1.5 (Na(H 2 O) 2 ) 0.5<br />

(In(H 2 O) 2 ) 2 (SbW 9 O 33 ) 2 ] n·28H 2 O <strong>and</strong> Na 2 K 2 H 2 [(In(H 2 O) 3 ) 2 (In(H 2 O) 2 ) 2 (AsW 9 O 33 ) 2 ]·37H 2 O<br />

[30]. Although the solid state structures <strong>of</strong> ‘Krebs’ compounds <strong>and</strong> ours are different,<br />

119


Fig. 3.53: 183 W NMR spectrum <strong>of</strong> [In 4 (H 2 O) 10 (β-AsW 9 O 32 OH) 2 ] 4− at 293 K<br />

the latter contains a polyanion identical to 16. On the other h<strong>and</strong>, the antimony(III)-<br />

containing polyanion <strong>of</strong> Krebs et al. is different from 17 as the outer indium positions<br />

are partially (25%) occupied by sodium ions resulting in a polymeric solid state structure.<br />

Bond valence sum (BVS) calculations on 16 <strong>and</strong> 17 resulted in the conclusion that all<br />

terminal lig<strong>and</strong>s <strong>of</strong> the indium atoms are water molecules [85]. For 17 we could not identify<br />

any other protonation sites, but for 16 we noticed that two In-O-W bridging oxygens<br />

(O5I2, O5I2’) are actually hydroxo groups. This results in a charge <strong>of</strong> -4 for 16 <strong>and</strong> -6 for<br />

17, which is somewhat surprising as both species were synthesized at exactly the same<br />

pH. It is <strong>of</strong> interest to compare the synthetic conditions for the preparation <strong>of</strong> ‘Krebs<br />

<strong>and</strong> our compounds. Besides using a higher concentration <strong>of</strong> the reagents (by about a<br />

factor <strong>of</strong> 2), Krebs <strong>and</strong> coworkers also performed their syntheses at pH 6.5-7, whereas we<br />

worked in much more acidic medium (pH 2.0). We used such acidic conditions because<br />

former work <strong>of</strong> Krebs et al. <strong>and</strong> also our own has demonstrated that [α-XW 9 O 33 ] 9− <strong>and</strong><br />

[β-XW 9 O 33 ] 9− (X = As III , Sb III ) are in equilibrium in aqueous solution [33, 39, 69, 83].<br />

The former dominates in neutral medium whereas the latter is present in acidic solution.<br />

Our synthetic conditions <strong>of</strong> 16 <strong>and</strong> 17 are in full agreement with this, as both species<br />

are formed easily in acidic medium. The recent results <strong>of</strong> Krebs et al. indicate that 16<br />

is also stable in the neutral pH range, but this is apparently not the case for 17. This<br />

observation combined with the different degree <strong>of</strong> protonation for 16 <strong>and</strong> 17 deserves<br />

attention <strong>and</strong> is further supported by our 183 W-NMR studies (see 3.53). We have already<br />

observed previously that the products resulting from interaction <strong>of</strong> transition metal<br />

ions with [α-AsW 9 O 33 ] 9− <strong>and</strong> [α-SbW 9 O 33 ] 9− are not solely determined by the pH, but<br />

120


also by the type <strong>of</strong> transition metal ion <strong>and</strong> its preferred coordination geometry. For<br />

example, we were able to synthesize the Fe 3+ analogues <strong>of</strong> 16 <strong>and</strong> 17, but not the Cu 2+<br />

<strong>and</strong> Zn 2+ derivatives.[39, 83] Reaction <strong>of</strong> the latter metal ions with [α-XW 9 O 33 ] 9− (X =<br />

As III , Sb III ) resulted in the polyanions [M 3 (H 2 O) 3 (α-XW 9 O 33 ) 2 ] 12− (M = Cu 2+ , Zn 2+ ;<br />

X = As III , Sb III ), no matter what the pH. The Cu 2+ <strong>and</strong> Zn 2+ ions have a square pyramidal<br />

coordination geometry in this structure, whereas the coordination geometry <strong>of</strong> all<br />

four In 3+ ions in 16 <strong>and</strong> 17 is octahedral. Interestingly, Cr 3+ was the only other firstrow<br />

transition metal ion besides Fe 3+ for which we could synthesize the tetra-substituted<br />

structure with As III <strong>and</strong> Sb III as hetero atoms [83]. It becomes apparent, that all three<br />

ions which allow formation <strong>of</strong> this structural type are trivalent (In 3+ , Fe 3+ , Cr 3+ ). However,<br />

by using Se IV <strong>and</strong> Te IV as hetero atoms we could also incorporate divalent transiton<br />

metal ions (e.g. Mn 2+ , Co 2+ , Ni 2+ , Zn 2+ , Cd 2+ , Hg 2+ ) [83]. These results indicate that<br />

the tetra-substituted Krebs-type structure is most stable within the charge limits <strong>of</strong> -4<br />

<strong>and</strong> -8. The differences <strong>of</strong> 12A <strong>and</strong> 12B in their solution <strong>and</strong> solid state properties (e.g.<br />

protonation, charge, pH-dependent stability) led us to check if the different hetero atoms<br />

(As III , Sb III ) perhaps initiate small structural changes (e.g. bond lengths, angles) in the<br />

polyanions. As expected, the As-O bond lengths are shorter than the Sb-O bonds (see<br />

Table 3.13.) <strong>and</strong> the observed hetero atom separations X···X within 16 (As···As = 6.26<br />

Å) <strong>and</strong> 17 (Sb···Sb = 5.87 Å) are fully consistent with this. On the other h<strong>and</strong>, the type<br />

<strong>of</strong> hetero group has no significant effect on the W-O <strong>and</strong> In-O bond lengths in 16 <strong>and</strong> 17<br />

(see Table 3.13). Therefore we conclude that there are no obvious structural differences in<br />

16 <strong>and</strong> 17. Furthermore, there is no significant lone-pair/lone-pair repulsion involving the<br />

lone pairs <strong>of</strong> the two hetero atoms. Both conclusions are fully consistent with our observations<br />

on the iron(III)-substituted derivatives <strong>of</strong> 16 <strong>and</strong> 17.[83] In fact, the hetero atom<br />

separations in 16 <strong>and</strong> 17 (see above) are even larger than in [Fe 4 (H 2 O) 10 (β-AsW 9 O 33 ) 2 ] 6−<br />

(As···As = 6.03 Å) <strong>and</strong> in [Fe 4 (H 2 O) 10 (β-SbW 9 O 33 ) 2 ] 6− (Sb···Sb = 5.68 Å), which reflects<br />

the larger In-O bond lengths compared to Fe-O (the ionic radii <strong>of</strong> In 3+ <strong>and</strong> Fe 3+ are 0.94<br />

Å <strong>and</strong> 0.79 Å, respectively).<br />

121


3.12.4 Conclusion<br />

We have synthesized <strong>and</strong> structurally characterized four novel indium(III)-substituted,<br />

s<strong>and</strong>wich-type polyoxoanions. The tri-indium substituted tungstophosphates [In 3 Cl 2 (Bα-PW<br />

9 O 34 ) 2 ] 11− 14 <strong>and</strong> [In 3 Cl 2 (P 2 W 15 O 56 ) 2 ] 17− 15 belong to the class <strong>of</strong> Weakley-type<br />

s<strong>and</strong>wich polyanions <strong>and</strong> they represent the first indium-containing derivatives. Polyanions<br />

14 <strong>and</strong> 15 are also stable in solution as indicated by a one-line (1) <strong>and</strong> a two-line<br />

(2) 31 P NMR pattern. The most important structural feature <strong>of</strong> 14 <strong>and</strong> 15 is the fact<br />

that they are chiral (C 2 symmetry), which results from a vacancy in one <strong>of</strong> the inner<br />

indium positions. Therefore 14 <strong>and</strong> 15 are <strong>of</strong> interest for catalytic <strong>and</strong> medicinal applications.<br />

Currently we investigate if chiral, transition metal substituted derivatives<br />

<strong>of</strong> 14 <strong>and</strong> 15 can be formed. The lone pair containing, tetra-indium substituted polyoxotungstates<br />

[In 4 (H 2 O) 10 (β-AsW 9 O 32 OH) 2 ] 4− 16 <strong>and</strong> [In 4 (H 2 O) 10 (β-SbW 9 O 33 ) 2 ] 6− 17<br />

belong to the class <strong>of</strong> Krebs-type s<strong>and</strong>wich polyanions. Polyanions 16 <strong>and</strong> 17 consist <strong>of</strong><br />

two B-β-(XW 9 O 33 ) (X = As III , Sb III ) Keggin moieties linked via four octahedral indium<br />

atoms leading to a dimeric structure with nominal C 2h symmetry. Polyanion 16 is also<br />

stable in solution as indicated by a five-line pattern (1:2:2:2:2) in 183 W-NMR. We have<br />

also synthesized the Se(IV) <strong>and</strong> Te(IV) analogues <strong>of</strong> 16 <strong>and</strong> 17.<br />

122


Bibliography<br />

[1] M. T. Pope. Heteropoly <strong>and</strong> Isopoly Oxometalates. Springer-Verlag, Berlin, 1983.<br />

[2] C. L. Hill. Chem. Rev., 98(1):1–390, 1998.<br />

[3] M. T. Pope <strong>and</strong> A. Müller. Polyoxometalates: from Platonic Solids to Anti-Retroviral Activity.<br />

Kluwer: Dordrecht, The Netherl<strong>and</strong>s, 1994.<br />

[4] M. T. Pope <strong>and</strong> A. Müller. Polyoxometalate Chemistry: From Topology via Self-Assembly to<br />

Applications. Kluwer: Dordrecht, The Netherl<strong>and</strong>s, 2001.<br />

[5] T. Yamase <strong>and</strong> M. T. Pope. Polyoxometalate Chemistry for Nano-Composite Design. Kluwer:<br />

Dordrecht, The Netherl<strong>and</strong>s, 2002.<br />

[6] M. T. Pope <strong>and</strong> A. Müller. Angew. Chem. Int. Ed., 30(1):34–48, 1991.<br />

[7] J. J. Berzelius. Poggendorfs Ann. Phys. Chem., 6:369, 1826.<br />

[8] C. Marignac. Ann. Chim. Phys., 3:1, 1864.<br />

[9] L.C. Pauling. J. Am.Chem. Soc., 51:2868, 1929.<br />

[10] J. F. Keggin. Nature., 131:908, 1933.<br />

[11] J. F. Keggin. Proc. Roy. Soc. A, 144:75, 1934.<br />

[12] T. R. Halbert, T. C. Ho, E. I. Stiefel, R. R. Chianelli, <strong>and</strong> M. J. Daage. J. Catal., 130:116, 1991.<br />

[13] E. Cadot, V. Béreau, B. Marg, S. Halut, <strong>and</strong> F. Sécheresse. Inorg. Chem., 95:3099, 1996.<br />

[14] E. Cadot, V. Béreau, <strong>and</strong> F. Sécheresse. Inorg. Chim. Acta., 252:101, 1996.<br />

[15] V. Béreau, E. Cadot, H. Bögge, A. Müller, <strong>and</strong> F. Sécheresse. Inorg. Chem., 38:5803, 1999.<br />

[16] E. Cadot, B. Salignac, S. Halut, <strong>and</strong> F. Sécheresse. Angew. Chem. Int. Ed., 37(5):611–612, 1998.<br />

[17] A. Bino, M. Ardon, D. Lee, B. Spingler, <strong>and</strong> S. J. Lippard. J. Am. Chem. Soc., 124:4578, 2002.<br />

[18] R. J. Errington, R. L. Wingad, W. Clegg, <strong>and</strong> M. R. J. Elsegood. Angew. Chem. Int. Ed., 39:3884,<br />

2000.<br />

[19] C. J. Clark <strong>and</strong> D. Hall. Acta Crystallogr. B., 32:1454, 1976.<br />

[20] M.-M. Rohmer, M. Bénard, J. P. Blaudeau, J. M. Maestre, <strong>and</strong> J. M. Poblet. Coord. Chem. Rev.,<br />

178-180:1019, 1998.<br />

[21] D. L. Kepert. Inorg. Chem., 8:1556, 1962.<br />

[22] A. Tézé <strong>and</strong> G. Hervé. J. Inorg. Nucl. Chem., 39:2151, 1977.<br />

[23] M. T. Pope. J. Inorg. Nucl. Chem., 15:2068, 1976.<br />

[24] V. W. Day <strong>and</strong> W. G. Klemperer. <strong>Science</strong>., 228:533, 1985.<br />

[25] A. Müller. Nature., 352:115, 1991.<br />

[26] M. T Pope. Nature., 355:27, 1992.<br />

[27] J. A. Jansen <strong>and</strong> S. H. Singh, D. J. Wang. Chem. Mater., 6:146, 1994.<br />

[28] J. M. Maestre, X. López, C. Bo, N. Casañ-Pastor, <strong>and</strong> J. M. Poblet. J. Am. Chem. Soc., 123:3749,<br />

2001.<br />

[29] X. López, J. M. Maestre, C. Bo, <strong>and</strong> J. M. Poblet. J. Am. Chem. Soc., 123(39):9571–9576, 2001.<br />

[30] I. Lindqvist. Acta Cryst., 5:247, 1952.<br />

[31] J.S. Anderson. Nature., 140:850, 1937.<br />

[32] B. Dawson. Acta Cryst., 6:113, 1953.<br />

[33] U. Kortz, M. G. Savelieff, B. S. Bassil, <strong>and</strong> M. H. Dickman. Angew. Chem. Int. Ed., 40:3384–3386,<br />

2001.<br />

[34] Loose I. Pohlmann H. Krebs B. Bösing, M. Chem. Eur. J., 3:1232, 1997.<br />

[35] Y. Jeannin <strong>and</strong> J. Martin-Frère. J. Am. Chem. Soc., 103:1664, 1981.<br />

[36] J. M. Clemente-Juan, E. Coronado, J. R. Galán-Mascarós, <strong>and</strong> C. J. Gómez-Garca. Inorg. Chem.,<br />

38:55, 1999.<br />

123


[37] D. Gatteschi, O. Kahn, J. Miller, <strong>and</strong> F. Palacio. Magnetic Molecular Materials. NATO ASI Series<br />

<strong>and</strong> Kluwer: Dordrecht, The Netherl<strong>and</strong>s, 1991.<br />

[38] C. L. Hill <strong>and</strong> Prosser McCartha C. M. Coord. Chem. Rev., 143:407, 1995.<br />

[39] U. Kortz, N. K. Al-Kassem, M. G. Savelieff, N. A. Al Kadi, <strong>and</strong> M. Sadakane. Inorg. Chem.,<br />

40:4742, 2001.<br />

[40] M. Inoue <strong>and</strong> T. Yamase. Bull. Chem. Soc. Jpn., 68:3055, 1995.<br />

[41] T. Yamase, H. Fujita, <strong>and</strong> K. Fukushima. Inorg. Chim. Acta., L15:151, 1998.<br />

[42] S. G. Sarafianos, U. Kortz, M. T. Pope, <strong>and</strong> M. J. Modak. Biochem. J, 319:619–626, 1996.<br />

[43] C. R. Mayer, S. Neveu, F. Secheresse, <strong>and</strong> V. Cabuil. J. Colloid Interface Sci., 273:350–355, 2004.<br />

[44] F. B. Xin <strong>and</strong> M. T. Pope. Organomet, 13(12):4881, 1994.<br />

[45] F. B. Xin, M. T. Pope, G. J. Long, <strong>and</strong> U. Russo. Inorg. Chem., 35(5):1207, 1996.<br />

[46] F. B. Xin <strong>and</strong> M. T. Pope. Inorg. Chem., 35(19):5693, 1996.<br />

[47] G. Sazani, M. H. Dickman, <strong>and</strong> M. T. Pope. Inorg. Chem., 39:939, 2000.<br />

[48] Q. H. Yang, H. C. Dai, <strong>and</strong> J. F. Liu. Transit. Metal Chem., 23:93, 1998.<br />

[49] X. H. Wang, H. C. Dai, <strong>and</strong> J. F. Liu. Polyhedron, 18:2293, 1999.<br />

[50] X. H. Wang, H. C. Dai, <strong>and</strong> J. F. Liu. Transit. Metal Chem., 24:600, 1999.<br />

[51] X. H. Wang <strong>and</strong> J. F. Liu. J. Coord. Chem., 51:73, 2000.<br />

[52] X. H. Wang, J. T. Liu, R. C. Zhang, B. Li, <strong>and</strong> J. F. Liu. Main Group Met. Chem., 25:535, 2002.<br />

[53] S. Bareyt, S. Piligkos, B. Hasenknopf, P. Gouzerh, E. Lacote, S. Thorimbert, <strong>and</strong> M. Malacria.<br />

Angew. Chem. Int. Ed., 42(29):3404–3406, 2003.<br />

[54] G. Sazani <strong>and</strong> M. T. Pope. Dalton Trans., page 1989, 2004.<br />

[55] S. Bareyt, S. Piligkos, B. Hasenknopf, P. Gouzerh, E. Lacote, S. Thorimbert, <strong>and</strong> M. Malacria. J.<br />

Am. Chem. Soc., 127:6788, 2005.<br />

[56] M. Rusu, A. R. Tomsa, D. Rusu, <strong>and</strong> I. Haiduc. Syn. React. Inorg. Met., 29:951, 1999.<br />

[57] J. T. Rhule, C. L. Hill, <strong>and</strong> D. A. Judd. Chem. Rev., 98(1):327–357, 1998.<br />

[58] P. Gouzerh <strong>and</strong> A. Proust. Chem. Rev., 98(1):77–111, 1998.<br />

[59] B. Hasenknopf, R. Delmont, P. Herson, <strong>and</strong> P. Gouzerh. Eur. J. Inorg. Chem., (5):1081–1087,<br />

2002.<br />

[60] U. Kortz, G. Savelieff, F. Y. A. Ghali, M. Khalil, S. A. Maalouf, <strong>and</strong> D. I. Sinno. Angew. Chem.<br />

Int. Ed., 41:4070, 2002.<br />

[61] B. B. Xu, Y. G. Wei, C. L. Barnes, <strong>and</strong> Z. H. Peng. Angew. Chem. Int. Ed., 40(12):2290, 2001.<br />

[62] X. Wei, M. H. Dickman, <strong>and</strong> M. T. Pope. J. Am. Chem. Soc., 120:10254, 1998.<br />

[63] P. R. Marcoux, B. Hasenknopf, J. Vaissermann, <strong>and</strong> P. Gouzerh. Eur. J. Inorg. Chem., (13):2406–<br />

2412, 2003.<br />

[64] R. K. C. Ho <strong>and</strong> Klemperer. J. Am. Chem. Soc., 100:6772, 1978.<br />

[65] W. H. Knoth. J. Am. Chem. Soc., 101:759, 1979.<br />

[66] J. F. W. Keana <strong>and</strong> M. D. Ogan. J. Am. Chem. Soc., 108:7951, 1986.<br />

[67] J. F. W. Keana, M. D. Ogan, M. Beer, Y. Lu, <strong>and</strong> J. Varkey. J. Am. Chem. Soc., 108:7957, 1986.<br />

[68] S. X. Cai, Y. Wu, <strong>and</strong> J. F. W. Keana. New J. Chem., 17:325, 1993.<br />

[69] M. Bösing, I. Loose, H. Pohlmann, <strong>and</strong> B. Krebs. Chem. Eur. J., 3(8):1232–1237, 1997.<br />

[70] C. Tourné, A. Revel, G. Tourné, <strong>and</strong> M. Vendrell. C. R. Acad. Sc. Paris,Ser. C, (277):643–645,<br />

1973.<br />

[71] U. Kortz, M. G. Savelieff, B. S. Bassil, <strong>and</strong> M. H. Dickman. Angew. Chem. Int. Ed. Engl., (40):3384–<br />

, 2001.<br />

[72] L.-H. Bi, R.-D. Huang, J. Peng, E.-B. Wang, Y.-H. Wang, <strong>and</strong> C.-W. Hu. J. Chem. Soc., Dalton<br />

Trans., pages 121–129, 2001.<br />

[73] P. J. Domaille. Inorg. Synth., volume 27. 1990.<br />

[74] R Acerete <strong>and</strong> J. Server-Carrió. J. Am. Chem. Soc., 112:9386, 1990.<br />

[75] R. Contant. Inorg. Synth., 27:108, 1990.<br />

[76] R. Contant <strong>and</strong> A. Tézé. Inorg. Chem., 24:4610, 1985.<br />

[77] Hervé. G <strong>and</strong> Tézé. A. Inorg. Synth., volume 27. 1990.<br />

[78] P. J. Domaille. Inorg. Synth., volume 27. 1990.<br />

[79] R. Contant. Can. J. Chem., 65:568–573, 1987.<br />

[80] C. M. Tourné <strong>and</strong> G. F. Tourné. J. Chem. Soc., Dalton. Trans., page 2411, 1988.<br />

[81] G. M. Sheldrick. <strong>University</strong> <strong>of</strong> Göttingen, Germany, 1996.<br />

124


[82] I. D. Brown <strong>and</strong> D. Altermatt. Acta Crys., B41:244–247, 1985.<br />

[83] U. Kortz, M. G. Savelieff, B. S. Bassil, B. Keita, <strong>and</strong> L. Nadjo. Inorg. Chem., 41:783–789, 2002.<br />

[84] B. Keita, F. Girard, L. Nadjo, R. Contant, R. Belghiche, <strong>and</strong> M. Abbessi. J. Electroanal. Chem.,<br />

508:70, 2001.<br />

[85] I. D. Brown <strong>and</strong> D Altermatt. Acta Cryst., B41:244, 1985.<br />

[86] M. H. Alizadeh, S. P. Harmalker, Y. Jeannin, J. Martinfrere, <strong>and</strong> M. T. Pope. J. Am. Chem. Soc.,<br />

107(9):2662–2669, 1985.<br />

[87] B. Keita <strong>and</strong> L. Nadjo. Mater. Chem. Phys., 22:77, 1989.<br />

[88] B. Keita, Y-W. Lu, L. Nadjo, <strong>and</strong> R. Contant. Eur. J. Inorg. Chem., (12):2463–2471, 2000.<br />

[89] B. Keita, Y. W. Lu, L. Nadjo, <strong>and</strong> R. Contant. Electrochem. Commun., 2:720, 2000.<br />

[90] G. S. Chorghade <strong>and</strong> M. T. Pope. J. Am. Chem. Soc., 109:5134–5138, 1987.<br />

[91] B. Keita, F. Girard, L. Nadjo, R. Contant, J. Canny, <strong>and</strong> M. Richet. J. Electroanal. Chem.,<br />

478(1-2):76–82, 1999.<br />

[92] K. C. Kim <strong>and</strong> M. T. Pope. J. Chem. Soc. Dalton Trans., (7):986–990, 2001.<br />

[93] R. Contant <strong>and</strong> G. Hervé. Inorg.Chem., 22:63–111, 2002.<br />

[94] D. Jabbour, B. Keita, I. M. Mbomekalle, L. Nadjo, <strong>and</strong> U. Kortz. Eur. J. Inorg. Chem., pages<br />

2036–2044, 2004.<br />

[95] E. Abdeljalil, B. Keita, L. Nadjo, <strong>and</strong> R. Contant. J. Solid State Electrochem., 5(2):94–106, 2001.<br />

[96] K. Wassermann, M. H. Dickman, <strong>and</strong> M. T. Pope. Angew. Chem. Int. Ed., 36(13-14):1445–1448,<br />

1997.<br />

[97] A. Müller, P. Kogerler, <strong>and</strong> A. W. M. Dress. Coord. Chem. Rev., 222:193–218, 2001.<br />

[98] G. Binnig, H. Rohrer, Ch. Gerber, <strong>and</strong> E. Weibel. Phys. Rev. Lett., 49:57, 1982.<br />

[99] H. J. Chiang, S in: Güntherodt <strong>and</strong> R. Wiesendanger. Scanning Tunneling Microscopy I. Springer-<br />

Verlag, Berlin.<br />

[100] T. A. Jung, F. J. Himpsel, R. R. Schlittler, <strong>and</strong> J. K. in: R. Wiesendanger Gimzewski. Scanning<br />

Probe Microscopy. Springer-Verlag, Berlin.<br />

[101] I. K. Song, R. B. Shnitser, J. J. Cowan, C. L. Hill, <strong>and</strong> M. A. Barteau. Inorg, Chem., 41:1292,<br />

2002.<br />

[102] I. K. Song <strong>and</strong> M. A. Barteau. Korean J. Chem. Eng., 19:567, 2002.<br />

[103] I. K. Song, J. E. Lyons, <strong>and</strong> M. A. Barteau. Catalysis Today, 81:137, 2003.<br />

[104] M. S. Kaba, I. K. Song, D. C. Duncan, C. L. Hill, <strong>and</strong> M. A. Barteau. Inorg. Chem., 37:398, 1998.<br />

[105] I. K. Song, M. S. Kaba, <strong>and</strong> M. A. Barteau. Langmuir, 18:2358, 2002.<br />

[106] M. S. Kaba, I. K. Song, <strong>and</strong> M. A. Barteau. J. Phys. Chem. B, 106:2337, 2002.<br />

[107] A. J. Fisher <strong>and</strong> P. E. Blochl. Phys. Rev. Lett., 70:3263, 1993.<br />

[108] I. B. Ó. Paz, J. M. Brihuega, J. M. Gómez-Rodríguez, <strong>and</strong> J. M. Soler. Phys. Rev. Lett., 94:056103,<br />

2005.<br />

[109] S. D. Feyter <strong>and</strong> F. C. D. Schryver. J. Phys. Chem. B, 109:4290, 2005.<br />

[110] S. Hembacher, F. J. Giessibl, <strong>and</strong> J. Mannhart. Phys. Rev. Lett, 94:056101, 2005.<br />

[111] S. Novokmet, M. S. Alam, V. Dremov, F. W. Heinemann, P. Müller, <strong>and</strong> R. Alsfasser. Angew.<br />

Chem. Int. Ed., 44:803, 2005.<br />

[112] A. M. Ako, H. Maid, S. Sperner, S. H. H. Zaidi, R. W. Saalfrank, M. S. Alam, P. Müller, <strong>and</strong> F. W.<br />

Heinnemann.<br />

[113] R. J. Errington, S. S. Petkar, B. R. Horrocks, A. Houlton, L. H. Lie, <strong>and</strong> S. N. Patole. Angew.<br />

Chem. Int. Ed., 117:1280, 2005.<br />

[114] B. Keita <strong>and</strong> L. Nadjo. Surface <strong>Science</strong>, 254:L443, 1991.<br />

[115] S. De Feyter <strong>and</strong> F. C De Schryver. Chem. Soc. Rev., 32:139, 2003.<br />

[116] P. Samorí. Chem. Soc. Rev., 34:551, 2005.<br />

[117] W. Ho. J. Chem. Phys., 117:11033, 2002.<br />

[118] P. Wahl, L. Diekhöner, M. A. Schneider, L. Vitali, G. Wittich, <strong>and</strong> K. Kern. Phys. Rev. Lett,<br />

93:176603, 2004.<br />

[119] A. Miura, Z. Chen, S. H-Uji-i, De Feyter, M Sdanowska, A. P. H. J. Jonkhejm, P. Schenning,<br />

B. Mejer, F. Würthner, <strong>and</strong> F. C. De Schryver. J. Am. Soc. Chem., 125:14968, 2003.<br />

[120] A. Gesquiere, S. De Feyter, <strong>and</strong> F. C. De Schryver. Nano Lett, 1:201, 2001.<br />

[121] F. Jäckel, M. D. Watson, K. Müllen, <strong>and</strong> J. P. Rabe. Phys. Rev. Lett., 92:188303, 2004.<br />

[122] T. Matsui, H. Kambara, Y. Niimi, K. Tagami, M. Tsukada, <strong>and</strong> H. Fukuyama. Phys. Rev. Lett.,<br />

125


94:226403, 2005.<br />

[123] R. J. Hamers, R. M. Tromp, <strong>and</strong> J. E. Demuth. Phys. Rev. Lett., 56:19743, 1986.<br />

[124] M. S. Alam, S. Strömsdörfer, V. Dremov, P Müller, J. Kortus, M. Ruben, <strong>and</strong> J. M. Lehn. Angew.<br />

Chem. Int. Ed., 44:7896, 2005.<br />

[125] J. A. Stroscio <strong>and</strong> W. J. Kaiser. Scanning Tunneling Microscopy. Academic Press, New York, 1993.<br />

[126] M. Rivera, R. L. Williamson, <strong>and</strong> M. J. Miles. J. Vac. Sci. Technol. B, 14:1472, 1996.<br />

[127] Z. Klusek <strong>and</strong> W. Kozlowski. J. Elect. Spect. <strong>and</strong> Relt. Phenom., 107:63, 2000.<br />

[128] C. Tourné, A. Revel, G. Tourné, <strong>and</strong> M. Vendrell. C. R. Acad. Sc. Paris, Ser. C, 277:643, 1973.<br />

[129] C. M. Tourné <strong>and</strong> G. F. Tourné. C. R. Acad. Sc.Paris, Ser. C., 281:933, 1975.<br />

[130] L. G. Detusheva, L. I. Kuznetsova, V. A. Likholobov, A. A. Vlasov, N. N. Boldyreva, S. G. Poryvaev,<br />

<strong>and</strong> V. V. Malakhov. Russ. J. Coord. Chem., 25:569, 1999.<br />

[131] P. Mialane, J. Marrot, A. Mallard, <strong>and</strong> G. Herveé. Inorg. Chim. Acta, 328:81–86, 2002.<br />

[132] F. Zonnevijlle <strong>and</strong> M. T. Pope. J. Am. Chem. Soc., 101:2731–2732, 1979.<br />

[133] P. Mialane, J. Marrot, E. Rivière, J. Nebout, <strong>and</strong> G. Hervé. Inorg. Chem., 40:44, 2001.<br />

[134] M. T. Pope <strong>and</strong> A. Müller. Angew. Chem-Int. Ed. Engl., 30(1):34–48, 1991.<br />

[135] J. J. Borrás-Almenar, E. Coronado, A. Müller, <strong>and</strong> M. T. Pope. Polyoxometalate Molecular <strong>Science</strong>.<br />

Kluwer: Dordrecht, Netherl<strong>and</strong>s, 2004.<br />

[136] L. Barcza <strong>and</strong> M. T. Pope. J. Phys. Chem., 77:1795, 1973.<br />

[137] K. Nomiya, M. Pohl, N. Mizuno, D. K. Lyon, <strong>and</strong> R. G. Finke. Inorg. Synth., 31:186, 1997.<br />

[138] K. Nomiya, C. Nozaki, K. Miyazawa, Y. Shimizu, T. Takayama, <strong>and</strong> K. Nomura. Bull. Chem. Soc.<br />

Jpn., 70:1369, 1997.<br />

[139] H. Weiner, J. D. Aiken III, <strong>and</strong> R. G. Finke. Inorg. Chem., 114:237, 1996.<br />

[140] T. Yamase, E. Ishikawa, Y. Asai, <strong>and</strong> S. Kanai. J. Mol. Catal. A, 114:237, 1996.<br />

[141] K. Nomiya, K. Ohsawa, T. Taguchi, M. Kaneko, <strong>and</strong> T. Takayama. Bull. Chem. Soc. Jpn., 71:2603,<br />

1998.<br />

[142] K. Nomiya, C. Nozaki, A. Kano, T. Taguchi, <strong>and</strong> K. Ohsawa. J. Organomet. Chem., 533:153, 1997.<br />

[143] Y. Lin <strong>and</strong> R. G. Finke. Inorg. Chem., 33:4891, 1994.<br />

[144] Y. Lin <strong>and</strong> R. G. Finke. J. Am. Chem. Soc., 116:8335, 1994.<br />

[145] J. D. Aiken III, Y. Lin, <strong>and</strong> R. G. Finke. J. Mol. Catal. A, 114:29, 1996.<br />

[146] J. H. He, X. H. Wang, Y. G. Chen, J. F. Liu, N. H. Hu, <strong>and</strong> H. Q. Jia. Inorg. Chem. Commun.,<br />

5(10):796–799, 2002.<br />

[147] F. Hussain, B. S. Bassil, L. H. Bi, M. Reicke, <strong>and</strong> U. Kortz. Angew. Chem. Int. Ed., 43(26):3485–<br />

3488, 2004.<br />

[148] K. Nomiya, M. Takahashi, K. Ohsawa, <strong>and</strong> J. A. Widegren. J. Chem. Soc. Dalton Trans., page<br />

2872 2878, 2001.<br />

[149] Y. Lin, T. J. R. Weakley, B. Rapko, <strong>and</strong> R. G. Finke. Inorg. Chem., 32:50955101, 1993.<br />

[150] T. Yamase, T. Ozeki, H. Sakamoto, S. Nishiya, <strong>and</strong> A. Yamamoto. Bull. Chem. Soc. Jpn., 66:103–<br />

108, 1993.<br />

[151] O. A. Kholdeeva, G. M. Maksimov, R. I. Maksimovskaya, L. A. Kovaleva, M. A. Fedotov, V. A.<br />

Grigoriev, <strong>and</strong> C. L. Hill. Inorg. Chem., 39:3828 3837, 2000.<br />

[152] W. H. Knoth, P. J. Domaille, <strong>and</strong> D. C. Roe. Inorg. Chem., 22:198–201, 1983.<br />

[153] T. Yamase, T. Ozeki, <strong>and</strong> S. Motomura. Bull. Chem. Soc. Jpn., 65:1453 1459, 1992.<br />

[154] P. J. Domaille <strong>and</strong> W. H. Knoth. Inorg. Chem., 22:818822, 1983.<br />

[155] T. Ozeki <strong>and</strong> T. Yamase. Acta Cryst. Sect. C, 47:693696, 1991.<br />

[156] F. X. Gao, T. Yamase, <strong>and</strong> H. Suzuki. J. Mol. Catal. A, 180:97108, 2002.<br />

[157] E. Ishikawa <strong>and</strong> T. Yamase. J. Mol. Catal. A, 142:6176, 1999.<br />

[158] T. Yamase, E. Ishikawa, Y. Asai, <strong>and</strong> S. Kanai. J. Mol. Catal. A, 114:237245, 1996.<br />

[159] T. Yamase <strong>and</strong> M. Sugeta. Inorg. Chim. Acta., (172):131, 1990.<br />

[160] Y. Take, Y. Tokutake, Y. Inouye, T. Yoshida, A. T. Yamamoto, Y. Yamase, <strong>and</strong> S. Nakamura.<br />

Antiviral Res., (16):327, 1991.<br />

[161] H. Hattori. Chem. Rev, (95):537, 1995.<br />

[162] S. Shigeta, S. Mori, E. Kodama, J. Kodama, K. Takahashi, <strong>and</strong> T. Yamase. Antiviral Res.,<br />

58(3):265–271, 2003.<br />

[163] U. Kortz, S. S. Hamzeh, <strong>and</strong> N. A. Nasser. Chem. Eur. J., 9(13):2945–2952, 2003.<br />

[164] Y. Sakai, K. Yoza, C. N. Kato, <strong>and</strong> K. Nomiya. Chem. Eur. J., 9:4077–4083, 2003.<br />

126


[165] Y. Sakai, K. Yoza, C. N. Kato, <strong>and</strong> K. Nomiya. J. Chem. Soc. Dalton Trans., page 35813586, 2003.<br />

[166] F. Hussain, U. Kortz, <strong>and</strong> R. J. Clark. Inorg. Chem., 43:3237–3241, 2004.<br />

[167] L. H. Bi, U. Kortz, B. Keita, L. Nadjo, <strong>and</strong> H. Borrmann. Inorg. Chem., 43:8367–8372, 2004.<br />

[168] I. Texier, C. Giannotti, S. Malato, C. Richter, <strong>and</strong> J. Delaire. Catal. Today, 54(2-3):297–307, 1999.<br />

[169] R. R. Ozer <strong>and</strong> J. L. Ferry. Environ. Sci. Technol., 35:32423246, 2001.<br />

[170] D. A. Friesen, L. Morello, J. V. Headley, <strong>and</strong> C. H. Langford. J. Photochem. Photobiol. A, 133:213–<br />

220, 2000.<br />

[171] U. Kortz <strong>and</strong> S. Matta. Inorg. Chem., 40:815–817, 2001.<br />

[172] U. Kortz, S. Isber, M. H. Dickman, <strong>and</strong> D. Ravot. Inorg. Chem., 39:2915–2922, 2000.<br />

[173] U. Kortz, Y. P. Jeannin, A. Tézé, G. Hervé, <strong>and</strong> S. Isber. Inorg. Chem., 38:3670–3675, 1999.<br />

[174] Y. Jeannin <strong>and</strong> M. Fournier. Pure Appl. Chem., 59:15291548, 1987.<br />

[175] M.T. Pope. Comp. Coord. Chem. II, 4:635, 2003.<br />

[176] T.J.R. Weakley, H.T.jun. Evans, J.S. Showell, <strong>and</strong> C.M. Tourné G. F. Tourné. J. Chem. Soc.,<br />

Chem. Commun., page 139, 1973.<br />

[177] R.G. Finke, M. Droege, J.R. Hutchinson, <strong>and</strong> O. Gansow. J. Am. Chem. Soc., 103:1587, 1981.<br />

[178] R.G. Finke <strong>and</strong> M.W. Droege. Inorg. Chem., 22:1006, 1983.<br />

[179] H.T. Evans, C.M. Tourné, <strong>and</strong> T. J.R. Weakley. G.F. Tourné. J. Chem. Soc., Dalton Trans., page<br />

2699, 1986.<br />

[180] R.G. Finke, M.W. Droege, <strong>and</strong> P.J. Domaille. Inorg. Chem., 26:3886, 1987.<br />

[181] S.H. Wasfi, A.L. Rheingold, G.F. Kokoszka, <strong>and</strong> A.S. Goldstein. Inorg. Chem., 26:2934, 1987.<br />

[182] T.J.R. Weakley <strong>and</strong> R.G. Finke. Inorg. Chem., 29:1235, 1990.<br />

[183] C.J. Gómez-García, E. Coronado, <strong>and</strong> J.J Borrás-Almenar. Inorg. Chem., 31:1667, 1992.<br />

[184] N. Casa n-Pastor, J. Bas-Serra, E. Coronado, G. Pourroy, <strong>and</strong> L.C.W. Baker. J. Am. Chem. Soc.,<br />

114:10308, 1992.<br />

[185] C.J. Gómez-García, E. Coronado, P. Gómez-Romero, <strong>and</strong> N. Casa n-Pastor. Inorg. Chem., 32:3378,<br />

1993.<br />

[186] C.J. Gómez-García, J.J. Borrás-Almenar, E. Coronado, <strong>and</strong> L. Ouahab. Inorg. Chem., 33:4016,<br />

1994.<br />

[187] X.-Y. Zhang, G.B. Jameson, C.J. O’Connor, <strong>and</strong> M.T. Pope. Polyhedron., 15:917, 1996.<br />

[188] X. Zhang, Q. Chen, D.C. Duncan, C. Campana, <strong>and</strong> C. L Hill. Inorg. Chem., 36:4208, 1997.<br />

[189] X. Zhang, Q. Chen, D. C. Duncan, R. J. Lachicotte, <strong>and</strong> C. L. Hill. Inorg. Chem., 36(20):4381–4386,<br />

1997.<br />

[190] J. M. Clemente-Juan, E. Coronado, J. R. Galán-Mascarós, <strong>and</strong> C. J. Gómez-García. Inorg. Chem.,<br />

38(1):55–63, 1999.<br />

[191] L.-H. Bi, E.-B. Wang, J. Peng, R.-D. Huang, L. Xu, <strong>and</strong> C.-W. Hu. Inorg. Chem., 39:671, 2000.<br />

[192] L.-H. Bi, R.-D. Huang, J. Peng, E.-B. Wang, Y.-H. Wang, <strong>and</strong> C.-W. Hu. J. Chem. Soc., Dalton<br />

Trans., 39:121, 2001.<br />

[193] E.M. Limanski, M. Piepenbrink, E. Droste, K. Burgemeister, <strong>and</strong> B. Krebs. J. Clust. Sci., 13:369,<br />

2002.<br />

[194] C. Rosu, D.C. Crans, <strong>and</strong> T. J. R. Weakley. Polyhedron., 21:959, 2002.<br />

[195] U. Kortz, S. Nellutla, A. C. Stowe, N. S. Dalal, U. Rauwald, W. Danquah, <strong>and</strong> D. Ravot. Inorg.<br />

Chem., 43:2308–2317, 2004.<br />

[196] F. Robert, M. Leyrie, <strong>and</strong> G. Hervé. Acta. Cryst., B38:358, 1982.<br />

[197] M. Bösing, A. Noh, I. Loose, <strong>and</strong> B. Krebs. J. Am. Chem. Soc., 120(29):7252–7259, 1998.<br />

[198] U. Kortz, N. K. Al-Kassem, M. G. Savelieff, N. A. Al Kadi, <strong>and</strong> M. Sadakane. Inorg. Chem.,<br />

40:4742–4749, 2001.<br />

[199] B. Botar, T. Yamase, <strong>and</strong> E. Ishikawa. Inorg. Chem. Commun., 4:551, 2001.<br />

[200] T. Yamase, B. Botar, E. Ishikawa, <strong>and</strong> K. Fukaya. Chem. Lett., page 56, 2001.<br />

[201] P. Mialane, J. Marrot, E. Riviére, J. Nebout, <strong>and</strong> G. Hervé. Inorg. Chem., 40:44, 2001.<br />

[202] U. Kortz, S. Nellutla, A. C. Stowe, N. S. Dalal, J. van Tol, <strong>and</strong> B. S. Bassil. Inorg. Chem.,<br />

43:144–154, 2004.<br />

[203] D. Drewes, E. M. Limanski, M. Piepenbrink, <strong>and</strong> B. Krebs. Z. Anorg. Allg. Chem., 630(1):58–62,<br />

2004.<br />

[204] L.-H. Bi, M. Reicke, U. Kortz, B. Keita, L. Nadjo, <strong>and</strong> R. J. Clark. Inorg. Chem., 43:3915, 2004.<br />

[205] E. M. Limanski, D. Drewes, E. Droste, R. Bohner, <strong>and</strong> B. Krebs. J. Mol. Struct., 656:17–25, 2003.<br />

127


[206] I. Loose, E. Droste, M. Bösing, H. Pohlmann, M.H. Dickman, C. Roşu, M.T. Pope, <strong>and</strong> B. Krebs.<br />

Inorg. Chem., 38:2688, 1999.<br />

[207] B. Krebs, E. Droste, M. Piepenbrink, <strong>and</strong> G. Vollmer. C. R. Acad. Sci. Paris. Ser, IIc 3:205, 2000.<br />

[208] W.H. Knoth, P.J. Domaille, <strong>and</strong> R.D. Farlee. Organometalics, 4:62, 1985.<br />

[209] W. H. Knoth, P. J. Domaille, <strong>and</strong> R. L. Harlow. Inorg. Chem., 25:1577–1584, 1986.<br />

[210] F. Xin <strong>and</strong> M. T. Pope. J. Am. Chem. Soc., 118:7731, 1996.<br />

[211] R.G. Finke, B. Rapko, <strong>and</strong> T.J.R. Weakley. Inorg. Chem., 28:1573, 1989.<br />

[212] N. Laronze, J. Marrot, <strong>and</strong> G. Hervé. Inorg. Chem., 42:5857, 2003.<br />

[213] R. Contant. J. Chem. Res., S:120, 1984.<br />

[214] R. Contant. J. Chem. Res., M:1063, 1984.<br />

[215] G.M. Maksimov, G.N. Kustova, K.I. Matveev, <strong>and</strong> T.P. Lazarenko. Koord. Khim., 15:788, 1989.<br />

[216] J. F. Kirby <strong>and</strong> L. C. W. Baker. J. Am. Chem. Soc., 117:10010–10016, 1995.<br />

[217] M.H. Alizadeh, H. Razavi, F.M. Zonoz, <strong>and</strong> M.R. Mohammadi. Polyhedron., 22:933, 2003.<br />

[218] J. G. Liu, F. Ortega, P. Sethuraman, D. E. Katsoulis, C. E. Costello, <strong>and</strong> M. T. Pope. J. Chem.<br />

Soc., Dalton Trans., pages 1901–1906, 1992.<br />

[219] T. M. Anderson, X. Zhang, K. I. Hardcastle, <strong>and</strong> C. L. Hill. Inorg. Chem., 41:2477–2488, 2002.<br />

[220] U. Kortz, I. M. Mbomekalle, B. Keita, L. Nadjo, <strong>and</strong> P. Berthet. Inorg. Chem., 41:6412–6416,<br />

2002.<br />

[221] L. Meng, J. Liu, Y. Wu, D. Zhao, <strong>and</strong> Y. Xiao. Polyhedron., 14:2127, 1995.<br />

[222] S.M. Wasfi <strong>and</strong> S.A. Tribbitt. Inorg. Chim. Acta., 268:329, 1998.<br />

[223] E. M. Limanski, D. Drewes, <strong>and</strong> B. Krebs. Z. Anorg. Allg. Chem., 630(4):523–528, 2004.<br />

[224] C. M. Tourné, G. F. Tourné, <strong>and</strong> F. Zonnevijlle. J. Chem. Soc., Dalton. Trans., page 143, 1991.<br />

[225] W. Adam, P. L. Alsters, R. Neumann, C. R. Saha-Moller, D. Seebach, A. K. Beck, <strong>and</strong> R. Zhang.<br />

J. Org. Chem., 68(21):8222–8231, 2003.<br />

128


APPENDIX


NMR Spectra<br />

Fig. 3.54: 13 C NMR spectrum <strong>of</strong> polyanion 1<br />

130


Fig. 3.55: 13 C NMR spectrum <strong>of</strong> polyanion 2<br />

131


Fig. 3.56: 31 P NMR spectrum <strong>of</strong> polyanion 3<br />

132


Fig. 3.57: 119 Sn NMR spectrum <strong>of</strong> polyanion 3<br />

133


Fig. 3.58: 13 C NMR spectrum <strong>of</strong> polyanion 3<br />

134


Fig. 3.59: 1 H NMR spectrum <strong>of</strong> polyanion 3<br />

135


Fig. 3.60: 1 H NMR spectrum <strong>of</strong> polyanion 5<br />

136


Fig. 3.61: 13 C NMR spectrum <strong>of</strong> polyanion 5<br />

137


Fig. 3.62: 119 Sn NMR spectrum <strong>of</strong> polyanion 5<br />

Fig. 3.63: 1 H NMR spectrum <strong>of</strong> polyanion 6<br />

138


Fig. 3.64: 13 C NMR spectrum <strong>of</strong> polyanion 6<br />

Fig. 3.65: 119 Sn NMR spectrum <strong>of</strong> polyanion 6<br />

139


Fig. 3.66: 1 H NMR spectrum <strong>of</strong> polyanion 7<br />

Fig. 3.67: 13 C NMR spectrum <strong>of</strong> polyanion 7<br />

140


Fig. 3.68: 183 W NMR spectrum <strong>of</strong> polyanion 7<br />

Fig. 3.69: 1 H NMR spectrum <strong>of</strong> polyanion 8<br />

141


Fig. 3.70: 13 C NMR spectrum <strong>of</strong> polyanion 8<br />

Fig. 3.71: 119 Sn NMR spectrum <strong>of</strong> polyanion 8 in presence <strong>of</strong> sodium chloride<br />

142


Incomplete Results<br />

1-The hybrid organic-inorganic 1-D material: {K 7 [{Sn(CH 3 ) 2 } 3 (H 2 O) 2 (PW 9 O 36 )]} ∞<br />

Experimental<br />

Preparation <strong>of</strong> {[{Sn(CH 3 ) 2 } 3 (H 2 O) 2 (PW 9 O 36 )] 7− } ∞ (18): A 1.46 g (0.600 mmol) sample<br />

<strong>of</strong> Na 9 [A-PW 9 O 34 ] [73] was added with stirring to a solution <strong>of</strong> 0.435 g (1.98 mmol)<br />

(CH 3 ) 2 SnCl 2 in 20 mL H 2 O. The pH was adjusted to 6 by addition <strong>of</strong> 1 M NaOH solution.<br />

This solution was heated to ∼80 ◦ C for 1 hour <strong>and</strong> then cooled to room temperature <strong>and</strong><br />

filtered. Addition <strong>of</strong> 0.5 mL <strong>of</strong> 1.0 M KCl solution to the colorless filtrate <strong>and</strong> slow<br />

evaporation at room temperature led to a white crystalline product after about a week<br />

or two. FTIR spectroscopy: 1065(s), 1007(m), 935(s), 914(sh), 837(m), 789(s), 656(s),<br />

593(w), 575(w), 515(m) cm −1 .<br />

Fig. 3.72: FTIR spectra <strong>of</strong> compound K-18(red) <strong>and</strong> Na 9 [A-PW 9 O 34 ](blue)<br />

X-ray Crystallography<br />

A crystal <strong>of</strong> compound K-18 was mounted on a glass fiber for indexing <strong>and</strong> intensity data<br />

collection at 163 K on a Bruker D8 SMART APEX CCD single-crystal diffractometer<br />

using Mo K radiation ( λ = 0.71073 Å). Direct methods were used to solve the structure<br />

<strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining atoms were found from<br />

successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong> polarization corrections<br />

143


were applied <strong>and</strong> an absorption correction was performed using the SADABS program<br />

[81]. Crystallographic data are summarized in Table 3.15.<br />

Table 3.15: Crystal Data <strong>and</strong> Structure Refinement for compound K-18<br />

Empirical formula W 36 Sn 8 P 4 K 8 C 16 O 188<br />

fw 11205<br />

space group (No.) P 2 1 /c (14)<br />

a (Å) 12.3879(8)<br />

b (Å) 14.1573(9)<br />

c (Å) 32.6941(20)<br />

β ( ◦ ) 90.826(1)<br />

vol (Å 3 ) 5733.3(63)<br />

Z 4<br />

temp ( ◦ C) -110<br />

wavelength (Å) 0.71073<br />

d calcd (mg m −3 ) 3.25<br />

abs coeff. (mm −1 ) 20<br />

R [I > 4 σ(I)] a 0.113<br />

R w (all data) b 0.258<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

Solution NMR<br />

Polyanion 18 is diamagnetic <strong>and</strong> contains four spin 1 nuclei 2 (183 W, 119 Sn, 13 C, 1 H) <strong>and</strong><br />

therefore represents a good c<strong>and</strong>idate for solution NMR studies at room temperature.<br />

We examined the solution properties <strong>of</strong> 18 by 31 P NMR (D 2 O, ∼20 ◦ ). Our 31 P NMR<br />

measurements resulted in one singlet at (-12.3 ppm). Further solution NMR <strong>of</strong> differerent<br />

nuclei are to be done.<br />

HPPS measurement<br />

The dynamic light scattering measurements was done on a freshly prepared solution <strong>of</strong><br />

polyanion 18 in collaboration with Pr<strong>of</strong>. M. Winterhalter <strong>and</strong> his group (IUB, Germany).<br />

High performance particle sizer(HPPS) provided by Malvern instruments, suggest that<br />

the particle were monodispersed. The size distribution by intensity showed two kinds<br />

<strong>of</strong> particle, a size <strong>of</strong> diameter 3.1 nm <strong>and</strong> 284 nm but size distribution by number <strong>and</strong><br />

144


Fig. 3.73: 31 P NMR spectrum <strong>of</strong> polyanion 18<br />

volume suggest that the particle <strong>of</strong> size 3.1 nm is more populated.<br />

All the measurement were done in aqueous medium in a polystyrene cuvette.<br />

Fig. 3.74: Size distribution by intensity <strong>of</strong> polyanion 18<br />

Results <strong>and</strong> discussion<br />

The structure <strong>of</strong> the polyanion 18 can be best described as a 1-D polymeric composed <strong>of</strong><br />

monopolyanionic building blocks ({{Sn(CH 3 ) 2 } 3 }(H 2 O) 2 (α-PW 9 O 36 ) 7− ) that are linked<br />

via Sn-O-(W’) bridges. The three organo-tin groups attached to each monomeric unit<br />

<strong>of</strong> 18 contain tin centers that are octahedrally coordinated by four oxygen atoms <strong>and</strong><br />

two methyl groups <strong>and</strong> two water molecules. Furthermore the two methyl groups on<br />

145


Fig. 3.75: Size distribution by number <strong>of</strong> polyanion 18<br />

Fig. 3.76: Size distribution by volume <strong>of</strong> polyanion 18<br />

each tin atom are positioned trans to each other. All the tin centers are grafted on<br />

the A-type [PW 9 O 34 ] 9− via coordination to the terminal oxygen atoms <strong>of</strong> the two edgeshared<br />

octahedra. Out <strong>of</strong> the three dimethyltin units, two tin atoms are coordinated to<br />

the neighboring polyanion by 2Sn-O-(W’) bridges <strong>and</strong> one tin atom possess two water<br />

molecules as terminal lig<strong>and</strong>s. Bond-valence-sum (BVS) calculations for 18 indicated<br />

Fig. 3.77: Ball/stick <strong>and</strong> polyhedron representation <strong>of</strong> solid state <strong>of</strong> 18 showing 1-D chain<br />

that no oxygen <strong>of</strong> the building blocks (PW 9 O 34 ) caps is protonated [85].<br />

146


2-The hybrid organic-inorganic 1-D material:(K 10 [{Sn(CH 3 ) 2 } 2 (P 2 W 12 O 48 )]) ∞<br />

Experimental<br />

Preparation <strong>of</strong> ({[{Sn(CH 3 ) 2 } 2 (P 2 W 12 O 48 )]} 10− ) ∞ (19): A 2.36 g (0.600 mmol) sample<br />

<strong>of</strong> K 12 [H 2 P 2 W 12 O 48 ] [75] was added with stirring to a solution <strong>of</strong> 0.435 g (1.98 mmol)<br />

(CH 3 ) 2 SnCl 2 in 20 mL H 2 O. The pH was adjusted to 6 by addition <strong>of</strong> 1 M NaOH solution.<br />

This solution was heated to ∼50 ◦ C for 30 minutes <strong>and</strong> then cooled to room temperature<br />

<strong>and</strong> filtered. Addition <strong>of</strong> 0.5 mL <strong>of</strong> 1.0 M KCl solution to the colorless filtrate <strong>and</strong> slow<br />

evaporation at room temperature led to a white crystalline product after about a week or<br />

two. FTIR spectroscopy: 1129(m), 1084(s), 1052(w), 1016(m), 940(vs), 917(vs), 885(sh),<br />

804(vs), 732(vs), 566(sh), 522(w), 463(w) cm −1 .<br />

Fig. 3.78: FTIR spectra <strong>of</strong> compound K-19(red) <strong>and</strong> K 12 [H 2 P 2 W 12 O 48 ](blue)<br />

X-ray Crystallography<br />

A crystal <strong>of</strong> compound K-19 was mounted on a glass fiber for indexing <strong>and</strong> intensity data<br />

collection at 163 K on a Bruker D8 SMART APEX CCD single-crystal diffractometer<br />

using Mo K radiation ( λ = 0.71073 Å). Direct methods were used to solve the structure<br />

<strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining atoms were found from<br />

successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong> polarization corrections<br />

were applied <strong>and</strong> an absorption correction was performed using the SADABS program<br />

[81]. Crystallographic data are summarized in Table 3.16.<br />

147


Table 3.16: Crystal Data <strong>and</strong> Structure Refinement for compound K-19<br />

Empirical formula W 48 Sn 8 P 8 K 36 O 264 C 16 H 48<br />

fw 11205<br />

space group (No.) C 2/c (15)<br />

a (Å) 14.5700(14)<br />

b (Å) 24.3039(14)<br />

c (Å) 20.4474(14)<br />

β ( ◦ ) 100.682(2)<br />

vol (Å 3 ) 7115.12(3)<br />

Z 4<br />

temp ( ◦ C) -110<br />

wavelength (Å) 0.71073<br />

d calcd (mg m −3 ) 3.71<br />

abs coeff. (mm −1 ) 20.38<br />

R [I > 2 σ(I)] a 0.051<br />

R w (all data) b 0.123<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

Results <strong>and</strong> discussion<br />

The structure <strong>of</strong> polyanion 19 is best described as a polymer composed <strong>of</strong> monopolyanionic<br />

building blocks ({{Sn(CH 3 ) 2 } 2 }(H 2 O) 2 [H 2 P 2 W 12 O 48 ] 10− ) that are linked via Sn-O-(W’)<br />

bridges. This arrangement leads to a 1-D polymeric chain. The two organo-tin groups<br />

attached to each monomeric unit <strong>of</strong> 19 contain tin centers that are five coordinated<br />

by two methyl groups <strong>and</strong> three oxo groups, respectively. Furthermore the two methyl<br />

groups on each tin atom are positioned trans to each other <strong>and</strong> all the organotin groups<br />

are structurally equivalent. The tin atoms are connected to the adjacent polyanion by<br />

a single Sn-O-(W’) bridges in such a way that they are coordinated to two terminal<br />

oxygens <strong>of</strong> the cap <strong>of</strong> the (H 2 P 2 W 12 O 48 ) fragment <strong>and</strong> one terminal oxygen <strong>of</strong> the belt<br />

<strong>of</strong> the adjacent polyanion. Alternatively 19 can be best described as described as a<br />

hexalacunary [H 2 P 2 W 12 O 48 ] 12− fragment which has taken up two (CH 3 ) 2 Sn 2+ units. All<br />

the tin atoms are five coordinated (trigonal pyramidal) <strong>and</strong> all the methyl groups are<br />

trans to each other. All the tin atoms are coordinated to the neighboring polyanion by<br />

single Sn-O-(W’) bridge. Bond-valence-sum (BVS) calculations for 19 indicated that no<br />

oxygen <strong>of</strong> the building blocks (H 2 P 2 W 12 O 48 ) caps is protonated [85].<br />

148


Fig. 3.79: Ball/stick <strong>and</strong> polyhedron representation <strong>of</strong> solid state <strong>of</strong> 19 showing 1-D chain<br />

3-The cyclic trimeric-titanium substituted, tungstophosphate:<br />

[{Ti 3 O(A-α-PW 9 O 37 )(OH)} 3 ] 10−<br />

Experimental<br />

The precursor K 14 [P 2 W 19 O 69 (H 2 O)] was synthesized according to the published procedure<br />

<strong>of</strong> Tourné et al. <strong>and</strong> the purity was confirmed by infrared spectroscopy [80]. All<br />

other reagents were used as purchased without further purification. K 10 [{Ti 3 O(A-α-<br />

PW 9 O 37 )(OH)} 3 ] (K-20). The title compound was synthesized by dissolving 0.141 g<br />

(0.88 mmols) <strong>of</strong> TiO(SO 4 ) in 40 mL potassium acetate buffer followed by addition <strong>of</strong> 2.26<br />

g (0.40 mmols) K 14 [P 2 W 19 O 69 (H 2 O)]. This solution (pH 4.8) was heated to ∼80 ◦ C for<br />

1 h <strong>and</strong> then cooled to room temperature. 2 mL <strong>of</strong> 30% H 2 O 2 was added to the solution.<br />

The color <strong>of</strong> the solution changed to golden yellow. The solution was filtered <strong>and</strong> a 1 mL<br />

<strong>of</strong> 0.1 M NH 4 Cl solution were added <strong>and</strong> then the solution was allowed to evaporate in<br />

an open vial at room temperature. A white crystalline product started to appear after a<br />

week or two. Evaporation was continued until the solvent approached the solid product.<br />

FTIR spectra for K 10 [{Ti 3 O(A-α-PW 9 O 37 )(OH)} 3 ] : 1165(sh), 1064(s), 1031(w), 959(s),<br />

882(w), 783(s), 668(s), 652(s), 617(w), 587(sh) cm −1 . The FTIR spectrum was recorded<br />

on a Nicolet Avatar FTIR spectrophotometer in a KBr pellet.<br />

X-ray Crystallography<br />

A crystal <strong>of</strong> compound K-20 was mounted on a glass fiber for indexing <strong>and</strong> intensity data<br />

collection at 173 K on a Bruker D8 SMART APEX CCD single-crystal diffractometer<br />

using Mo K α radiation (λ = 0.71073 Å). Direct methods were used to solve the structure<br />

149


Fig. 3.80: FTIR spectra <strong>of</strong> compound K-20(red) <strong>and</strong> K 14 [P 2 W 19 O 69 (H 2 O)](blue)<br />

<strong>and</strong> to locate the heavy atoms (SHELXS97). Then the remaining atoms were found from<br />

successive difference maps (SHELXL97). Routine Lorentz <strong>and</strong> polarization corrections<br />

were applied <strong>and</strong> an absorption correction was performed using the SADABS program<br />

[81]. Crystallographic data are summarized in Table 3.17<br />

Table 3.17: Crystal Data <strong>and</strong> Structure Refinement for compound K-20<br />

Empirical formula K 45 O 369 P 9 Ti 27 W 81<br />

fw 15507.27<br />

space group R 3m (160)<br />

a (Å) 29.8461(7)<br />

c (Å) 13.6781(8)<br />

volume (Å 3 ) 10551.9(8)<br />

Z 3<br />

temp. ( ◦ C) -100<br />

wavelength (Å) 0.71073<br />

dcalc (mg m −3 ) 2.440<br />

abs. coeff. (mm −1 ) 15.72<br />

R [I > 2 σ(I)] a 0.071<br />

R w (all data) b 0.2079<br />

R = ∑ ||F o |-|F c ||/ ∑ |F o |. b R w = [ ∑ w(F 2 o-F 2 c) 2 / ∑ w(F 2 o) 2 ] 1/2<br />

Results <strong>and</strong> discussion<br />

The novel polyanion 20 is isostructural to the polyanion 10. The polyanion 20 is composed<br />

<strong>of</strong> three [A-α-PW 9 O 34 ] units <strong>and</strong> nine TiO 6 octahedra. The nine TiO 6 octahedra<br />

150


Fig. 3.81: Ball/stick <strong>and</strong> polyhedral representation <strong>of</strong> polyanion 20<br />

are connected to each other by Ti-O-Ti bridges. Three TiO 6 octahedra <strong>of</strong> each trilacunary<br />

[A-α-PW 9 O 34 ] unit fills the lacuna to form a complete Keggin cluster. Of the<br />

three TiO 6 octahedra two are connected to Ti-centers <strong>of</strong> adjacent polyanions by Ti-O-Ti<br />

bridges, the unique TiO 6 octahedra in each Keggin subunit has terminal hydroxolig<strong>and</strong>.<br />

Bond-valence-sum (BVS) calculations for polyanion 20 indicated that the unique Ti is<br />

monoprotonated no oxygen <strong>of</strong> the building blocks [A-α-PW 9 O 34 ] caps is protonated [85].<br />

The polyanion has a nominal symmetry <strong>of</strong> C 3v (see Figure. 3.74<br />

Catalytic studies are on progress by our collaborator Pr<strong>of</strong>. O. A.<br />

Kholdeeva, Boreskov Institute Catalysis, Novosibirsk, Russia.<br />

151


FIRASAT HUSSAIN<br />

International <strong>University</strong> Bremen<br />

Campus Ring 8, Res. III, Room 110<br />

Bremen, D-28759, Germany.<br />

Curriculum Vitae<br />

Education<br />

Ph.D. (Chemistry), May 2006, International <strong>University</strong> Bremen, Germany.<br />

Advisor: Pr<strong>of</strong>. Ulrich Kortz, International <strong>University</strong> Bremen.<br />

M.Phil. (Chemistry), 2001, Utkal <strong>University</strong>, India.<br />

M.S. (Chemistry), 2000, Utkal <strong>University</strong>, India.<br />

B.S. (Physics, Chemistry <strong>and</strong> Mathematics), 1998, Sambalpur <strong>University</strong>, Orissa, India.<br />

Research Experience<br />

Research Scholar, International <strong>University</strong> Bremen, Germany, 01/2003- 31/2005<br />

• Synthesized novel hybrid inorganic-organic polyoxoanions based on dimethyltin.<br />

• Synthesized novel supramolecular inorganic-organic polyoxoanions based on dimetyltin.<br />

• Synthesized titanium, indium, cadmium <strong>and</strong> copper containing novel poloxoanions.<br />

Research Assistant, Indian Institute <strong>of</strong> Technology, Bombay, India. 12/2001-<br />

03/2003<br />

• Synthesized <strong>and</strong> studied the catalytic properties <strong>of</strong> titanium <strong>and</strong> chromium containing<br />

hexagonal mesoporous aluminophosphates.<br />

• Synthesized Al-, Ga-containing novel mesoporous MCM-48 molecular sieves. These<br />

catalysts tested for the vapour phase phenol alkylation.<br />

Personal Information<br />

Indian,unmarried <strong>and</strong> born on 19 th February, 1975, in Bhawanipatna (Orissa), India.<br />

152


Teaching Experience<br />

At International <strong>University</strong> Bremen, Bremen, Germany<br />

Teaching Assistant - (4 semesters) Tutorials For B.S. student in General Chemistry <strong>and</strong><br />

laboratory. For a semester Physical Chemistry laboratory<br />

Instrumentations known<br />

Experienced in h<strong>and</strong>ling different essential analytical tools(XRD, UV-vis spectroscopy,<br />

infrared spectroscopy, Multi nuclear NMR required for characterization <strong>of</strong> polyoxoanions.<br />

Conservant with other analytical techniques ( Vapour phase reactor, Thermal techniques<br />

(TG-DTA) Transmission electron microscopy.<br />

Computer skills<br />

S<strong>of</strong>twares: Diamond crystallographic s<strong>of</strong>tware, Origin, Delta s<strong>of</strong>tware for NMR, Latex<br />

<strong>and</strong> MSOffice, ChemDraw, photoshop.<br />

Academic Honours<br />

• (i)Qualified - Graduate Aptitude Test In <strong>Engineering</strong> (GATE) 90.43, Conducted by<br />

Indian Institute <strong>of</strong> Technology, India, 2000.<br />

• (ii) Gold Medalist for outst<strong>and</strong>ing performance in M.Phil in Chemistry.<br />

• (iii) 5th position in the <strong>University</strong> for B.S.<br />

Hobbies<br />

Cricket, Table tennis, Soccer, Reading, Watching television <strong>and</strong> listening to music.<br />

153


Publications<br />

1. Observation <strong>of</strong> a Half Step Magnetization in the (Cu 3 )-Type Triangular Spin Ring<br />

K.-Y. Choi, Y. H. Matsuda, H. Nojiri*, U. Kortz*, F. Hussain, A. C. Stowe, C.<br />

Ramsey, N. S. Dalal*, Phys. Rev. Lett., 2006, 96, 107202.<br />

2. STM/STS Observation <strong>of</strong> Polyoxoanions on HOPG Surfaces: The Wheel-shaped<br />

[Cu 20 Cl(OH) 24 (H 2 O) 12 (P 8 W 48 O 184 )] 25− <strong>and</strong> the Ball-shaped<br />

[{Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-PW 9 O 34 ) 12 ] 36−<br />

M. S. Alam, V. Dremov, P. Müller*, A. V. Postnikov, S. S. Mal, F. Hussain, U.<br />

Kortz*, Inorg. Chem., 2006, 45, 2866-2872.<br />

3. Tetrakis-Dimethyltin Containing Tungstophosphate [{Sn(CH 3 ) 2 } 4 (H 2 P 4 W 24 O 92 ) 2 ] 28− :<br />

First Evidence for Lacunary Preyssler Ion<br />

F. Hussain, U. Kortz*, B. Keita, L. Nadjo*, M. T. Pope, Inorg. Chem., 2006, 45,<br />

761-766.<br />

4. Ball-Shaped Heteropolytungstates [{Sn(CH 3 ) 2 (H 2 O)} 24 {Sn(CH 3 ) 2 } 12 (A-XW 9 O 34 ) 12 ] 36−<br />

(X = P V , As V )<br />

U. Kortz*, F. Hussain, M. Reicke, Angew. Chem. Int. Ed. 2005, 44, 3773-3777.<br />

5. Structure <strong>and</strong> Solution Properties <strong>of</strong> the Cadmium(II)-Substituted Tungstoarsenate<br />

[Cd 4 Cl 2 (B-α-AsW 9 O 34 ) 2 ] 12−<br />

F. Hussain, L. Bi, U.Rauwald, M. Reicke <strong>and</strong> U. Kortz*, Polyhedron, 2005, 24,<br />

847-852.<br />

6. Polyoxoanions Functionalized By Diorganotin Groups: The Tetrameric,<br />

Chiral Tungstoarsenate(III) [{Sn(CH 3 ) 2 (H 2 O)} 2 {Sn(CH 3 ) 2 }As 3 (α-AsW 9 O 33 ) 4 ] 21−<br />

F. Hussain <strong>and</strong> U. Kortz*. Chem. Commun., 2005, 1191-1193.<br />

7. Some Indium(III)-Substituted Polyoxotungstates <strong>of</strong> the Keggin <strong>and</strong> Dawson Type<br />

F. Hussain, M. Reicke, V. Janowski, S. de Silva, J. Futuwi, U. Kortz* Comptes<br />

Rendus Chimie, 2005, 8, 1045-1056.<br />

154


8. A Novel Isopolytungstate Functionalized by Ruthenium:[HW 9 O 33 Ru II<br />

2 (dmso) 6 ] 7−<br />

L. Bi, F. Hussain, U. Kortz*, M. Sadakane, M. H. Dickman, Chem. Commun.,<br />

2004, 1420.<br />

9. Structural Control on the Nanomolecular Scale: Self-Assembly <strong>of</strong> The Polyoxotungstate<br />

Wheel [(β-Ti 2 SiW 10 O 39 ) 4 ] 24−<br />

F. Hussain,B. S. Bassil, L. Bi, M. Reicke, U. Kortz*. Angew. Chem. Int. Ed., 2004,<br />

43, 3485.<br />

10. The Bis-Phenyltin Substitued, Lone Pair Containing Tungstoarsenate<br />

[Na(H 2 O)(C 6 H 5 Sn) 2 As 2 W 19 O 67 (H 2 O)] 7−<br />

F. Hussain,U. Kortz*, R. J. Clark. Inorg. Chem., 2004, 43, 3237.<br />

11. Polyoxoanions Functionalized by Diorganotin Groups. 1. The Hybrid Organic-<br />

Inorganic 2-D material (CsNa 4 [{Sn(CH 3 ) 2 } 3 O(H 2 O) 4 (β-XW 9 O 33 )] ·5H 2 O) ∞ (X =As III ,<br />

Sb III ) <strong>and</strong> its Solution Properties<br />

F. Hussain, M. Reicke, U. Kortz*. Eur. J. Inorg. Chem., 2004, 2733.<br />

12. Synthesis, characterization, <strong>and</strong> catalytic properties <strong>of</strong> chromium- containing hexagonal<br />

mesoporous aluminophosphate molecular sieves.<br />

S. K. Mohapatra, F. Hussain, P.Selvam* Catalysis Letters Vol. 85, Nos. 3-4, Feb.<br />

2003, 217-222.<br />

13. Titanium substituted hexagonal mesoporous aluminophosphates: Highly efficient<br />

<strong>and</strong> selective heterogeneous catalysts for the oxidation <strong>of</strong> phenols at room temperature.<br />

S.K. Mohapatra, F. Hussain, P. Selvam* Catalysis Communications 4, 2003, 57-62.<br />

155


PAPERS IN CONFERENCE/ SYMPOSIA / WORK-<br />

SHOPS<br />

1. F. Hussain, M. Reicke <strong>and</strong> U. Kortz*, Hybrid Organic-Inorganic Polyoxometalates<br />

functionalized by Organotin Moieties, Extended Abstract-7th Norddeutsches<br />

Doktor<strong>and</strong>en- Kolloqium der anorganisch-chemischen Institute, Organisation vom<br />

Institute fur Anorganische Chemie der Christian-Alberchts-Universitat zu Kiel, Hamburg,<br />

Germany, Sept 30-Oct 01, 2004, p12.<br />

2. F. Hussain, L. Bi, B. Bassil <strong>and</strong> U. Kortz*, Discrete Polyoxoanions: Nanomolecular<br />

structures with Multiple Functions, Extended Abstract-7th Norddeutsches<br />

Doktor<strong>and</strong>en- Kolloqium der anorganisch-chemischen Institute, Organisation vom<br />

Institute fur Anorganische Chemie der Christian-Alberchts-Universitat zu Kiel, Hamburg,<br />

Germany, Sept 30-Oct 01, 2004, p12.<br />

3. B. S. Bassil, F. Hussain, U. Kortz*, S. Nellutla, A. C. Stove, N. S. Dalal, Transition<br />

Metal Substituted Polyoxotungstates <strong>and</strong> Their Unique Magnetic Properties,<br />

Extended Abstract- Fourth International Conferences on Inorganic Materials, <strong>University</strong><br />

<strong>of</strong> Antwerp, Belgium, September 19-21, 2004, P48.<br />

4. F. Hussain, L. Bi, B. Bassil <strong>and</strong> U. Kortz*, Discrete Polyoxoanions: Nanomolecular<br />

structures with Multiple Functions,Extended Abstract-7th Iinternational Conference<br />

on Nanostructured Materials, Wiesbaden, Germany, June 20-24, 2004, p.161<br />

5. F.Hussain <strong>and</strong> U.Kortz*, Novel Hybribs Of Inorganic-Organic Polyoxometalates,<br />

1st Open Graduate Students Conference, International <strong>University</strong> Bremen, Bremen.<br />

Germany.<br />

6. F. Hussain <strong>and</strong> P. Selvam*, Tertiary butylation <strong>of</strong> phenol over versatile solid acid<br />

catalysts H-GaMCM48, Extended Abstract-1st Indo-German Conference on Catalysis,<br />

IICT-Hyderabad, Feb 6-8, 2003, PO-32.<br />

156


CONFERENCE / SYNOPSIA / WORKSHOPS AT-<br />

TENDED<br />

1. 1st Indo-German Conference on Catalysis, February 6-8, 2003, IICT-Hyderabad,India.<br />

2. Conference on Electron Microscopy (EMSI), In-House Symposia February 19-22,<br />

2003, I.I.T. Bombay, Powai, India.<br />

3. 1st Open Graduate Students Conference, International <strong>University</strong> Bremen, Bremen,<br />

Germany, October 2003.<br />

4. 7th International Conference on Nanostructured Materials, Wiesbaden, Germany,<br />

June 20-24, 2004.<br />

5. Fourth International Conferences on Inorganic Materials, <strong>University</strong> <strong>of</strong> Antwerp,<br />

Belgium, September 19-21, 2004.<br />

6. Norddeutsches Doktor<strong>and</strong>en- Kolloqium der anorganisch-chemischen Institute, Organisation<br />

vom Institute fur Anorganische Chemie der Christian-Alberchts-Universitat<br />

zu Kiel, Hamburg, Germany, Sept 30-Oct 01, 2004, p12.<br />

157

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!