Ibsa Naseer - Thesis

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 86

Synthesis of Polyoxometalate Complexes and their Application in

Electrochemical Water-splitting

By

Ibsa Naseer

Regn. No. 00000365111

This thesis is submitted to the National University of Sciences and Technology, Islamabad,

in partial fulfillment of the requirements for the degree of

Master of Science (MS) in

Chemistry

Supervisor: Prof. Manzar Sohail

Department of Chemistry, School of Natural Sciences

National University of Science and Technology (NUST)

Islamabad, Pakistan

(2024)
“C’est le temps que tu as perdu pour ta rose

qui fait ta rose si importante”

Antoine de Saint-Exupéry

Le Petit Prince

i
Dedication

To my parents, my mom, my first and favorite teacher, and forever cheerleader, my dad, who

never gave up on us, and my siblings, for never-ending laughter

ii
Acknowledgment

While I sit down here to write this acknowledgment page after completing my thesis, I’m once

again reminded of ‘how infinite is Allah’. Words can’t describe the gratefulness and humbleness

I feel right now that Allah has granted me the strength, resources, and knowledge to complete

another chapter of my life.

I want to thank the School of Natural Sciences (SNS), NUST for this amazing opportunity to

complete my master's degree in such a prestigious institute. I’d especially like to honor my

worthy supervisor Prof. Dr. Manzar Sohail for his excellent guidance, skills, and patience with

me. None of this would have been possible without his guidance. I am also thankful to my GEC

members Dr. Muhammad Adil Mansoor and Dr. Faheem Amin for their help and

knowledgeable insight in polishing my work. I want to extend my gratitude to SNS faculty and

staff especially laboratory staff for always being available in time of need.

My family and friends have been my pillar of support and encouragement throughout this

journey, for which I am incredibly thankful. I want to express my gratitude to my family for their

unending love, tolerance, and faith in me—even on the days when I found it difficult to see

myself, they never doubted my potential and strength. My parents Naseer Ahmed Khursheed

and Shamiza Kousar, you both are superhumans and I’m honored to have been born as your

daughter. You have consistently provided me with solace and inspiration. My siblings, thank you

for being my safe harbor and for empowering me to fly high, knowing I always have a secure

place to return. This achievement is as much yours as it is mine.

To my friends and colleagues of Lab-110, I express my gratitude for your unending love, and

cooperation and for providing me with several stress-relieving breaks that allowed me to

iii
decompress. In addition to making my days happier, your friendship has been a key factor in my

achievement. I want to acknowledge my seniors Amna Altaf, Hafiza Komal Zafar, Sara

Zainab, Ghulam Murtaza, and others for helping me learn all the big and small things. Last but

not least, my friends Aiman, Sadam, Ifra, Sania, and Nida who brilliantly mastered the art of

driving me insane—just enough to keep me perfectly sane through this entire journey. Thanks for

the madness that brought the method to my days!

Regards

Ibsa Naseer Ahmed.

iv
Abstract
This thesis describes the synthesis, structure characterization, and electrochemical properties of

new Rubidium salt of Cobalt-based Polyoxometalate (POM) and Sodium salt of Chromium-

based POM. The complexes Rb2.5[(Co(H2O)6)0.5(PW12O40)].16H2O (Rb-1) and

[(Cr3O(O2CC(CH3)3)6(H2O)3]3[P2W18O40]. xH2O (Na-1) were successfully synthesized in a

single-pot reaction. The synthesis was done using pre-synthesized [Co2(μ-

OH2)(O2CCMe3)4(HO2CCMe3)4], [(Cr3O(O2CC(CH3)3)6(H2O)3](O2CCMe3)3 and tri-lacunary

Keggin anion Na9[A-α-PW9O34].7H2O. The structure of Rb-1 and Na-1 was verified through

FTIR and single-crystal X-ray crystallography. The crystal structure revealed a cubic crystal

system with a central Co (II) surrounded by six water molecules for Rb-1 and a primitive-type

monoclinic crystal structure for Na-1.

Moreover, Rb-1 and Na-1 were deposited on a Ni-foam electrode and functioned as an anode for

efficient OER in water splitting. The Rb-1/Ni-Foam electrode showed excellent overpotential

values of 207 mV at 10 mA.cm-2 and 390 mV at 100 mA.cm-2 current densities. The electrode

had fast kinetics, as indicated by a low Tafel Slope of 51.76 mVdec-1 and a small charge transfer

resistance (RCT) of 5.904 ohms. The Na-1/Ni-Foam electrode showed overpotential values of

226 mV at 10 mA.cm-2 and 310 mV at 100 mA.cm-2 current densities. The electrode had fast

kinetics, indicated by a low Tafel Slope of 54.6 mVdec-1 and an RCT of 12.17 ohms. Both the

catalysts exhibited remarkable stability over an 8-hour test period, emphasizing their practical

value and possible uses in energy applications.

v
Table of Contents

Abstract ........................................................................................................................................... v

Chapter 1: Introduction ................................................................................................................... 1

1.1. Polyoxometalates (POMs) .............................................................................................. 1

1.1.1. A History of Polyoxometalates ................................................................................... 2

1.1.2. Structure & Nomenclature .......................................................................................... 4

1.1.3. Lacunary Structures .................................................................................................... 8

1.1.4. Synthesis ................................................................................................................... 10

1.1.5. Self-Assembly ........................................................................................................... 12

1.1.6. Counterions ............................................................................................................... 14

1.2. Classical POMs ............................................................................................................. 14

1.2.1. Lindqvist ................................................................................................................... 14

1.2.2. Strandberg ................................................................................................................. 15

1.2.3. Keggin ....................................................................................................................... 16

1.2.4. Wells-Dawson ........................................................................................................... 17

1.2.5. Anderson-Evans ........................................................................................................ 19

1.3. Single Crystal XRD (SCXRD) ..................................................................................... 20

1.3.1. Sampling, Instrumentation, and Data Collection ...................................................... 22

1.3.2. Structure Determination Methodology ..................................................................... 22

vi
1.3.3. Applications of Single-Crystal XRD ........................................................................ 24

1.4. Introduction to Electrochemical Water Splitting .......................................................... 26

1.4.1. Oxygen Evolution Reaction ...................................................................................... 26

1.4.2. Reactions involved in OER:...................................................................................... 28

Chapter 2: Literature Review ........................................................................................................ 31

Chapter 3: Experimental Section .................................................................................................. 36

3.1. Chemicals and Solvents: .................................................................................................... 36

3.2. Apparatus and Glassware: .................................................................................................. 36

3.3. Procedure of Precursors ..................................................................................................... 36

1. Synthesis of trilacunary Keggin POM precursor Na9[A-α-PW9O34].7H2O ....................... 36

2. Synthesis of [Co2(μ-OH2)(O2CCMe3)4(HO2CCMe3)4] ..................................................... 38

3. Synthesis of [(Cr3O(O2CC(CH3)3)6(H2O)3](O2CCMe3)3] ................................................. 39

3.4. Procedure of POM Complexes ........................................................................................... 40

1. Synthesis of Archimedean rhombicuboctahedron Rb2.5[(Co(H2O)6)0.5(PW12O40)].16H2O40

2. Synthesis of [(Cr3O(O2CC(CH3)3)6(H2O)3]3[P2W18O62]. xH2O ........................................ 41

3.5. Fabrication of Electrodes ................................................................................................... 42

Chapter 4: Results and Discussion ................................................................................................ 44

4.1. Single Crystal X-ray Diffraction (SCXRD) ....................................................................... 44

4.1.1. Single crystal XRD of Rb-1......................................................................................... 44

4.1.2. Single crystal XRD of Na-1......................................................................................... 47

vii
4.2. FT-IR .................................................................................................................................. 50

4.2.1. FTIR Spectrum of Rb-1 ............................................................................................... 50

4.2.2. FTIR Spectrum of Na-1 ............................................................................................... 51

4.3. Thermogravimetric Analysis (TGA) .................................................................................. 52

4.4. Oxygen Evolution Reaction (OER).................................................................................... 54

4.4.1. Electrochemical activity of Rb-1 ................................................................................. 54

4.4.2. Electrochemical activity of Na-1 ................................................................................. 59

Chapter 5: Conclusion................................................................................................................... 62

Chapter 6: References ................................................................................................................... 64

viii
Table of Figures
Figure 1: Timeline of POM history. ................................................................................................ 4

Figure 2: Structure of a POM.......................................................................................................... 5

Figure 3: Ball and stick and polyhedral representation of a POM. ................................................. 6

Figure 4: Polyhedral units of POMs made by different metals. ..................................................... 7

Figure 5: Different forms of interaction between terminal O-atoms. ............................................. 8

Figure 6: Lacunary POM structures. ............................................................................................. 10

Figure 7: Synthesis process of POMs. ...........................................................................................11

Figure 8: Condensation of Pd84. .................................................................................................... 13

Figure 9: A Strandberg POM. ....................................................................................................... 16

Figure 10: Keggin POM structure................................................................................................. 17

Figure 11: Dawson-type POM structure. ...................................................................................... 18

Figure 12: Dawson-type POM isomers......................................................................................... 19

Figure 13: Structure of Anderson-Evans POM. ............................................................................ 20

Figure 14: Schematic diagram of Electrochemical water splitting. .............................................. 27

Figure 15: Synthesis scheme for trilacunary POM precursor Na9[A-α-PW9O34].7H2O............... 38

Figure 16: Synthesis scheme for [Co2(μ-OH2)(O2CCMe3)4(HO2CCMe3)4]. ................................ 39

Figure 17: Synthetic diagram for [(Cr3O(O2CC(CH3)3)6(H2O)3](O2CCMe3)3]. ........................... 40

Figure 18: Synthesis scheme for Rb2.5[(Co(H2O)6)0.5(PW12O40)].16H2O..................................... 41

Figure 19: Synthetic scheme for [(Cr3O(O2CC(CH3)3)6(H2O)3]3[P2W18O62]. xH2O .................... 42

Figure 20: Washing of Ni-Foam and fabrication of electrodes. .................................................... 43

ix
Figure 21: (a) Packing of Rb-1 (b) Cubic representation (c) Polyhedral representation of Co (II)

containing rhombicuboctahedron Archimedean solid coordination sphere. (Color coding, W:

blue, O: red, P: yellow, Co: green, Rb: purple, H: grey) .............................................................. 45

Figure 22: Polyhedral / Ball and stick representation of different building blocks and different

directional views of Rb-1. (Color coding, W: blue, O: red, P: yellow, Co: green, Rb: purple, H:

grey) .............................................................................................................................................. 47

Figure 23: Crystal Structure of [(Cr3O(O2CC(CH3)3)6(H2O)3]3[P2W18O62]. xH2O and Dawson-

type POM anion. ........................................................................................................................... 48

Figure 24: Wheel-shaped structure showing the metal centers. ................................................... 49

Figure 25: Monoclinic unit cell packing of Na-1.......................................................................... 49

Figure 26: FTIR Spectrum of Rb-1. .............................................................................................. 51

Figure 27: FTIR Spectrum of Na-1. .............................................................................................. 52

Figure 28: TGA analysis of Rb-1. ................................................................................................. 53

Figure 29: (a) Cyclic Voltammogram for Rb-1/Ni-Foam at different scan rates ranging from 5

mV s-1 to 100 mV s-1. (b) Linear Sweep Voltammogram of Rb-1/Ni-Foam (black) and bare Ni-

Foam (red). (c) Tafel Slope for Rb-1/Ni-Foam (black) and bare Ni-Foam (red). (d) A comparison

of Rb-1/Ni-Foam (black) overpotential values and bare Ni-Foam (red) at 10 mA, 50 mA, and

100 mA, respectively. ................................................................................................................... 56

Figure 30: (a) EIS for Rb-1/Ni-Foam with fitted circuit. (b) Cdl for Rb-1/Ni-Foam from linear

fitted data. ..................................................................................................................................... 58

Figure 31: (a) Chronopotentiometry scan for Rb-1/Ni-Foam for 8 hrs. (b) XRD pattern, before &

after checking stability via Chronopotentiometry. ........................................................................ 59

x
Figure 32: (a) Linear Sweep Voltammogram of Na-1/Ni-Foam (red) and bare Ni-Foam (black).

(b) A comparison of Na-1/Ni-Foam (black) overpotential values and bare Ni-Foam (red) at 10

mA, 50 mA, and 100 mA, respectively. (c) Tafel Slope for Na-1/Ni-Foam (red) and bare Ni-

Foam (black). (d) EIS for Na-1/Ni-Foam with fitted circuit. ....................................................... 60

Figure 33: Chronopotentiometry scan for Na-1/Ni-Foam for 8 hrs. ............................................ 61

xi
List of Abbreviations

No. Abbreviation Full Name No. Abbreviation Full Name


FTIR Fourier Transform
1. POM Polyoxometalate 14.
Spectroscopy Infra-red Spectroscopy
Single-Crystal X-Ray Thermogravimetric
2. SCXRD 15. TGA
Diffration Analysis
Powder X-Ray Water Oxidation
3. pXRD 16. WOC
Diffraction Catalysis
Metal-organic f
4. MS Mass Spectrometry 17. MOF
Framework
Nuclear Magnetic
5. NMR 18. TBA Tetrabutylammonium
Resonance
Least Unoccupied
6. LUMO 19. DI water Deionized Water
Molecular Orbital

7. CV Cyclic Voltammetry 20. R.M.S. Root Mean Square

Reversible Hydrogen
8. DMF Dimethylformamide 21. RHE
Electrode

9. 3D Three-Dimensional 22. DC Direct Current

Linear Sweep
10. CIF Crystal Information File 23. LSV
Voltammetry
Electron Impedence
11. DNA Deoxyribonucleic Acid 24. EIS
Spectroscopy
Oxygen Evolution
12. OER 25. CP Chronopotentiometry
Reaction
Hydrogen Evolution Electro-chemical
13. HER 26. ECSA
Reaction Surface Area

xii
Chapter 1: Introduction
Nearly two centuries ago [1], the initial reports on polyoxometalates (POMs) marked the

beginning of their significant presence across diverse scientific disciplines. Today, POMs garner

substantial attention and find applications in various fields such as catalysis [2], materials science

[3], electrochemistry [4, 5], and medicine [6, 7]. Representing a class of discrete metal-oxo

cluster anions, or polyoxoanions, POMs are constructed from early transition metals, most

commonly those in groups 5 and 6, exhibiting their higher oxidation states, especially Mo(VI),

W(VI), and V(V) [8]. Their extensive structural diversity, remarkable stability, and adaptable

chemical and physical properties position them as excellent building blocks for the development

of functional materials [9, 10].

1.1.Polyoxometalates (POMs)

Polyoxometalates are inorganic compounds consisting of transition metals in their high oxidation

states and oxygen. These heteropolyanions come together to form interconnected structures

which gives rise to the formation of discrete clusters [11]. These clusters, typically carrying a

negative charge, exhibit a range of sizes from molecular to nano-scale [12], giving rise to diverse

architectures that garner significant attention. The synthesis of POMs involves easy and simple

one-pot reactions, wherein basic building blocks act on the principle of spontaneously happening

self-assembly and form complex polyanions that are subsequently isolated via crystallization.

Despite a partial understanding of the process of self-assembly, scientists can manipulate and

modify POM structures by changing simple experimental conditions like concentration, pH, and

temperature since POMs are sensitive to changes like these. Beyond their intricate structures,

POMs are intriguing due to their distinctive electronic and redox properties arising from the

inherent characteristics of the transition metals, most commonly used metals being Tungsten,

1
Vanadium and Molybdenum. The composition of POMs, influenced by other elements and

balancing cations, can vary significantly, impacting their electronic behavior. Consequently, the

diverse electronic features of POMs make them useful in electronics [13, 14] catalysis [15, 16]

materials science [17-20] magnetism [21, 22] and medicine [23] as demonstrated in various

studies.

1.1.1. A History of Polyoxometalates

The first recorded sighting of a POM was "molybdenum blues" that was associated with

ilsemmanite, a mineral found in nature. Unidentified blue molybdenum oxides were originally

observed by Swedish scientist Carl Wilhelm Scheele in 1778, marking the first known

occurrence of POM molecules [24]. When ammonium molybdate and phosphoric acid reacted, in

1826, Jöns Jacob Berzelius saw yellow precipitates that was subsequently determined to be

(NH4)3[PMo12O40].xH2O, also called the Keggin, a traditional POM structure [1]. Given the

analytic limits of the time, the Swiss scientist Jean-Charles Galissard de Marignac demonstrated

the composition of elements in 1862 when he used titration method. Galissard de Marignac

postulated the presence of two geometries, which are presently known as the α- and β- species

while being unable to identify the precise structure [25, 26].

The basic ideas of coordination compounds were not developed until 1893, when Swiss scientist

Alfred Werner postulated the notion that metal ions in a molecule might be connected to more

than one oxygen atom [27]. Werner's ideas were extended to metal oxide clusters by Miolati

fifteen years later [28]. The Miolati-Rosenheim hypothesis (1917) built on this work, proposing

that [MO42-] or [M2O72-] replace the oxo ligands of a parent acid to generate heteropolyacids

[29]. Ten years later, Linus Pauling offered a different idea in which he suggested that the

clusters were formed by the surrounding octahedral [MO6] units of a center tetrahedral [XO4]

2
unit. Pauling's restriction, which said that these units only permitted oxo ligands to share corners

rather than edges or faces, has been contested, though, since current research indicates that

polyoxometalate complexes frequently exhibit this kind of sharing [30].

The development of Single-crystal XRD (SCXRD) in 1913 was a crucial step in obtaining a

more comprehensive knowledge of these clusters' structures [31, 32]. The pioneering elucidation

of a polyoxometalate structure occurred in 1933 when J. F. Keggin solved the structure of 12-

phosphotungstic acid [33, 34]. This revelation showcased a structure comprising twelve

polyhedra, sharing edges and corners. Although it took approximately a decade for this

breakthrough to gain full acceptance, the solved structure paved the way for predicting the

structures of several other clusters before their confirmation through XRD, including the Wells-

Dawson [35] and the Anderson-Evans structures [36, 37].

A selected number of groups of research fellows, including Souchay in Paris and Baker and Pope

at Georgetown, played pivotal roles in advancing the concepts and techniques that form the basis

of polyoxometalate (POM) chemistry [38, 39]. The foundation for the field's advancement was

built by these pioneers. Significant discoveries include the identification of large clusters like the

"blue lemon" and wheel-shaped molybdenum blues [12] that not only confirmed POMs as

nanomaterials [40] but also sparked general interest in the field. More recently, the combination

of analytical techniques such as MS, NMR, and computational modeling has allowed us to gain

even more insight into the structural properties and assembly mechanisms of polyoxometalates

[41]. The key turning points in polyoxometalate chemistry that are described here are also briefly

displayed in the form of a timeline.

3
Figure 1: Timeline of POM history.
1.1.2. Structure & Nomenclature

The nomenclature employed in the realm of POMs lacks complete standardization, though

several frequently used phrases exist. The term "polyoxometalate" itself denotes a compound

featuring three or more metal centers interconnected through bridging O-atoms. While an

alternative, self-explanatory term, "metal-oxygen cluster," exists, it is seldom utilized [40].

4
Instead, the more prevalent terms encountered are "polyanion" or, "polyoxoanion." While these

terms acknowledge that POMs are generally anionic, they do not emphasize their metal-

containing nature [35, 41].

Addenda are the atoms present in POMs that are usually early transition metals at their highest

oxidation states, usually in the configurations of d0 and d1, such as MoVI and WVI [42].

Figure 2: Structure of a POM.


Heteroatoms are the atoms that are not addenda but remain within the structure of POM,

commonly belonging to elements in the p-block mostly Silicon and Phosphorus, though their

inclusion is not limited to these elements. Virtually any element, whether metal or non-metal, can

serve as a heteroatom, provided it can form bonds with three or more atoms. A heteroatom may

be classified as either primary or central, playing a crucial role in the POM structure, or

secondary or peripheral, such that its removal does not disrupt the integrity of the cluster. When a

secondary atom leaves its place, a cavity called a lacuna forms, and the POM is called a lacunary.

5
Polyoxometalates may be divided into two distinct categories: isopolyoxometalates and

heteropolyoxometalates. These groups are represented by the generic formulae [MnOy]p and

[XaMnOy]p respectively, where p is the cluster's overall charge, which is usually negative, and a <

n, M = addenda, and X = heteroatoms. Whereas heteropolyoxometalates have both secondary

heteroatoms at the peripheral and primary heteroatoms that are structurally important,

isopolyoxometalates only contain primary heteroatoms and oxygen atoms.

Figure 3: Ball and stick and polyhedral representation of a POM.


Finally, a POM cluster's oxygen atoms can be classified as either bridge or terminal. While a

terminal O-atom establishes a double bond with just one metal center on the cluster's border, a

bridging O-atom is connected to 2 or more addenda. A polyhedron is a geometric configuration

of Oxygen atoms around a single metal core. It is frequently depicted in images to help see the

POM molecule more clearly (Figure 3). Generally speaking, these [MOx] polyhedral units, are

regarded as the basic structural component (or synthon) of a POM. Even in the largest and most

complex POM clusters, they always resurface.

6
Figure 4: Polyhedral units of POMs made by different metals.
To understand why metal oxides, unlike the iron oxides that cause rust, form discrete clusters

rather than longer structures, coordination theory and the nature of metal-to-oxygen bonding are

crucial [27].

In this case, the trans effect is important because it describes how the ligand on the other end of

the metal weakens the metal-ligand interaction in an octahedral or square planar geometry

(Figure 4) [42]. Because the ligands linked to each other in a trans manner occupy the same

orbital of the central metal, their capacity to receive and donate electrons affects the stability of

their bonds. This effect may result in kinetic or structural modifications, increasing the bond's

labile and reactive properties.

By shortening M=O bonds and distorting polyhedrons, the trans-effect in polyoxometalates

brings addenda atoms in proximity to terminal O-atoms. As a result, the cluster becomes

7
polarized throughout, with the inner core being somewhat electron-rich and the outer surface

having a larger positive charge than anticipated [42]. The flexibility of the electron-rich core

allows for the accommodation of heteroatoms with a range of shapes, which accounts for the

wide variety of POM structures. Rather than from simple polyhedra synthons, bulkier POM

building blocks usually result from a joining of simpler parts.

Figure 5: Different forms of interaction between terminal O-atoms.


As Figure 5 shows, corner, edge, or face sharing can cause the condensation process in an

environment full of acid to result in polymerization of [MOx] (usually MO6) building blocks [43–

45]. This leads to the possibility of three types of POM clusters: type I, which has octahedra with

a single M=O terminal link; type II, which has two such bonds; and type III, which is a

combination of the two [46]. This categorization takes into account variations in chemical

behavior in addition to morphological distinctions.

Because LUMO of these clusters is non-bonding, type I & type III POMs—the molybdenum

blues and browns, for example—are reversibly redox-active [14, 42–44]. Conversely, type II

clusters tend to break down during reduction because of dioxo units present in cis positions [38,

43].

1.1.3. Lacunary Structures

Lacunary structures are POM structures without heteroatoms, such as the Wells-Dawson or

Keggin structures. The initially formed clusters themselves can be used to create these lacunary

8
structures; this is often accomplished by modifying the pH to significantly destabilize the POM.

As an alternative, they can be made in a one-pot reaction in circumstances that are detrimental to

the creation of the whole structure. Generally speaking, the removal of an extra metal center

never happens at the opposing ends of the cluster and always happens next to an addenda site

that has previously been cleared.

Lacunary structures allow extra addendum metals to occupy the lacuna, hence enabling the

creation of mixed-metal POMs. By condensing lacunary units, they also aid in the creation of

more complex POM structures. Furthermore, POM hybrids may be formed using lacunary

structures as a building block. This adaptability results from the unequal charge distribution in a

lacunary POM, where a lacuna's increased reactivity with electrophiles is caused by the exposed

oxygen atoms' increased nucleophilicity.

9
Figure 6: Lacunary POM structures.
1.1.4. Synthesis

The Lewis acid character of the metal centers increases in acidified solution form of transition

metal oxyanions, [MOx]n-, which triggers condensation events that unite the metal centers

forming bigger molecules. Structure of these complexes vary significantly depending on

circumstances.

10
Figure 7: Synthesis process of POMs.
The most common method for synthesizing POMs involves a one-pot reaction happening in an

aqueous medium. There are several options for the circumstances of the experiment, such as pH

& temperature, duration, and concentration, all of which can have a substantial effect on the

results of the experiment [8].

Beyond the conventional synthesis process, alternative approaches can be highly effective. While

water remains the most frequently employed solvent for the synthesis of POMs, conducting

experiments for synthesis in polar organic solvents—particularly MeCN, DMF, or

dichloromethane—can lead to the formation of distinct POM clusters compared to an equivalent

aqueous setup [42-46].

Many POMs are synthesized at relatively low temperatures, including room temperature.

However, deliberately inducing reactions far from equilibrium conditions can lead to the

crystallization of metastable POMs that would not be obtained under normal circumstances.

11
Reactions are mostly carried out in hydrothermal or solvothermal conditions, and at greatly

increased temperatures and pressures, to accomplish this. This process has been effectively

shown using a variety of solvents, such as ionic liquids [47-49], organic solvents [50-52], and

water [53-55]. Unconventional approaches, such as those employed by Cronin and colleagues,

include continuous flow methods [56, 57] or utilizing purpose-built 3D-printed cartridges [58,

59].

Crystallization plays a pivotal role in POM synthesis, as well as in the processes of purification

and isolation. For certain polyoxometalates, crystallization is the sole method of obtaining them,

and, in many cases, the desired clusters are formed under the specific guidance of the

crystallization process itself. Certain crystals might only develop at one time of the year, or it

might be difficult to replicate them outside of the lab where they were first isolated [60].

The POM area frequently faces the reproducibility challenge, which arises from the necessity of

experimental accuracy. This highlights the significance of precise recording of experimental

techniques. Some POMs can be synthesized quite quickly and have numerous legitimate

synthesis routes, whereas many only form under a highly particular and limited range of reaction

circumstances. Determining these parameters may be difficult, and even a seasoned researcher

may find it difficult to synthesize a compound without the direct supervision and guidance of

another specialist in the area who is familiar with that specific POM.

1.1.5. Self-Assembly

Much research using X-ray crystallography has been conducted to explore POMs assembly

processes, which has resulted in the creation of several hypotheses. Recently, research using

MS has given rise to a more solid knowledge of these clusters' self-assembly processes [61, 62].

For example, a proposed mechanism suggests sequential metal-center assembly with alternating

12
steps exhibiting endo and exothermic pathways until a large exothermic reaction creates the

possibility of cluster formation [63].

In a different instance, Cronin and associates examined and suggested that the Pd84 wheel is

formed gradually over a few days from 14 Pd6 subunits, thanks to this method (Figure 8). A

similar procedure was then used to examine the development of a sequence of smaller Pd wheels

[64, 65].

Figure 8: Condensation of Pd84.


The Lindqvist type of POMs have a complicated assembly process, which emphasizes how

difficult it is to comprehend how bigger clusters of POMs self-assemble. For POM chemists, this

lack of conceptual understanding poses a major hurdle. Although directed self-assembly entails

planning a response with a particular target in mind, this approach is not infallible due to the

unpredictable nature of POM production. An equally legitimate method for investigating the

chemical space of POMs is to create reactions by adjusting synthetic conditions without any

preconceived notions about the structure. This kind of undirected or spontaneous self-assembly

makes it possible to find previously undiscovered pockets of POM cluster families [66, 67].

13
1.1.6. Counterions

Since the majority of polyoxometalates are negatively charged, cations are necessary to keep the

charge balance in both structures and solutions. These counterions range in hardness from soft to

hard, including organic molecules and alkali metals like Na+ and K+ as well as derivatives of

amines that are either protonated or quaternary salts. The structure and crystallization of species

in a polyoxometalate solution can be significantly impacted by these counterions. Apart from

controlling the formation of clusters, cations affect the supramolecular microstructures that POM

species self-assemble into [68-72].

1.2. Classical POMs

In the realm of polyoxometalate (POM) research, certain structures emerge consistently because

of their stability and repeatability. These classic POMs not only manifest with diverse addenda as

well as heteroatoms but also serve as the foundational substructures for a majority of larger

polyoxometalates and their derivatives. Additionally, these structures are the basis from which

typical POM hybrids originate. The upcoming sections will provide a concise introduction to

some of these prototype POMs.

1.2.1. Lindqvist

The smallest and most basic archetype among polyoxometalates is a "super-octahedral" assembly

comprising 6 edge-sharing [MO6] units, each containing 6 metal centers, 12 bridging oxo

ligands, and 6 terminal oxo ligands. This assembly was first described by Swedish chemist

Ingvar Fritz Lindqvist in 1950 [73] (Figure 3). The Lindqvist structure is an iso polyoxometalate

that may be created with the addition of Nb [74, 75], Ta [76], Mo [76, 77], and W [78]. Its typical

formula is [M6O19]n-. Although V can also be used to generate the Lindqvist structure, [V6O19]8-,

it tends to break down into a hybrid in the absence of organic ligand stability [79, 80].

14
Like other traditional POM categories, the Lindqvist structure is used as a building block for

more complex POM structures. Examples of this use are the synthesis of a polyoxometalate-

based SMM composed of a Lindqvist dimer connected by lanthanide, a coordination polymer

including mixed metals, and a sandwich structure containing cobalt.

1.2.2. Strandberg

The structure and effective isolation of Na6Mo5P2O23(H2O)13, subsequently known as the

Strandberg polyoxometalate, Mo5X2O236-, was accomplished by Rolf Strandberg in 1973. This

POM is made up of a ring with five [MoO6-] octahedral subunits that share all but one corner.

Three bridging oxo ligands are terminated at both ends connecting tetrahedral [XO4-] subunits.

The resultant structure has a double rotating axis of symmetry, and the clusters are spherical due

to the twelve extending oxygen atoms. To balance the charge, counterions usually join these

polyanions to create a crystalline structure. Notably, the Strandberg has a greater charge density

and is comparatively smaller in size than other POMs that are regularly investigated [81].

In the Strandberg-type POM, phosphorus atoms often make up the two heteroatom units [XO4-],

while other atoms are also possible. There aren't many structures made using the Strandberg-type

POM, despite how simple it is to make hybrids with this structure. But in the last several years,

several investigations have looked at other macrostructures based on the Strandberg structure. A

few papers have mentioned biological or catalytic activity together with antiferromagnetic

characteristics [82-84].

15
Figure 9: A Strandberg POM.
1.2.3. Keggin

Given its straightforward structure and range of derivatives, the Keggin structure is an example

of a hetero-polyoxometalate that is commonly encountered in polyoxometalate chemistry. The

Keggin cluster is usually made up of Mo, W, or V with a heteroatom of a p-block element.

However, structures made completely of other transition metals have also been reported; these

structures are not technically POMs. The Keggin structure has a core heteroatom encircled by 12

metal centers plus 40 oxygen atoms, with the formula [XM12O40]p-. Of them, twelve create

terminal oxo ligands, 24 form bridges between metal centers, and four connect metal centers to

the heteroatom [85].

Four [M3O9] subunits are revealed when dissecting the structure. By sharing a corner atom and

tetrahedrally coordinating with the central heteroatom through an oxygen atom shared by the

three metal centers of each subunit, these subunits create a cage. The Keggin cluster isomer is

16
determined by the orientation of subunits concerning one another; stability decreases as the W-

O-W bond strain rises.

Figure 10: Keggin POM structure.


1.2.4. Wells-Dawson

This configuration, known as the Wells-Dawson structure (abbreviated as "Dawson" and

symbolized by [X2M18O62]p-), arises in circumstances that are comparable to those that favor the

Keggin structure. It looks like two fused Keggins, neither of which has a [M3O9] triad. Instead,

they form a central belt with 12 octahedral addenda that share a corner, and two edge-sharing

[M3O13] units on top of it. This structure has two heteroatomic tetrahedral sites [XO4], where X

is often one of the following: Si, P, S, or As. Eight heteroatom-bonded oxygens, 36 bridging oxo

ligands between metal centers, and 18 terminal oxo ligands make up the Dawson structure's total

of 62 oxygen atoms. Interestingly, the addenda's surroundings inside the Dawson structure are

17
not comparable, resulting in unique features like the increased vulnerability of the center belt to

reduction (Figure 11) [86-91].

Figure 11: Dawson-type POM structure.


The Dawson structure has numerous isomeric forms, as do many classic POM structures; six

possibilities were suggested by Baker and Figgis. The [M3O3] caps rotate 60° concerning one

another, giving rise to the α, β, and γ isomers.

18
Figura 12: Dawson-type POM isomers.
1.2.5. Anderson-Evans

The structure of a Molybdenum POM with six edge-sharing polyhedra organized in a ring

around a core heteroatom that is octahedrally coordinated was predicted by John Stuart Anderson

in 1937. About ten years later, Howard T. Evans Jr. isolated and validated the suggested

structure, K6[TeMo6O24], using crystallography. Later, several POMs with a similar structure

were found; they are denoted as [Hy(XO6)M6O18]n-, where n = 2–8, y = 0–6, M = addenda atoms

19
of MoVI or WVI, and X = core heteroatom. Even though both people are honored in the POM's

name, it is typically shortened to "Anderson" for ease. Since its original discovery by Anderson,

this cluster-type with different centers has been synthesized using molybdenum addenda and, to

a lesser extent, tungsten addenda. Twelve of the cluster's 24 oxygens are terminal oxo ligands

that are positioned around the edge, while the remaining six are triple-bridged, double-bridged,

and bridging oxo ligands (Figure 13). Depending on the heteroatom present, the Anderson

structure can be either protonated (B-type) or non-protonated (A-type). The Anderson structure's

chemical and physical characteristics are mostly determined by heteroatom, cation, and

hybridization [92-94].

Figura 13: Structure of Anderson-Evans POM.


1.3.Single Crystal XRD (SCXRD)

An essential tool for examining the structural characteristics of materials is diffraction testing

which uses X-rays or other particles like electrons and neutrons. These lab investigations usually

utilize X-rays with wavelengths between 0.5 to 2.5 Ångströms, which are similar to the smallest

gaps between atoms in a material. This wavelength similarity is important because it makes it

possible to use diffraction patterns to observe atomic configurations in detail [95].

20
A concentrated beam of X-rays, electrons, or neutrons can scatter waves when it strikes a

material, especially a crystalline solid, because of the interaction between the incoming waves

and the substance's atomic structure. A diffraction pattern is produced as a result of this

scattering, and it is simply a representation of the atomic structure of the material projected onto

a region called reciprocal space.

However, this pattern needs to be mathematically changed back from reciprocal space to direct

space to be represented in a format that is suitable for understanding the three-dimensional

distribution of atoms within the crystal. Through this process of transformation, scientists can

piece together a clear picture of the atomic lattice of the crystal, showing the exact placement

and orientation of the atoms. Scientists may better understand and alter material characteristics

for a variety of applications because of this in-depth structural information [96-98].

While it is mostly used to investigate solid samples, XRD examination is a flexible approach that

may be used for a wide range of materials. Under some circumstances, XRD analysis may also

be performed on liquid samples and, in rare cases, gasses. The properties of the solid under

investigation have a significant impact on the kind and quantity of information obtained from an

XRD experiment.

If a sample is optically transparent, has no grain boundaries, and has a continuous lattice

structure, it is categorized as single-crystal or monocrystalline. The best range of crystal sizes for

SCXRD investigation is between 0.1 and 0.2 millimeters. When it comes to clarifying and

improving the crystalline structure of novel materials, single-crystal XRD is especially useful.

Over the last fifteen years, there have been notable improvements in the success rates of this

non-destructive approach due to breakthroughs in experimental equipment and methods for

solving and improving crystal structures. Comprehending a material's crystal structure is

21
essential since it immediately affects and clarifies the material's qualities, supporting the creation

and use of these materials in a variety of applications [99, 100].

1.3.1. Sampling, Instrumentation, and Data Collection

The unfractured, optically transparent single crystals that are employed are usually between 0.1

and 0.2 millimeters in size in all the different dimensions of space. Each crystal's compatibility is

assessed by turning the microscope stage 90 degrees and seeing if the light consistently goes out.

After being chosen, the crystal is either put in a loop that holds a certain oil or firmly affixed to

the end of a thin fiber of glass using epoxy or cement. The crystal is then meticulously aligned

with the direction of the beam. The stability of the crystals must be evaluated as they could be

susceptible to light, air, or humidity, or they might lose their crystallization solvent. When these

types of sensitivities are found, more care must be taken. For example, to maintain stability and

integrity, data gathering may need to be done at low temperatures, or crystals might be put within

sealed glass capillaries.

The crystal is put on the diffractometer, and then the Ewald sphere is adjusted and the distance to

the detector is determined for each measurement. Next, using molybdenum radiation, the

intensity of the diffracted rays is measured. Typically, data between 3 degrees and 30 degrees 2θ

are captured. Depending on the sample's properties and the diffractometer's capabilities, a full

data collection can take three to twelve hours to complete. The unit cell's properties are

computed from the measured intensities. Then, every intensity is indexed to provide an extensive

list of hkl reflections that have been seen.

1.3.2. Structure Determination Methodology

The structural factor is the X-ray intensity in a pattern of diffraction that is exclusively

determined by the crystal structure:

22
Where h, k, and l are the Miller Indices. The total number of atoms in the unit cell is denoted by

N. The position of each atom j, is represented by its scattering factor (fj) and fractional

coordinates (xj, yj, zj).

The diffraction pattern may be used to calculate the electron density function at each location

inside a single unit cell to identify the crystal structure:

Every single value of h, k, and l is included in the total, and V stands for the unit cell's volume.

The resultant computation is an average of the contents of the unit cell, extrapolated from the

entire crystal, as X-rays are scattered by the entire crystal. In practical terms, the calculation of

the electron density yields maps. The positions of the atoms within the cell are shown by the

highest points on these maps. However, the phases of the structural factors are not revealed by

the diffraction pattern; only their amplitudes are. Since the electron density cannot be computed

directly from experimental data, further techniques are required to determine the absent phases.

The phase problem is the name given to this issue.

The most popular approaches for solving the phase problem include direct approaches, which are

essential in chemical crystallography, and approaches that make use of the Patterson function,

which are especially helpful in cases when the structure contains heavy atoms. These approaches

yield phases that need to be further refined because they are approximations. Once the structural

components' phases and amplitudes are identified, a preliminary, although imprecise, electron

23
density map is created. The locations of the atoms inside the structure are then determined using

this map.

The next step is to refine the structure using Fourier synthesis and modify the structural

parameters to improve the agreement between the diffraction pattern's observed and anticipated

intensity. The anisotropic vibration parameters and atomic positions are adjusted during this

refinement process. In this step, the positions of any hydrogen atoms that are a part of the

structure are also determined or calculated. By comparing the computed structure factors with

experimentally determined ones, the refinement process is evaluated for efficacy. The structural

refinement is considered finished when certain essential criteria are satisfied. Such as:

- The agreement factors are low enough.

- The chemical validity of the structural model is met.

- The geometric parameters' estimated standard deviations are kept to a minimum.

- The electron density map's leftover peaks are reduced.

After the structure has been defined and refined, various geometric parameters, including bond

lengths, bond angles, torsion angles, π-stacking interactions, and hydrogen bonds, are analyzed.

Tables and graphical representations of the structure are then prepared. A standard file, known as

a Crystal Information File (CIF), is generated to encapsulate all the structural details. This file is

utilized to assess the quality of the structure and identify any potential issues.

1.3.3. Applications of Single-Crystal XRD

The strong analytical method SCXRD is mostly applied in the fields of chemistry, materials

research, and mineralogy to ascertain the molecular and atomic structure of a crystal. The

24
technique uses the electrons in a crystalline sample to diffraction an X-ray beam, creating a

pattern that may be used to determine the locations of the atoms within the crystal. Applications

of Single-Crystal XRD include mainly the structure determination in Crystals in Organic,

Inorganic, and Biochemistry. In organic compounds, it is crucial to determine the three-

dimensional structure of complex organic compounds because it helps scientists understand the

stereochemistry, conformation, and relative configuration. For industries like pharmaceuticals,

where a drug's structure can have a significant impact on its activity, this knowledge is essential.

In the case of Inorganic Chemistry, understanding the coordination environments and geometry

around metal centers in inorganic complexes is crucial for comprehending their catalytic,

electrical, and magnetic characteristics. Single-crystal XRD aids in this process.

Single-crystal XRD is also a very important technique that is used in Materials Science, where it

is used to determine the structure of materials such as ceramics, semiconductors, and

superconductors, giving an insight into the potential applications of the materials. Not only that,

this technique also provides information about the lattice defects and impurity placements which

play a vital role in the electrical and mechanical properties of materials.

In academic geological studies and practical applications such as mining and resource extraction,

Single-crystal XRD is important for its role in determining the crystal structure of minerals to

help understand their formation conditions and physical properties. In Environmental sciences,

SCXRD is important to identify the crystal phases that participate in environmental issues like

pollution absorption and remediation. In Biochemistry and Molecular Biology, SCXRD is crucial

for its use in the determination of complex and helical structures of biological molecules such as

DNA, proteins, and several enzymes. Apart from these major fields, SCXRD is also used in other

25
application areas such as Catalysis, Forensics and Manufacturing on a wider scale that proves its

worth as one of the state-of-the-art technique used in academic research and industrial area.

1.4. Introduction to Electrochemical Water Splitting

In the study of sustainable energy, electrochemical water splitting is a key procedure that offers a

productive and clean way to produce hydrogen and oxygen gasses. This method uses electrical

energy to break down water (H2O) into its constituent elements, oxygen (O2) and hydrogen (H2).

Two half-reactions are involved in the process: the OER at the anode, where water molecules are

oxidized to generate oxygen gas, and the HER at the cathode, where protons are reduced to

hydrogen gas.

The main reasons electrochemical water splitting is being studied are its possible uses in energy

storage and as a sustainable fuel source. This process can create hydrogen, which may be used as

a clean energy source and substitute for fossil fuels. In order to expedite reaction rates and reduce

total energy consumption, this technique needs electrocatalysts that are effective, long-lasting,

and financially viable. Improvements in electrochemical engineering, materials science, and

catalysis are essential for streamlining this procedure, cutting expenses, and raising the

effectiveness of water electrolysis systems. Because of this, research on electrochemical water

splitting is leading the way toward a sustainable and energy-secure future.

1.4.1. Oxygen Evolution Reaction

Along with the Hydrogen Evolution Reaction (HER), the Oxygen Evolution Reaction (OER) is a

crucial half-reaction in the electrochemical splitting of water. The chemical equation:

2H₂O → O₂ + 4H⁺ + 4e⁻

26
describes OER, which takes place at the anode of an electrolysis cell and includes the oxidation

of water molecules to create oxygen gas and protons. The method of water electrolysis, which is

used to turn water into hydrogen fuel, is based on this reaction. OER is important for several

main reasons. First, it has a direct impact on the electrochemical water-splitting process's

efficiency and total energy usage. OER is usually the less efficient and more kinetically slow

half-reaction of the two, needing greater overpotentials to continue at desired rates. This

inefficiency reduces hydrogen's feasibility as a cost-competitive fuel source, increasing the

energy needed to split water, which is the main bottleneck in hydrogen generation.

Figure 14: Schemetic diagram of Electrochemical water splitting.


Second, the oxygen evolution reaction involves intricate mechanisms and severe oxidative

conditions, making it a chemically demanding reaction. It is necessary to employ strong, long-

lasting materials in these conditions so they can resist deterioration and corrosion over time.

Therefore, one of the main goals of the research is to increase the activity, stability, and

affordability of OER catalysts to improve the overall viability and sustainability of electrolytic

water splitting as a method of producing hydrogen [101-104].

27
1.4.2. Reactions involved in OER:

1.4.2.1. Voltammetry Cyclic (CV)

To determine the redox characteristics of an analyte in solution, a commonly used

electrochemical technique called cyclic voltage monitoring, or CV, entails sweeping the potential

of a working electrode linearly vs time in a cycle. Currents from oxidation and reduction

processes at the electrode surface are recorded when the voltage is swept, resulting in a

voltammogram with peaks corresponding to these processes. CV provides a wealth of

information on a substance's electrochemical activity, including redox potentials, reaction rates,

and potential reaction processes. The method may be used to characterize novel materials,

especially in the context of battery and supercapacitor research, as well as investigate the

stability of redox species and the impact of various solvents or electrolytes on redox behavior.

Cyclic voltammetry in the Non-Faradic region is used to find out the crucial electrochemical

surface area of an electrocatalyst [105].

1.4.2.2. Linear Sweep Voltammetry (LSV)

Similar to CV but usually without reversing the sweep's direction, Linear Sweep Voltammetry

(LSV) records the current flow while linearly varying the voltage of an electrode. LSV is

essential for examining a system's electrochemical reaction kinetics, especially for figuring out

the onset potentials of particular electrochemical processes. This method is frequently employed

to evaluate the performance of electrocatalysts, particularly in energy-related applications such

as electrolyzers and fuel cells. LSV is a vital tool in the development and assessment of catalysts

for hydrogen evolution and oxygen reduction processes because the onset potential sheds light

on how well catalysts initiate electrochemical reactions. One of the most important readings LSV

28
provides is the overpotential that determines the goodness of an electrocatalyst and whether the

reaction is kinetically and thermodynamically favorable.

An essential metric from the LSV data is Tafel analysis, which is a technique for understanding

the kinetics of electrochemical processes and especially useful for clarifying the rate-determining

step in electron transfer reactions, is the Tafel Slope. The Tafel plot, from which the Tafel slope

may be determined, is created by graphing the electrode potential against the logarithm of the

current density. This slope, which shows how the overpotential influences the reaction's pace,

provides important insights into the electrochemical reaction's process. Because it aids in the

selection and development of more effective catalysts as well as the design of better energy

conversion and storage devices, the Tafel Slope is essential for improving electrochemical

systems, such as batteries and fuel cells [101-104].

1.4.2.3. Electrochemical Impedance Spectroscopy (EIS)

The sensitive and potent method of EIS is employed to examine the intricate resistive and

capacitive characteristics of electrochemical systems. The system's impedance is found across a

variety of frequencies by measuring the current that results from applying a modest amplitude

AC voltage to an electrochemical cell. A Nyquist plot, which is commonly used to illustrate data,

may be useful in comprehending a number of phenomena, including diffusion properties, double

layer capacitance, and charge transfer resistances. When researching battery lifespans, corrosion

prevention, fuel cell efficiency, and coating integrity, EIS is helpful in understanding of the

electrochemical reactions occurring in the electrolytes and at the electrode interfaces [106, 107].

1.4.2.4. Chronopotentiometry

Chronopotentiometry is an electrochemical method in which an electrode receives a continuous

current and the potential change over time is measured. This technique is very helpful for

29
investigating the pace of electrochemical reactions and how stable an electrocatalyst is.

Transitions in the recorded potential in chronopotentiometry might signify the development of

new phases, such the deposition of a metal, or the exhaustion of reactants at the electrode

surface.

30
Chapter 2: Literature Review
Chemists have been fascinated with polyoxometalates (POMs), a special and adaptable family of

metal-oxygen cluster compounds, because of their amazing structural variety and wide range of

potential applications, from catalysis to medicine. It is still very difficult to create clusters that

are concentrated on a single metal species, even after thousands of POMs have been synthesized

to far. The intrinsic chemical characteristics and coordination preferences of various metals can

result in complicated and frequently unpredictable formation paths when attempting to

synthesize POMs, which is the main cause of this problem. The scientific literature frequently

features metals that are frequently employed in POMs, including as tungsten, molybdenum, and

vanadium, because of their favorable reaction conditions that readily result in stable, high-

nuclearity clusters. On the other hand, creating POMs based on uncommon metals may need

adjusting temperature, pH, and reaction conditions as well as selecting suitable organic or

inorganic ligands or counter ions, all of which play a part in the complex dance of molecule

assembly. Furthermore, the absence of specialized, customized synthesis techniques for less

common metal ions—which frequently lead to poorer yields and stability—can be blamed for the

scarcity of single-metal POMs. Therefore, although POM chemistry is a broad area, coordination

chemistry's intriguing but challenging frontier is revealed when one concentrates on clusters that

mostly include a single metal type. Differentially incorporated polyoxometalate (POM)

complexes have improved and shown frequently distinct features over conventional POMs,

rendering them highly important in cutting-edge scientific applications. Through the

incorporation of several metal ions into their compositions, these hybrid POMs are able to

provide a mutual benefit that capitalizes on the unique properties of each metal. For instance,

POM frameworks' catalytic activity may be greatly increased by adding transition metals like

31
copper, iron, or cobalt. These metals can increase the effectiveness of catalytic processes such as

water oxidation, reduction reactions, and synthesis transformations by adding additional active

sites or changing electrical characteristics. The included metals in these hybrid POMs bring up

new functional possibilities while retaining the basic benefits of conventional POMs, such as

high stability, toughness, and simplicity of synthesis.

A new POM was reported by Kortz et al. in 2006 in which a cobalt ion was lodged into a dimer

of two polyoxotungstate units joined by a W-O-W bridge. The same group reported the synthesis

of Co-based POM unit [Co(H2O)2(γ-SiW10O35)2]10−, in which Co metal ion was embedded into a

dimer formed by two (γ-SiW10O35) units [108]. M. T. Pope and co-workers prepared another type

of P.O.M. in which, along with a dimer of poly-tungstate unit, a third entity is involved in the

form of cobalt-ethylenediamine complex as in [K{Co(en)}W2O5(H2O)(PW9O34)2]12- and

[{Co(en)(O.H.)2Co(en)}-{Co(en)}2(PW10O37)2]12- [109]. In other reports, this dimer is also

replaced with a trimer of polyoxotungstates [110]. Co9(H2O)6(OH)3(HPO4)2(PW9O34)3]16– has

been synthesized by Soriano-López and colleagues and was used to alter amorphous carbon

paste electrodes. This polyoxometalate's catalytic activity is still potent in the solid state. At mild

overpotentials, the catalyst reaches adequate catalytic rates. With prolonged action, it exhibits

outstanding durability as a heterogeneous catalyst [111]. Numerous investigations have focused

on the tetracobalt Weakley sandwich, [CoII4(H2O)2(B-α-PW9O34)2]10−. To examine the electronic

parameters influencing their catalytic activities, K. Azmani et al. have compared its performance

to that of its iron equivalent, [FeIII4(H2O)2(B-α-PW9O34)2]6−. Electrocatalytic experiments

utilizing electrodes implanted with solid-state POMs and associated computational analysis

demonstrated that CoII-based POMs perform better than their FeIII counterparts for water

oxidation catalysis (WOC). Furthermore, nuclearity appears to have little effect on the activity of

32
the POMs, since weakly sandwich structures show somewhat superior WOC characteristics than

Keggin-type clusters [112]. The cobalt-inclusive Keggin-type polyoxometalates [CoW12O40]6−

and [Co(H2O)SiW11O39]6− were presented in a work published by A. Barros and coworkers. The

synthesis was performed with a simple and very efficient technique. Furthermore, prepared

catalysts were shown to have encouraging electrochemical properties using the one-pot

manufacturing method [113].

POMs have been used as electrode materials for various electrochemical and energy applications

[114]. POMs have the potential to be excellent electrode materials for challenging

electrochemical reactions due to their redox activity, mixed oxidation states, and maintaining

structural integrity during the electrochemical reactions [115-117]. These features are essential

for fast and reversible electron transfers and high stability as an electrode material.

Electrochemical water splitting for hydrogen production is an area of immense interest to the

scientific community. OER, one of the two half-cell reactions in waster splitting, is a more

challenging reaction due to sluggish electrode kinetics, and there is a significant focus on

developing electrocatalysts for the OER [118, 119]. In this regard, POMs can be the materials of

choice due to their large surface area, high density of active sites, and better stability [2, 5, 115,

120, 121]. However, only some reports exist on synthesizing POMs suitable for electrocatalytic

water splitting.

Hill et al. reported a Co-based POM, [Co4(H2O)2(PW9O34)2], in 2010, for the electrochemical

water splitting. However, the stability of this POM remained an issue, and with prolonged usage,

nanostructures of CoOx appeared at the electrode surface [122]. Later, in 2013, an inter-cluster

POM, [Co9(H2O)6(OH)3(HPO4)2(PW9O34)3]16-, was reported to exhibit good stability and

improved catalytic activity for the electrochemical water oxidation reaction [111]. Another

33
conceptually-related cobalt containing POM, Na30.5K1.5[{Co4(O.H.)3PO4}4(A-α-

GeW9O34)4].16H2O showed good electrocatalytic performance for the OER reaction of water

splitting [123, 124]. The mononuclear cobalt POM anion [Co(H2O)(H3PW11O39)]2− as reported

by Li et al. in 2017, showed modest electrocatalytic activity for water oxidation in neutral

phosphate buffer [125]. In contrast, the cobalt Keggin anion [CoW12O40]6− when confined in

ZIF-8 metal-organic framework (MOF), exhibited excellent stability for water oxidation at a

neutral pH [126]. Two Mo-based POMs [CoMo6O24H6]3− and [Co2Mo10O38H4]6− were tested for

the OER reaction. A comparison of the catalytic performance of these two POMs revealed that

the high nuclearity of cobalt was not necessary for better catalytic activity [127].

Apart from that, W. Liu et al. reported their work in which, by combining the appropriate

components, two monochromium(III) heteropolytungstate compounds, [CrIII(HPVW7O28)2]13-

and [CrIII(HAsVW7O28)2]13-, were created via simple, one-pot processes in a fundamental

aqueous environment. The hydrated sodium salts were the resultant compounds that were

isolated later. These compounds were then subjected to single crystal analysis and magnetic

studies [128]. R. Afrasiabi et al. reported the synthesis of TBA salt of a chromium-substituted

POM (Keggin-type) for the oxidation of sulfides into sulfones [129]. In 2020, Nadiia I.

Gumerova and coworkers reported the chromium-centered Keggin anion [α-CrW12O40]5–, a new

addition to the Keggin archetype. This molecule exhibited unusual magnetic characteristics due

to its peculiar tetrahedral coordination of CrIII, a configuration not previously described in

polyoxometalates [130].

In POM chemistry, there is comparatively less documentation available for polyoxometalate

(POM) complexes integrated with chromium than the other widely utilized metals. The reasons

for this underrepresentation include many difficulties related to the chemistry of chromium.

34
Because of chromium's ability to adopt several oxidation states and form different coordination

complexes, the hard and well-defined structures of POMs frequently exhibit unexpected

behavior. Furthermore, the synthesis of chromium-integrated POMs is more difficult and less

studied as it requires exact control over the experimental parameters. These elements hinder the

development of chromium-based POMs despite their promising nature for special chemical and

physical features, which is why this field hasn’t been explored much.

35
Chapter 3: Experimental Section
This chapter includes the step-by-step synthesis of polyoxometalate precursors and the synthesis

of polyoxometalate complexes using those precursors.

3.1. Chemicals and Solvents:

Solvents used in this experimental section were DI water, Acetic acid (CH3COOH), Acetonitrile

(MeCN), and Ethanol (C2H5OH). CH3COOH and MeCN were purchased from Merck and used

without further purification. All the synthesis and catalytic studies were conducted using DI

water unless stated otherwise.

Chemicals used were Sodium Tungstate Dihydrate (Na2WO4.2H2O), Phosphoric acid H3PO4

(85%), Glacial Acetic acid (CH3COOH), Cobalt Carbonate (CoCO3), Pivalic acid

((CH3)3CCO2H), and Chromium (III) nitrate nonahydrate (Cr(NO3)3.9H2O).

3.2. Apparatus and Glassware:

Weighing balance, Vacuum oven, Vacuum pump, Fume hood, Hot plates, Drying oven, pH

papers, pH meter, Filter paper, Clamps, Stands, Centrifuge machine, Beakers, Single and double

neck round bottom flasks, Condenser, Pipes, Rubber stoppers, Magnetic stirrer, Glass vials,

Reagent bottles, Spatula, Pipettes, Thermometer, Oil bath, Measuring cylinder, Eppendorf tubes,

Falcon tubes, Micropipettes, Petri dishes etc.

3.3. Procedure of Precursors

1. Synthesis of trilacunary Keggin POM precursor Na9[A-α-PW9O34].7H2O

A trilacunary Keggin POM precursor was prepared using the following method:

Using a 100 ml beaker, 12 g (0.036 mol) of sodium tungstate dihydrate Na2WO4.2H2O was

dissolved in 15 ml of water and the mixture was stirred using a magnetic stirrer bar until the solid

36
salt was completely dissolved. When a homogenized mixture was prepared, then 0.4 ml (0.006

mol) of 85% Phosphoric acid (H3PO4) was added dropwise with a continuous stirring. After the

addition of acid was completed, the pH of the mixture was checked using a pH meter. The

measured pH of the solution at that point was 8.9 – 9.0. Glacial acetic acid (2.25 ml, 0.04 mol)

was added dropwise with an ongoing vigorous stirring. A large quantity of white precipitates

formed during the addition of acetic acid. The final pH of the mixture was checked which was

7.5 ± 0.3. The solution was left on stirring for 1 h, and the precipitates were collected via vaccum

filtration and dried overnight in a drying oven at 60 ℃. The yield of this reaction was 9 g (75%).

The crude product Na9-xHx[A-PW9O34].xH2O was used in the subsequent reactions without

further purification.

37
Figura 15: Synthesis scheme for trilacunary POM precursor Na9[A-α-PW9O34].7H2O.
1. Synthesis of [Co2(μ-OH2)(O2CCMe3)4(HO2CCMe3)4]

2 g of CoCO3 which is equal to 34 mmol was mixed with an excess amount of pivalic acid (10 g)

in the presence of DI water and the mixture was subjected to a reflux process for 24 hours at a

38
temperature of 100 ℃ until the dissolution of cobalt salt occurred. After this, the solution was

cooled to room temperature, and 25 ml of acetonitrile was added with continuous stirring. The

solution was then cooled and placed in the fridge at a temperature of 5 ℃ to get a crop of pink

crystals. The yield of this reaction was 68%.

Figura 16: Synthesis scheme for [Co2(μ-OH2)(O2CCMe3)4(HO2CCMe3)4].


3. Synthesis of [(Cr3O(O2CC(CH3)3)6(H2O)3](O2CCMe3)3]

2.5 g (6.25 mmol) of Cr(NO3)3.9H2O was mixed with 2.5 g of pivalic acid in the presence of

water and subjected to heating via reflux method at 100 ℃ for 1 hour. The temperature was

subsequently raised to 150 ℃ and reaction was continued for further 15 - 20 minutes. After

allowing it to cool, the solution was filtered and the green residue was collected to redissolve in a

4:1 solution of Ethanol/DI water. This mixture was heated on a hot plate at 80 ℃ and filtered

while hot. The filtrate was placed for filtration at room temperature and green crystals were

39
collected after a few days. The yield of this reaction was 45%.

Figure 17: Synthetic diagram for [(Cr3O(O2CC(CH3)3)6(H2O)3](O2CCMe3)3].

3.4. Procedure of POM Complexes

1. Synthesis of Archimedean rhombicuboctahedron Rb2.5[(Co(H2O)6)0.5(PW12O40)].16H2O

0.1 g (0.04 mmol) of Na9[A-α-PW9O34].7H2O was dissolved in a 1:1 solution of Acetic acid/DI

water (5/ 5 ml). With continuous stirring, 0.05 g (0.052 mmol) of [Co2(μ-

OH2)(O2CCMe3)4(HO2CCMe3)4] was added into the mixture and heated at ~80 ℃ for 3 hours in

an oil bath then cooled the solution at room temperature. A few drops of 0.25 M RbCl were

added and placed for crystallization at room temperature. After one month, red crystals were

collected and dried in the air. The yield of this reaction was 18%. The prepared compound is

named Rb-1 for convenience.

40
Figure 18: Synthesis scheme for Rb2.5[(Co(H2O)6)0.5(PW12O40)].16H2O
2. Synthesis of [(Cr3O(O2CC(CH3)3)6(H2O)3]3[P2W18O62]. xH2O

0.04 g (0.035 mmol) of [(Cr3O(O2CC(CH3)3)6(H2O)3]3(O2CCMe3)3 was dissolved in a 1:1

mixture of Acetic Acid/Methanol (5/5 ml) and the mixture was left on stirring for 15 minutes.

Then 0.1 g (0.04 mmol) of Na9[A-α-PW9O34] was added to the solution. The mixture was then

heated at ~80 ℃ for 2 h in an oil bath. The white sediments in the green solution were formed

and the filtrate was separated by filtration. 1 M NaCl was added dropwise in the filtrate and the

solution was left undisturbed for crystallization at room temperature. After 10 days, green

crystals were obtained on filtration and dried in air. The yield of this reaction was 28%. The

prepared compound is named Na-1 for ease of use.

41
Figure 19: Synthetic scheme for [(Cr3O(O2CC(CH3)3)6(H2O)3]3[P2W18O62]. xH2O
3.5. Fabrication of Electrodes

The Rb-1-based working electrode was prepared by coating the material on a commercially

bought nickel foam (Ni-Foam) of a geometrical surface area of 1 cm-2 without using any

additional binder. Ni-Foam was thoroughly washed and cleaned by ultrasonication in 1 M HCl,

DI water, and analytical grade Ethanol for 20 minutes each, then dried in a vacuum oven at 60 ℃

for 2 hours. To make the slurry, 10 mg/mL of Rb-1 was mixed in water. The resulting blend was

ultrasonicated for 30 minutes to yield a homogeneous solution. This uniform solution was then

loaded on Ni-Foam via dip-coating. The process was repeated 4 times, and the prepared

electrodes were left in open air to dry. The average mass loading on electrodes was calculated to

be 3 mg. All the electrodes were then calcined in a chamber furnace at 450 ℃ in air for 2 hours.

Na-1 based working electrode was prepared by coating the material on a nickel foam of similar

42
area and dimensions without using any additional binder. To make the slurry, 10 mg/mL of Na-1

was mixed in water. The resulting mixture was subjected to ultrasonication for 30 minutes to

obtain a homogeneous solution. This solution was then loaded on Ni-Foam via drop-casting and

the process was repeated 4 times. Prepared electrodes were then left in the open air to dry. The

average mass loading for Na-1/Ni-Foam was 4 mg.

Figure 20: Washing of Ni-Foam and fabrication of electrodes.

43
Chapter 4: Results and Discussion
In this chapter, the characterization and analysis of synthesized polyoxometalate complexes have

been discussed. The techniques used were the following:

4.1. Single Crystal X-ray Diffraction (SCXRD)

The Single-crystal X-ray diffraction technique was used to study the three-dimensional

arrangement of atoms within the crystal lattice.

4.1.1. Single crystal XRD of Rb-1

A specimen of Co0.50H38O59PRb2.50W12 with the dimensions of 0.102 mm x 0.121 mm x 0.231

mm was studied with single crystal X-ray crystallography. The data on X-ray intensities were

collected. The cubic unit cell used to integrate the data and a total of 49819 reflections yielded a

maximum angle θ of 28.45° (0.75 Å resolution), of which 789 were independent (average

redundancy 63.142, completeness = 99.4%, Rint = 5.49%, Rsig = 1.06%) and 736 (93.28%) were

more significant than 2σ(F2). The cell constants were a = 22.5801(8) Å, b = 22.5801(8) Å, c =

22.5801(8) Å, volume = 11512.7(12) Å3. The Bruker SHELXTL Software Package was used to

solve and refine the structure using space group F m -3 m, with Z = 8 for the formula unit,

Co0.50H38O59PRb2.50W12. The goodness-of-fit was 1.258. The most significant hole was -2.040 e-

/Å3 with an R.M.S. deviation of 0.333 e-/Å3, and the most prominent peak in the final difference

electron density synthesis was 2.823 e-/Å3. Based on the final model, F(000), 12152 e- and

calculated density was 3.995 g/cm3.

44
(a) (b)

(c)

Figure 21: (a) Packing of Rb-1 (b) Cubic representation (c) Polyhedral representation of Co

(II) containing rhombicuboctahedron Archimedean solid coordination sphere. (Color coding,

W: blue, O: red, P: yellow, Co: green, Rb: purple, H: grey)

The single crystal XRD showed the unprecedented structure of Rb-1, which crystallized in cubic

crystal and had an F m-3m space group. The highly symmetrical Archimedean structure

rhombicuboctahedron contains centrally located Co (II) atoms. The structure showed 8 triangular

and 18 square faces, which shared vertices and corners. The oxygens (O5W—O5W) were

45
connected with bond lengths 1.29 Å, 1.73 Å and 0.85 Å. The crystal structure revealed that the

Cobalt atom was coordinated with 24 oxygen atoms. The six squares of oxygen atoms were

arranged around the cobalt atom with a (Co2—O5W) bond length of 2.08 Å. The water

molecules were also surrounding the coordination sphere of Cobalt. Each square of four oxygen

atoms further coordinated with six Rb atoms, forming a star-shaped structure.

46
Figure 22: Polyhedral / Ball and stick representation of different building blocks and different

directional views of Rb-1. (Color coding, W: blue, O: red, P: yellow, Co: green, Rb: purple, H:

grey)

The Rb atoms were connected through two oxygen atoms (O10W) with bond lengths of 0.9 Å.

The star shape was now at the center of the cube in which six Rb atoms connected with eight

Keggin polyanion {P2W12O40} through oxygen, which was at the cube's corners.

4.1.2. Single crystal XRD of Na-1

A specimen of empirical formula C72Cr18Na8O232P4W36 was studied with single-crystal X-ray

crystallography. The data on X-ray intensities were collected. The monoclinic unit cell was used

to integrate the data and a total of 237955 reflections were obtained. The cell constants of a =

17.6981 (3) Å, b = 29.3306 (4) Å, c = 29.4522 (4) Å, volume = 15281.5(4) Å3, were based upon

the refinement of the XYZ-centroids of 1045 parameters. The Bruker SHELXTL Software

Package solved and refined the structure using space group P 1 21/n 1, with Z = 1 for the formula

unit, C72Cr18Na8O232P4W36. The chemical formula weight was found to be 18829.600. The most

significant hole was -10.27 e-/Å3 with an R.M.S. deviation of 0.115 e-/Å3, and the most prominent

peak in the final difference electron density synthesis was 14.04 e-/Å3.

Single crystal XRD shows the structure of Na-1 which consists of Dawson-type POM anion

P2W18O62 joined with metal centers. Dawson anion makes a barrel-shaped structure consisting of

a bi-octahedron. There are two phosphorus atoms occupying the center of each octahedron. The

P-atom at the center forms a tetrahedron and is surrounded by WO6 octahedrons connected to the

P-atom via bridging O-atoms. Each Tungsten atom is joined with six Oxygen atoms via a bond

length of 2.426 Å. Terminal oxygens making a double bond with tungsten atoms have a shorter

47
bond length of 1.733 Å. Oxygen atoms joining W with P via W-O-P bonds have a W-O bond

length of 1.906 Å, meanwhile, the oxygen atoms connecting W-O-W atoms have a bond length

of 1.933 Å.

Figure 23: Crystal Structure of [(Cr3O(O2CC(CH3)3)6(H2O)3]3[P2W18O62]. xH2O and Dawson-


type POM anion.

48
Each P2W18O62 unit is bonded with three metal centers via bridging Oxygen atoms. The

chromium metal centers form a wheel structure because of the carbon atoms present in the

structure. Each metal centers contain three chromium atoms which are bridged together via an

oxygen atom present in the center of a wheel-like structure. The sodium atom acts like a counter-

ion in this highly negatively charged structure.

Figure 24: Wheel-shaped structure showing the metal centers.


The packing of Na-1 shows a monoclinic unit cell.

Figura 25: Monoclinic unit cell packing of Na-1.

49
4.2. FT-IR

Fourier Transform Infrared Spectroscopy (FTIR) is a highly versatile analytical technique used to

obtain the infrared spectrum of absorption or emission of a solid, liquid, or gas. High-resolution

spectral data over a broad spectral range are concurrently collected by an FTIR spectrometer.

This method converts the data into a format that is simple to evaluate and understand by applying

the concepts of the Fourier transform. In chemical and pharmaceutical research, it has grown to

be an essential technique for determining the structure of substances, examining sample

compositions, and comprehending molecular interactions.

The absorption and transmission of infrared light by molecules is the central idea of FTIR.

Certain infrared light wavelengths are absorbed by molecules, changing their dipole moment and

causing molecular vibrations at distinct frequencies. The molecular bonds can be bent, twisted,

and stretched during these vibrations. As a kind of molecular fingerprint, every kind of chemical

bond and molecule structure has an own collection of vibrational frequencies. Certain

frequencies of infrared light are absorbed by samples when they are transmitted through them,

which results in a reduction in the intensity of the light at certain frequencies. The outcome is a

spectrum that depicts the absorption and transmission of molecules, resulting in a unique

chemical fingerprint. This unique fingerprint is used to identify materials using their functional

groups.

4.2.1. FTIR Spectrum of Rb-1

The FTIR spectrum of Rb-1 shown below in the figure shows a wide and prominent peak

between 3200-3400 cm-1 depicting the presence of an O-H functional group in the material. This

OH shows the presence of water of crystallization in Rb-1 which is confirmed by SCXRD and

TGA as well [131]. A peak at 1076.28 cm-1 shows P-O vibration, 943 cm-1 shows W=O

50
stretching, and 890-860 cm-1 shows W-O-W stretching vibrations, confirming the structure of

POM anion. Meanwhile, Metal oxygen bonds (M-O) show peaks between a range of 400-700

cm-1 confirming the presence of Co-O bonds [113, 132].

Figure 26: FTIR Spectrum of Rb-1.


4.2.2. FTIR Spectrum of Na-1

A wide broad peak between 3200-3400 cm-1 represents the str. vibration of the OH functional

group present in the material due to the presence of water of crystallization. Other than that, a

peak at 775.26 cm-1 represents the W-O-W group (oxygen in the middle of the polyoxometalate

compound) and the peak at 904 cm-1 shows another W-O-W group (the oxygen at the edge of the

polyoxometalate compound). The peak at 955 cm-1 shows terminal oxygen present in the POM

anion which is double bonded with tungsten and a peak at 1090.27 cm-1 shows the P-O

51
vibrations. To confirm the metal-oxygen bond (Cr-O), 400 to 700 cm-1 is the range that shows the

presence of a Cr-O bond [133].

Figura 27: FTIR Spectrum of Na-1.


4.3. Thermogravimetric Analysis (TGA)

TGA is a thermal analysis technique that determines how a material's weight changes in response

to temperature or time in a controlled environment. Providing quantitative and qualitative

information on physical and chemical processes such as sublimation, adsorption, desorption,

absorption, and thermal degradation is its main goal. TGA is frequently used to evaluate

materials that show weight increase or loss as a result of oxidation, breakdown, or loss of

volatiles (such as moisture).

The fundamental idea behind TGA is not that complicated. It entails heating a sample over a

variety of temperatures and precisely measuring its mass throughout that process. The weight of

the sample may fluctuate as a result of different components disintegrating, oxidizing, or losing

volatiles as the temperature rises. Thus, crucial details on the material's composition, such as its

52
stability, the makeup of multi-component systems, and the estimation of lifetimes at various

temperatures, are provided by the TGA curve, which displays mass versus temperature. Materials

science, the polymer chemistry, medicines, food science, and any other discipline where

knowing the composition and stability of materials under heat stress is essential may all benefit

greatly from TGA.

The thermogram of Rb-1 was measured at the Ramp rate of 20 oC /min. The curve showed

weight loss up to 800 oC. A weight loss of 9.55 % was observed at 410 oC due to the loss of 17.4

chemically bonded water molecules. The TGA analysis showed 1.4 water molecules more as

compared to the single crystal XRD study which may be attributed to the physical adsorption

Figura 28: TGA analysis of Rb-1.

53
during sample handling. After the loss of water molecules, the further weight loss depicts the

ligand loss with the increase in temperature.

The synthesized POM complexes were also subjected to check their electrochemical water

splitting applications that consist of:

4.4. Oxygen Evolution Reaction (OER)

4.4.1. Electrochemical activity of Rb-1

All the electrochemical measurements were carried out in a standard three-electrode system

against Ag/AgCl (3 M KCl) used as the reference electrode, Pt wire as the counter electrode, and

Rb-1/Ni-Foam as the working electrode on a Potentiostat/Galvanostat using Gamry Interface

1000 electrochemical workstation (manufactured by Corrtest Instrument, Model CS350M). OER

activity was assessed in 50 mL of 1 M KOH (pH = 14), prepared using 18 MΩ DI water, at room

temperature (25 ℃). All the measured potentials reported in this thesis are converted from vs.

Ag/AgCl to vs. RHE (Reversible Hydrogen Electrode) using the Nernst Equation:

ERHE = EAg/AgCl + 0.059 × pH + E0Ag/AgCl

Where ERHE is the converted potential in vs. RHE, EAg/AgCl is the experimental value of potential

measured against Ag/AgCl reference electrode, 0.059 is a constant value and E0Ag/AgCl is the

standard potential of Ag/AgCl at 25 ℃, i.e., 0.1976 V [134].

The voltage window of 1.0 - 1.8 V vs. RHE was used to record cyclic voltammograms (CV) at

various scan rates (5, 10, 25, 50, 75, 100 mVs-1). The double-layer capacitance (Cdl) was

evaluated using this data. Linear Scan Voltammogram (LSV) was conducted from 1.0 V to 2.0 V

vs. RHE, at a scan rate of 10 mVs-1, to assess the electrochemical catalytic properties of the

OER. EIS was carried out in the frequency range of 100,000 Hz to 0.1 Hz while the given DC

54
voltage was 0.5 V vs. Ag/AgCl. A Nyquist plot was used to plot the data. By fitting the obtained

impedance data to an equivalent circuit, the charge transfer resistance (RCT) was determined.

Eight hours of stability testing were conducted via Chronopotentiometry.

CV measurements were conducted at different scan rates over the potential range of 1.0 to 1.8 V

(vs. RHE). CV curves of Rb-1/Ni-Foam recorded at different scan rates, exhibiting typical

features of faradic reactions occurring in an alkaline electrolyte [135]. By monitoring the

potential at which the oxidation and reduction processes take place in a sample, cyclic

voltammetry was utilized to estimate the redox potentials of the sample and investigate the

kinetics and mechanism of electrode processes. Peak current rises as the scan rate increases,

slightly varying both the cathodic and anodic peak potentials [135]. Comparable peak current

values of oxidation and reduction peaks indicate that the catalyst facilitates significant

reversibility. A high level of reversibility is often recommended in the context of OER since it

demonstrates that the catalyst can effectively speed up the OER process. A reversible catalyst

may move more quickly between its reduced and oxidized states, enabling participation in both

the forward and backward processes [136]. LSV was recorded using the scan rate 10 mV s-1. Rb-

1/Ni-Foam exhibits an on-set potential of 1.4 V (vs RHE). Overpotential (η), given by the

difference between electrode potential and the standard potential value for OER, was recorded to

be 207 mV at 10 mA.cm-2, 330 mV at 50 mA.cm-2 and 390 mV at 100 mA.cm-2, with maximum

current density reaching up to 595 mA cm-2 at a voltage of 2.0 V vs. RHE. In comparison, bare

Ni-Foam exhibited an overpotential of 430 mV for 10 mA.cm-2.

The equation η = b log (j) + a was used to fit the Tafel data, where η is the overpotential, j is the

current density, and b indicates the Tafel slope. Tafel slope was plotted using LSV data, which

gives a value of 51.76 mV dec-1, showing the kinetic parameters such as the effect of the rate of

55
increment of potential on the electrochemical reaction. The calculated Tafel slope value for bare

Ni-Foam electrode came out to be 172.35 mV dec-1. This shows that Rb-1 when used as the

electrocatalyst, promotes the evolution of oxygen by splitting water, 69.96% more than the bare

Ni-Foam.

Figura 29: (a) Cyclic Voltammogram for Rb-1/Ni-Foam at different scan rates ranging from 5

mV s-1 to 100 mV s-1. (b) Linear Sweep Voltammogram of Rb-1/Ni-Foam (black) and bare Ni-

Foam (red). (c) Tafel Slope for Rb-1/Ni-Foam (black) and bare Ni-Foam (red). (d) A comparison

56
of Rb-1/Ni-Foam (black) overpotential values and bare Ni-Foam (red) at 10 mA, 50 mA, and

100 mA, respectively.

Electron Impedance Spectroscopy (EIS) is a method used to investigate the electrical

characteristics of substances or systems and comprehend how electrochemical systems behave.

An electrochemical system's impedance as a function of frequency may be examined using EIS

to learn more about the system's electrical characteristics, including resistance, capacitance, and

reactance. It is also used to analyze a material's properties, including its surface properties,

charge transfer kinetics, and electronic and ionic conductivity [137]. EIS was carried out in the

frequency range of 100 kHz to 0.1 Hz; meanwhile, the given DC voltage was 0.5 V vs. Ag/AgCl.

A Nyquist plot of the imaginary part of impedance (Z") versus the real part of impedance (Z') is

shown in figure below. An essential step of fitting the suitable circuit is to gain insight into the

reaction kinetics and find charge transfer resistance (RCT) at the electrode-electrolyte interface,

which was 5.904 Ω. The smaller the RCT value, the faster the reaction kinetics [138].

Cyclic voltammetry in the non-Faradaic region can be used to study the electrochemical double-

layer capacitance and serve as a method for calculating the electrochemical active surface area

(ECSA), which aligns with earlier literature [139]. The characteristics of the electrode-electrolyte

interface and the electrolyte's properties are learned from the double-layer capacitance (Cdl)

[140]. Using CV at different scan rates, Cdl was examined using the difference of j values for

anodic and cathodic current according to equation (Δj = ((janodic - jcathodic)/2) and plotted it against

scan rates to give slope value, which was calculated to be 102.8 mF, indicating the core nature of

the electrolyte towards OER. Cdl was also used to find out ECSA using the equation ESCA =

Cdl/Cs where Cs is the specific capacitance of Ni-Foam. The value of ECSA for Rb-1/Ni-Foam

was 2,570 cm2.

57
Figura 30: (a) EIS for Rb-1/Ni-Foam with fitted circuit. (b) Cdl for Rb-1/Ni-Foam from linear

fitted data.

Durability is an important consideration that must be considered in addition to activity. The

stability of a catalyst is essential to knowing its working potential, cost-effectiveness, and

understanding of its possible commercial applications [141]. The catalyst exhibits excellent

stability in a robust alkaline electrolyte, as shown in the figure down below, which shows the

potential curve over 8 hours. The graph shows a very stable performance by Rb-1/Ni-Foam in

the CP scan, which shows no significant current change or degradation of the catalyst, indicating

a greater catalytic activity against aging. The working electrode was analyzed via pXRD before

and after undergoing a stability test to assess any possible changes that occurred. XRD pattern is

given in Fig. 31b, where comparative data is plotted. The graph shows no significant difference

in the XRD pattern of Rb-1/Ni-Foam before and after running a CP test, depicting its excellent

stability.

58
Figure 31: (a) Chronopotentiometry scan for Rb-1/Ni-Foam for 8 hrs. (b) XRD pattern, before &

after checking stability via Chronopotentiometry.

4.4.2. Electrochemical activity of Na-1

LSV was recorded using the scan rate of 5 mV s-1. Na-1/Ni-Foam exhibits an on-set potential of

1.46 V (vs RHE). Overpotential (η), given by the difference between electrode potential and the

standard potential value for OER, was recorded to be 226 mV at 10 mA.cm-2, 280 mV at 50

mA.cm-2 and 310 mV at 100 mA.cm-2, with maximum current density reaching up to 657 mA

cm-2 at a voltage of 1.7 V vs. RHE. In comparison, bare Ni-Foam exhibited an overpotential of

430 mV for 10 mA.cm-2.

The equation η = b log (j) + a was used to fit the Tafel data, where η is the overpotential, j is the

current density, and b indicates the Tafel slope. Tafel slope was plotted using LSV data, which

gives a value of 54.6 mV dec-1, showing the kinetic parameters such as the effect of the rate of

increment of potential on the electrochemical reaction. The calculated Tafel slope value for bare

Ni-Foam electrode came out to be 171 mV dec-1. This shows that Na-1 when used as the

electrocatalyst, promotes the evolution of oxygen by splitting water, 68.07% more than the bare

Ni-Foam.

59
EIS was carried out in the frequency range of 100 kHz to 0.1 Hz; meanwhile, the given DC

voltage was 0 V vs. Ag/AgCl. A Nyquist plot of the imaginary part of impedance (Z") versus the

real part of impedance (Z') is shown in figure below. An essential step of fitting the suitable

circuit is to gain insight into the reaction kinetics and find charge transfer resistance (RCT) at the

electrode-electrolyte interface, which was 12.17 Ω.

Figure 32: (a) Linear Sweep Voltammogram of Na-1/Ni-Foam (red) and bare Ni-Foam (black).

(b) A comparison of Na-1/Ni-Foam (black) overpotential values and bare Ni-Foam (red) at 10

mA, 50 mA, and 100 mA, respectively. (c) Tafel Slope for Na-1/Ni-Foam (red) and bare Ni-

Foam (black). (d) EIS for Na-1/Ni-Foam with fitted circuit.

60
The catalyst exhibits excellent stability in a robust alkaline electrolyte, as shown in the

Chronopotentiometry test in Fig. 34, which shows the potential curve over 8 hours. The graph

shows a very stable performance by Na-1/Ni-Foam in the CP scan, which shows no significant

current change or degradation of the catalyst, indicating a greater catalytic activity against aging.

Figure 33: Chronopotentiometry scan for Na-1/Ni-Foam for 8 hrs.

61
Chapter 5: Conclusion
In summary, two POM complexes, Rb-1 and Na-1 were synthesized using the trilacunary Keggin

POM precursor Na9[A-α-PW9O34].7H2O and [Co2(μ-OH2)(O2CCMe3)4(HO2CCMe3)4], and

Na9[A-α-PW9O34].7H2O and [(Cr3O(O2CC(CH3)3)6(H2O)3]3[P2W18O62]. xH2O in a single direct pot

reaction, respectively. The unique structure of Rb-1, which is classified as belonging to the Fm-

3m space group and has a cubic crystal lattice, was discovered via single crystal X-ray

diffraction (XRD) study. The crystallographic analysis showed that a cobalt atom and 24 oxygen

atoms were coordinated. Oxygen atoms and water molecules surrounded cobalt atoms to form

star-shaped structures, which happened to be at the cube's center, where six Rb atoms were

connected with eight Keggin polyanions PW12O40. The electrochemical performance of Rb-1

was assessed using various techniques such as Cyclic Voltammetry (CV), Linear Sweep

Voltammetry (LSV), Electron Impedance Spectroscopy (EIS), and Chronopotentiometry (CP).

Rb-1 showed a remarkable performance towards OER, exhibiting an overpotential of 207 mV @

10 mA.cm-2 with a current density of 595 mA.cm-2. The Tafel value for Rb-1 was 51.76 mV dec-1

compared to 172.35 mV dec-1 for bare Ni Foam electrode, showing good reaction kinetics. EIS

further supported the good performance of Rb-1, as it showed a low charge transfer resistance

(RCT) of 5.904 ohms. The electrochemical surface area (ECSA) of Rb-1/Ni-Foam was calculated

to be 2,570 cm2. Meanwhile, the single crystal XRD study for Na-1 showed a primitive type

structure with the monoclinic unit cell and P 1 21/n 1 space group. The study showed that a

Dawson-type POM anion P2W18O62 is connected with three metal centers each containing one

sodium and three chromium atoms. The metal centers make a wheel-shaped structure with

alternate Cr and C atoms. The electrochemical performance of Na-1 was observed via typical

techniques such as CV, LSV, EIS, and CP. It showed a good OER performance with an

62
overpotential of 226 mV @ 10 mA.cm-2 with a current density of 657 mA.cm-2. The Tafel slope

value for Na-1 was 54.6 mV dec-1 and RCT value was 12.17 Ω. Both catalysts, Rb-1 and Na-1

showed high stability in chronopotentiometry, and their good performance makes them a solid

and viable choice for water-splitting applications. This work explores the possibility of using

new POM-based materials as sustainable catalysts for electrocatalytic water-splitting reactions.

63
Chapter 6: References

1. Berzelius, J.J.P.A., The preparation of the phosphomolybdate ion [PMo12O40] 3−. 1826. 6: p.
369-371.
2. Wang, S.-S. and G.-Y. Yang, Recent Advances in Polyoxometalate-Catalyzed Reactions. Chemical
Reviews, 2015. 115(11): p. 4893-4962.
3. Song, Y.-F. and R. Tsunashima, Recent advances on polyoxometalate-based molecular and
composite materials. Chemical Society Reviews, 2012. 41(22): p. 7384-7402.
4. Li, Q., et al., Polyoxometalate-based materials for advanced electrochemical energy conversion
and storage. Chemical Engineering Journal, 2018. 351: p. 441-461.
5. Wang, D., et al., Polyoxometalate-based composite materials in electrochemistry: state-of-the-art
progress and future outlook. Nanoscale, 2020. 12(10): p. 5705-5718.
6. Bijelic, A., M. Aureliano, and A.J.C.C. Rompel, The antibacterial activity of polyoxometalates:
structures, antibiotic effects and future perspectives. 2018. 54(10): p. 1153-1169.
7. Rhule, J.T., et al., Polyoxometalates in medicine. 1998. 98(1): p. 327-358.
8. Pope, M.T., Y. Jeannin, and M. Fournier, Heteropoly and isopoly oxometalates. Vol. 8. 1983:
Springer.
9. Long, D.-L., E. Burkholder, and L.J.C.S.R. Cronin, Polyoxometalate clusters, nanostructures and
materials: From self assembly to designer materials and devices. 2007. 36(1): p. 105-121.
10. Reedijk, J. and K.R. Poeppelmeier, Comprehensive inorganic chemistry II: from elements to
applications. 2013.
11. Pope, M.T. and A. Müller, Polyoxometalate Chemistry From Topology via Self-Assembly to
Applications [electronic resource].
12. Müller, A., et al., Inorganic Chemistry Goes Protein Size: A Mo368 Nano-Hedgehog Initiating
Nanochemistry by Symmetry Breaking. Angewandte Chemie International Edition, 2002. 41(7): p.
1162-1167.
13. Proust, A., et al., Functionalization and post-functionalization: a step towards polyoxometalate-
based materials. Chemical Society Reviews, 2012. 41(22): p. 7605-7622.
14. Wang, H., et al., In Operando X-ray Absorption Fine Structure Studies of Polyoxometalate
Molecular Cluster Batteries: Polyoxometalates as Electron Sponges. Journal of the American
Chemical Society, 2012. 134(10): p. 4918-4924.
15. Hill, C.L., Progress and challenges in polyoxometalate-based catalysis and catalytic materials
chemistry. Journal of Molecular Catalysis A: Chemical, 2007. 262(1): p. 2-6.
16. Proust, A., R. Thouvenot, and P. Gouzerh, Functionalization of polyoxometalates: towards
advanced applications in catalysis and materials science. Chemical Communications, 2008(16): p.
1837-1852.
17. Busche, C., et al., Design and fabrication of memory devices based on nanoscale polyoxometalate
clusters. Nature, 2014. 515(7528): p. 545-549.
18. Du, D.-Y., et al., Recent advances in porous polyoxometalate-based metal–organic framework
materials. Chemical Society Reviews, 2014. 43(13): p. 4615-4632.
19. Fan, D., J. Hao, and Q. Wei, Assembly of Polyoxometalate-Based Composite Materials. Journal of
Inorganic and Organometallic Polymers and Materials, 2012. 22(2): p. 301-306.
20. Hutin, M., D.-L. Long, and L. Cronin, Controlling the Molecular Assembly of Polyoxometalates
from the Nano to the Micron Scale: Molecules to Materials. Israel Journal of Chemistry, 2011.
51(2): p. 205-214.

64
21. Clemente-Juan, J.M., E. Coronado, and A. Gaita-Ariño, Magnetic polyoxometalates: from
molecular magnetism to molecular spintronics and quantum computing. Chemical Society
Reviews, 2012. 41(22): p. 7464-7478.
22. Müller, A., P. Kögerler, and A.W.M. Dress, Giant metal-oxide-based spheres and their topology:
from pentagonal building blocks to keplerates and unusual spin systems. Coordination Chemistry
Reviews, 2001. 222(1): p. 193-218.
23. Müller, C.E., et al., Polyoxometalates—a new class of potent ecto-nucleoside triphosphate
diphosphohydrolase (NTPDase) inhibitors. Bioorganic & Medicinal Chemistry Letters, 2006.
16(23): p. 5943-5947.
24. Scheele, C.W., Sämmtliche physische und chemische Werke. Vol. 1. 1793: Mayer & Müller.
25. de Marignac, J.-C.G., Recherches sur les acides silicotungstiques et note sur la constitution de
l'acide tungstique. 1864: Gauthier-Villars.
26. Salles, R., Electroactive polyoxometalate-based molecular layers for nano-electronics. 2021.
27. Werner, A., Beitrag zur Konstitution anorganischer Verbindungen. Zeitschrift für anorganische
Chemie, 1893. 3(1): p. 267-330.
28. Miolati, A. and R.J.J.f.P.C. Pizzighelli, Zur Kenntnis der komplexen Säuren I. 1. Über die
Leitfähigkeit von molybdänsäurehaltigen Gemischen. 1908. 77(1): p. 417-456.
29. Rosenheim, A. and J.J.Z.f.a.u.a.C. Jaenicke, Zur Kenntnis der Iso‐und Heteropolysäuren. XV.
Mitteilung Über Heteropolywolframate und einige Heteropolymolybdänate. 1917. 101(1): p. 235-
275.
30. Pauling, L.J.J.o.t.A.C.S., The molecular structure of the tungstosilicates and related compounds.
1929. 51(10): p. 2868-2880.
31. Bragg, W.L. and W.H. Bragg, The structure of some crystals as indicated by their diffraction of X-
rays. Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical
and Physical Character, 1997. 89(610): p. 248-277.
32. Bragg, W.H., W.L.J.P.o.t.R.S.o.L.S.A. Bragg, Containing Papers of a Mathematical, and P. Character,
The reflection of X-rays by crystals. 1913. 88(605): p. 428-438.
33. Keggin, J.F. and W.L. Bragg, The structure and formula of 12-phosphotungstic acid. Proceedings
of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical
Character, 1997. 144(851): p. 75-100.
34. Keggin, J.F., Structure of the Crystals of 12-Phosphotungstic Acid. Nature, 1933. 132(3331): p.
351-351.
35. Dawson, B.J.A.C., The structure of the 9 (18)-heteropoly anion in potassium 9 (18)-
tungstophosphate, K6 (P2W18O62). 14H2O. 1953. 6(2): p. 113-126.
36. Evans Jr, H.T.J.J.o.t.A.c.s., The crystal structures of ammonium and potassium molybdotellurates.
1948. 70(3): p. 1291-1292.
37. Anderson, J.S., Constitution of the Poly-acids. Nature, 1937. 140(3550): p. 850-850.
38. Pope, M.T. and A. Müller, Polyoxometalate Chemistry: An Old Field with New Dimensions in
Several Disciplines. Angewandte Chemie International Edition in English, 1991. 30(1): p. 34-48.
39. Baker, L.C.W. and D.C. Glick, Present General Status of Understanding of Heteropoly Electrolytes
and a Tracing of Some Major Highlights in the History of Their Elucidation. Chemical reviews,
1998. 98(1): p. 3-50.
40. Dolbecq, A., et al., Hybrid Organic–Inorganic 1D and 2D Frameworks with ε-Keggin
Polyoxomolybdates as Building Blocks. Chemistry – A European Journal, 2003. 9(12): p. 2914-
2920.
41. Šimuneková, M., et al., Tetradecanuclear lanthanide-vanadium “nanochocolates”: catalytically-
active cationic heteropolyoxovanadium clusters. RSC Advances, 2013. 3(18): p. 6299-6304.

65
42. Abbas, H., et al., Controllable Growth of Chains and Grids from Polyoxomolybdate Building Blocks
Linked by Silver(I) Dimers. Chemistry – A European Journal, 2005. 11(4): p. 1071-1078.
43. Rosnes, M.H., et al., Mapping the synthesis of low nuclearity polyoxometalates from
octamolybdates to Mn-Anderson clusters. Dalton Transactions, 2012. 41(33): p. 10071-10079.
44. Mohs, T.R., et al., An organoimido derivative of the hexatungstate cluster: Preparation and
structure of [W6O18 (NAr)] 2-(Ar= 2, 6-(i-Pr) 2C6H3). 1995. 34(1): p. 9-10.
45. Errington, R.J., M.D. Kerlogue, and D.G.J.J.o.t.C.S. Richards, Chemical Communications, Non-
aqueous routes to a new polyoxotungstate. 1993(7): p. 649-651.
46. Kurata, T., et al., Cyclic Polyvanadates Incorporating Template Transition Metal Cationic Species:
Synthesis and Structures of Hexavanadate [PdV6O18]4-, Octavanadate [Cu2V8O24]4-, and
Decavanadate [Ni4V10O30(OH)2(H2O)6]4. Inorganic Chemistry, 2005. 44(7): p. 2524-2530.
47. Lin, S., et al., Preparation of polyoxometalates in ionic liquids by ionothermal synthesis. Dalton
Transactions, 2010. 39(7): p. 1740-1744.
48. Fu, H., et al., An Ionothermal Synthetic Approach to Porous Polyoxometalate‐Based Metal–
Organic Frameworks. 2012. 32(124): p. 8109-8113.
49. Ahmed, E. and M. Ruck, Ionothermal synthesis of polyoxometalates. Angewandte Chemie
(International ed. in English), 2012. 51(2): p. 308-309.
50. Wutkowski, A., et al., A New Mixed-Valent High Nuclearity Polyoxovanadate Cluster Based on the
{V18O42} Archetype. Zeitschrift für anorganische und allgemeine Chemie, 2011. 637(14-15): p.
2198-2204.
51. Antonova, E., C. Näther, and W. Bensch, Antimonato polyoxovanadates with structure directing
transition metal complexes: pseudopolymorphic {Ni(dien)2}3[V15Sb6O42(H2O)]·nH2O
compounds and {Ni(dien)2}4[V16Sb4O42(H2O)]. Dalton Transactions, 2012. 41(4): p. 1338-1344.
52. Zhao, J., et al., Synthesis, structure and magnetic property of the first 2-D organic–inorganic
hybrid sandwich-type polyoxotungstate containing O-donor organic ligands. Inorganic Chemistry
Communications, 2009. 12(6): p. 450-453.
53. Lisnard, L., et al., Hydrothermal syntheses and characterizations of 0D to 3D polyoxotungstates
linked by copper ions. Inorganica Chimica Acta, 2004. 357(3): p. 845-852.
54. Fraqueza, G., et al., Decavanadate, decaniobate, tungstate and molybdate interactions with
sarcoplasmic reticulum Ca2+-ATPase: quercetin prevents cysteine oxidation by vanadate but does
not reverse ATPase inhibition. Dalton Transactions, 2012. 41(41): p. 12749-12758.
55. Michailovski, A. and G.R. Patzke, Hydrothermal Synthesis of Molybdenum Oxide Based Materials:
Strategy and Structural Chemistry. Chemistry – A European Journal, 2006. 12(36): p. 9122-9134.
56. de la Oliva, A.R., et al., Assembly of a gigantic polyoxometalate cluster {W200Co8O660} in a
networked reactor system. 2012. 124(51): p. 12931-12934.
57. Richmond, C.J., et al., A flow-system array for the discovery and scale up of inorganic clusters.
Nature Chemistry, 2012. 4(12): p. 1037-1043.
58. Symes, M.D., et al., Integrated 3D-printed reactionware for chemical synthesis and analysis.
Nature Chemistry, 2012. 4(5): p. 349-354.
59. Kitson, P.J., et al., Combining 3D printing and liquid handling to produce user-friendly
reactionware for chemical synthesis and purification. Chemical Science, 2013. 4(8): p. 3099-
3103.
60. Pope, M.T.J.I.C., Heteropoly and isopoly anions as oxo complexes and their reducibility to mixed-
valence blues. 1972. 11(8): p. 1973-1974.
61. Waters, T., R.A.J. O'Hair, and A.G. Wedd, Catalytic Gas Phase Oxidation of Methanol to
Formaldehyde. Journal of the American Chemical Society, 2003. 125(11): p. 3384-3396.

66
62. J. Deery, M., O. W. Howarth, and K. R. Jennings, Application of electrospray ionisation mass
spectrometry to the study of dilute aqueous oligomeric anions and their reactions. Journal of the
Chemical Society, Dalton Transactions, 1997(24): p. 4783-4788.
63. Vilā-Nadal, L., et al., Nucleation Mechanisms of Molecular Oxides: A Study of the Assembly–
Dissassembly of [W6O19] 2Ā by Theory and Mass Spectrometry. 2009. 48: p. 5452-5456.
64. Christie, L.G., et al., Investigating the Formation of Giant {Pd72}Prop and {Pd84}Gly Macrocycles
Using NMR, HPLC, and Mass Spectrometry. Journal of the American Chemical Society, 2018.
140(30): p. 9379-9382.
65. Scullion, R.A., et al., Exploring the Symmetry, Structure, and Self-Assembly Mechanism of a
Gigantic Seven-Fold Symmetric {Pd84} Wheel. Angewandte Chemie International Edition, 2014.
53(38): p. 10032-10037.
66. Henson, A.B., P.S. Gromski, and L. Cronin, Designing Algorithms To Aid Discovery by Chemical
Robots. ACS Central Science, 2018. 4(7): p. 793-804.
67. Gromski, P.S., et al., How to explore chemical space using algorithms and automation. Nature
Reviews Chemistry, 2019. 3(2): p. 119-128.
68. Gabb, D., et al., Organic-soluble lacunary {M2(P2W15)2} polyoxometalate sandwiches showing a
previously unseen αββα isomerism. Dalton Transactions, 2012. 41(33): p. 10000-10005.
69. Li, F., et al., Cation induced structural transformation and mass spectrometric observation of the
missing dodecavanadomanganate(iv). Dalton Transactions, 2012. 41(33): p. 9859-9862.
70. Pradeep, C.P., D.-L. Long, and L. Cronin, Cations in control: crystal engineering polyoxometalate
clusters using cation directed self-assembly. Dalton Transactions, 2010. 39(40): p. 9443-9457.
71. Kirby, J.F. and L.C.W. Baker, Effects of Counterions in Heteropoly Electrolyte Chemistry. 1.
Evaluations of Relative Interactions by NMR on Kozik Salts. Inorganic Chemistry, 1998. 37(21): p.
5537-5543.
72. Knoth, W. and R.J.J.o.t.A.C.S. Harlow, New tungstophosphates: Cs6W5P2O23, Cs7W10PO36, and
Cs7Na2W10PO37. 1981. 103(7): p. 1865-1867.
73. Lindqvist, I.J.A.F.K., The structure of the tetramolybdate ion. 1950. 2(4): p. 349-355.
74. Nyman, M., et al., Solid-state Structures and Solution Behavior of Alkali Salts of
the$$[\hbox{Nb}_{6}\hbox{O}_{19}]^{8-}$$Lindqvist Ion. Journal of Cluster Science, 2006. 17(2):
p. 197-219.
75. Lindqvist, I.J.A.K., The structure of the hexaniobate ion in 7Na2O· 6Nb2O5· 32H2O. 1953. 5: p.
247-250.
76. Lindqvist, X. and B.J.A.K. Aronsson, The structure of the hexatantalate ion K 8 [Ta 6 O 19]· 16H 2
O. 1954. 7: p. 49-52.
77. Garner, C.D., et al., Studies in eight-co-ordination. Part 5. Crystal and molecular structure and
electron spin resonance spectra of tetrakis(diethyldithio-carbamato)molybdenum(V)
hexamolybdate and chloride. Journal of the Chemical Society, Dalton Transactions, 1978(11): p.
1582-1589.
78. Asami, M., H. Ichida, and Y.J.A.C.S.C.C.S.C. Sasaki, The structure of hexakis
(tetramethylammonium) dihydrogendodecatungstate enneahydrate,[(CH3) 4N] 6 [H2W12O40].
9H2O. 1984. 40(1): p. 35-37.
79. Daniel, C. and H. Hartl, Neutral and Cationic VIV/VV Mixed-Valence Alkoxo-polyoxovanadium
Clusters [V6O7(OR)12]n+ (R = −CH3, −C2H5): Structural, Cyclovoltammetric and IR-Spectroscopic
Investigations on Mixed Valency in a Hexanuclear Core. Journal of the American Chemical
Society, 2005. 127(40): p. 13978-13987.
80. Müller, A., et al., Cis-/Trans-Isomerie bei Bis-(trisalkoxy)-hexavanadaten: cis-
Na2[VO7(OH)6{(OCH2)3CCH2OH}2] · 8 H2O, cis-(CN3H6)3[VIVVO13{(OCH2)3CCH2OH}2] · 4,5

67
H2O und trans-(CN3H6)2[VO13{(OCH2)3CCH2OH}2] · H2O. Zeitschrift für anorganische und
allgemeine Chemie, 1995. 621(11): p. 1818-1831.
81. Strandberg, R., et al., Multicomponent Polyanions. IV. The Molecular and Crystal Structure of
Na6Mo5P2O23 (H2O) 13, a Compound Containing Sodium-coordinated
Pentamolybdodiphosphate Anions. 1973. 27: p. 1004-1018.
82. Ji, Y.-M., et al., Copper(II) and cadmium(II) complexes derived from Strandberg-type
polyoxometalate clusters: Synthesis, crystal structures, spectroscopy and biological activities.
Inorganic Chemistry Communications, 2017. 86: p. 22-25.
83. Li, Z.-L., et al., Three molybdophosphates based on Strandberg-type anions and Zn(ii)-
H2biim/H2O subunits: syntheses, structures and catalytic properties. Dalton Transactions, 2014.
43(15): p. 5840-5846.
84. Hu, G., et al., Hydrothermal synthesis, structure and properties of two new phosphomolybdates
based on Strandberg-type {P2Mo5O23}6− building units. Inorganic Chemistry Communications,
2015. 60: p. 33-36.
85. Tézé, A., G.J.J.o.I. Hervé, and N. Chemistry, Relationship between structures and properties of
undecatungstosilicate isomers and of some derived compounds. 1977. 39(12): p. 2151-2154.
86. López, X., et al., Relative Stability in α- and β-Wells−Dawson Heteropolyanions: A DFT Study of
[P2M18O62]n- (M = W and Mo) and [P2W15V3O62]n. Inorganic Chemistry, 2003. 42(8): p. 2634-
2638.
87. McDowell, J.J.A.C.S.B.S.C. and C. Chemistry, The crystal and molecular structure of
phenothiazine. 1976. 32(1): p. 5-10.
88. Mariotti, A.W.A., et al., Synthesis and Voltammetry of [bmim]4[α-S2W18O62] and Related
Compounds: Rapid Precipitation and Dissolution of Reduced Surface Films. Inorganic Chemistry,
2007. 46(7): p. 2530-2540.
89. Contant, R. and R. Thouvenot, A reinvestigation of isomerism in the Dawson structure: syntheses
and 183W NMR structural characterization of three new polyoxotungstates [X2W18O62]6−
(X=PV, AsV). Inorganica Chimica Acta, 1993. 212(1): p. 41-50.
90. Ding, Y., et al., Synthesis, structures and electrochemical properties of four organic–inorganic
hybrid polyoxometalates constructed from polyoxotungstate clusters and transition metal
complexes. Transition Metal Chemistry, 2009. 34(3): p. 281-288.
91. Minato, T., et al., Synthesis of α-Dawson-Type Silicotungstate [α-Si2W18O62]8− and Protonation
and Deprotonation Inside the Aperture through Intramolecular Hydrogen Bonds. Chemistry – A
European Journal, 2014. 20(20): p. 5946-5952.
92. Perloff, A.J.I.C., Crystal structure of sodium hexamolybdochromate (III) octahydrate, Na3
(CrMo6O24H6). 8H2O. 1970. 9(10): p. 2228-2239.
93. Evans, H.T.J.A.C.S.B.S.C. and C. Chemistry, The molecular structure of the hexamolybdotellurate
ion in the crystal complex with telluric acid (NH4) 6 [TeMo6O24]. Te (OH) 6.7 H2O. 1974. 30(9): p.
2095-2100.
94. Evans Jr, H.T.J.J.o.t.A.C.S., Refined molecular structure of the heptamolybdate and
hexamolybdotellurate ions. 1968. 90(12): p. 3275-3276.
95. Hahn, T., U. Shmueli, and J.W. Arthur, International tables for crystallography. Vol. 1. 1983: Reidel
Dordrecht.
96. Klug, H.P. and L.E. Alexander, X-ray diffraction procedures: for polycrystalline and amorphous
materials. 1974.
97. Cullity, B.J.b.M.C., Addison-Weslay, Boston, Elements of X-ray. 1978: p. 447-478.
98. Creagh, D. and W.J.J.V.C. McAuley, , Kluwer Academic Publishers, Boston, Table, International
tables for crystallography. 1992. 4(4.3): p. 200-206.

68
99. Muller, P., et al., Crystal Structure Refinement: A Crystallographer's Guide to SHELXL. Vol. 8. 2006:
OUP Oxford.
100. Clegg, W., et al., Crystal Structure Analysis. 2002: Oxford University Press.
101. Peng, X., et al., Strategies to improve cobalt-based electrocatalysts for electrochemical water
splitting. Journal of Catalysis, 2021. 398: p. 54-66.
102. Sun, H., et al., Electrochemical Water Splitting: Bridging the Gaps Between Fundamental
Research and Industrial Applications. ENERGY & ENVIRONMENTAL MATERIALS, 2023. 6(5): p.
e12441.
103. Wei, C. and Z.J. Xu, The Comprehensive Understanding of as an Evaluation Parameter for
Electrochemical Water Splitting. Small Methods, 2018. 2(11): p. 1800168.
104. Li, X., et al., Nanostructured catalysts for electrochemical water splitting: current state and
prospects. Journal of Materials Chemistry A, 2016. 4(31): p. 11973-12000.
105. Marken, F., A. Neudeck, and A.M. Bond, Cyclic Voltammetry, in Electroanalytical Methods: Guide
to Experiments and Applications, F. Scholz, et al., Editors. 2010, Springer Berlin Heidelberg:
Berlin, Heidelberg. p. 57-106.
106. Wang, S., et al., Electrochemical impedance spectroscopy. Nature Reviews Methods Primers,
2021. 1(1): p. 41.
107. Chang, B.-Y. and S.-M. Park, Electrochemical Impedance Spectroscopy. 2010. 3(Volume 3, 2010):
p. 207-229.
108. Bassil, B.S., et al., Transition metal containing decatungstosilicate dimer [M(H2O)2(γ-
SiW10O35)2]10− (M = Mn2+, Co2+, Ni2+). Dalton Trans., 2006(35): p. 4253-4259.
109. Belai, N. and M.T. Pope, Chelated heteroatoms in polyoxometalates and the topological
equivalence of {Coiii(en)} to type II cis-dioxometal centers. Synthesis and structure of [{Co(en)(μ-
OH)2Co(en)}{PW10O37Co(en)}2]8− and [K⊂{Co(en)WO4}{WO(H2O)}(PW9O34)2]12−. Chemical
Communications, 2005(46): p. 5760-5762.
110. Zheng, S.-T. and G.-Y. Yang, Recent advances in paramagnetic-TM-substituted polyoxometalates
(TM = Mn, Fe, Co, Ni, Cu). Chemical Society Reviews, 2012. 41(22): p. 7623-7646.
111. Soriano-López, J., et al., Cobalt Polyoxometalates as Heterogeneous Water Oxidation Catalysts.
Inorganic Chemistry, 2013. 52(9): p. 4753-4755.
112. Azmani, K., et al., Understanding polyoxometalates as water oxidation catalysts through iron vs.
cobalt reactivity. Chemical Science, 2021. 12(25): p. 8755-8766.
113. Barros, Á., et al. Systematic Approach to the Synthesis of Cobalt-Containing Polyoxometalates for
Their Application as Energy Storage Materials. Materials, 2023. 16, DOI: 10.3390/ma16145054.
114. Horn, M.R., et al., Polyoxometalates (POMs): from electroactive clusters to energy materials.
Energy & Environmental Science, 2021. 14(4): p. 1652-1700.
115. Zhao, J.-W., et al., Research progress on polyoxometalate-based transition-metal–rare-earth
heterometallic derived materials: synthetic strategies, structural overview and functional
applications. Chemical Communications, 2016. 52(24): p. 4418-4445.
116. Ahmed, K.M. and K. Amani, A novel amine-functionalized polyoxometalate-based metal-organic
framework: A reusable heterogeneous nanocomposite for selective oxidation of alcohols. Journal
of Molecular Structure, 2024. 1303: p. 137503.
117. Cui, H., et al., Assembly of some composites of Keggin-type polyoxometalate bearing Zn/Cu-
organic complexes as high capacity electrode materials. Journal of Molecular Structure, 2024.
1297: p. 136935.
118. Hassan, A., et al., Ferrocene-Boosted Nickel Sulfide Nanoarchitecture for Enhanced Alkaline
Water Splitting. Chem. Asian J. n/a(n/a): p. e202301051.
119. Zafar, H.K., et al., S-doped copper selenide thin films synthesized by chemical bath deposition for
photoelectrochemical water splitting. Applied Surface Science, 2023. 641: p. 158505.

69
120. Razzaque, S., et al., Selective Synthesis of Bismuth or Bismuth Selenide Nanosheets from a Metal
Organic Precursor: Investigation of their Catalytic Performance for Water Splitting. Inorganic
Chemistry, 2021. 60(3): p. 1449-1461.
121. Tahir, N., et al., Engineering Mn-Doped CdS Thin Films Through Chemical Bath Deposition for
High-Performance Photoelectrochemical Water Splitting. Chem. Asian J., 2024. n/a(n/a): p.
e202301100.
122. Yin, Q., et al., A Fast Soluble Carbon-Free Molecular Water Oxidation Catalyst Based on Abundant
Metals. Science, 2010. 328(5976): p. 342-345.
123. Ahmed, T., et al., High-nuclearity cobalt(II)-containing polyoxometalate anchored on nickel foam
as electrocatalyst for electrochemical water oxidation studies. Journal of Alloys and Compounds,
2022. 909: p. 164709.
124. Shi, J., et al., Polyoxometalate-based crown ether supramolecular hybrid electrocatalyst for the
hydrogen evolution reaction and robust nitrite reduction. Journal of Molecular Structure, 2024.
1296: p. 136831.
125. Li, J., et al., Determination of the stability constant of cobalt-substituted mono-lacunary Keggin-
type polyoxometalate and its electrocatalytic water oxidation performance. Journal of
Coordination Chemistry, 2017. 70(17): p. 2950-2957.
126. Mukhopadhyay, S., et al., A Keggin Polyoxometalate Shows Water Oxidation Activity at Neutral
pH: POM@ZIF-8, an Efficient and Robust Electrocatalyst. Angewandte Chemie International
Edition, 2018. 57(7): p. 1918-1923.
127. Tanaka, S., M. Annaka, and K. Sakai, Visible light-induced water oxidation catalyzed by
molybdenum-based polyoxometalates with mono- and dicobalt(iii) cores as oxygen-evolving
centers. Chemical Communications, 2012. 48(11): p. 1653-1655.
128. Liu, W., et al., Synthesis, Detailed Characterization, and Theoretical Understanding of
Mononuclear Chromium(III)-Containing Polyoxotungstates [CrIII(HXVW7O28)2]13– (X = P, As)
with Exceptionally Large Magnetic Anisotropy. Inorganic Chemistry, 2014. 53(17): p. 9274-9283.
129. Afrasiabi, R., M.R. Farsani, and B. Yadollahi, Highly selective and efficient oxidation of sulfides
with hydrogen peroxide catalyzed by a chromium substituted Keggin type polyoxometalate.
Tetrahedron Letters, 2014. 55(29): p. 3923-3925.
130. Gumerova, N.I., et al., Incorporation of CrIII into a Keggin Polyoxometalate as a Chemical
Strategy to Stabilize a Labile {CrIIIO4} Tetrahedral Conformation and Promote Unattended Single-
Ion Magnet Properties. Journal of the American Chemical Society, 2020. 142(7): p. 3336-3339.
131. Gebreaneniya, M.F., G.G. Berhe, and T. Teklu, Synthesis, Characterization, and Photocatalytic
Activity of Cu-Doped MgO Nanoparticles on Degradation of Methyl Orange (MO). Advances in
Materials Science and Engineering, 2024. 2024: p. 9969064.
132. Nayan, M.B., et al., Comparative Study on the Effects of Surface Area, Conduction Band and
Valence Band Positions on the Photocatalytic Activity of ZnO-
M&lt;sub&gt;x&lt;/sub&gt;O&lt;sub&gt;y&lt;/sub&gt; Heterostructures %J Journal of Water
Resource and Protection. 2019. Vol.11No.03: p. 14.
133. Trivedi, M., et al., Characterization of Physical, Thermal and Structural Properties of Chromium
(VI) Oxide Powder: Impact of Bio Field Treatment. Powder Metallurgy & Mining, 2015. 4.
134. Ci, S., et al., Rational design of mesoporous NiFe-alloy-based hybrids for oxygen conversion
electrocatalysis. 2015. 3(15): p. 7986-7993.
135. Wang, X., et al., High performance porous nickel cobalt oxide nanowires for asymmetric
supercapacitor. Nano Energy, 2014. 3: p. 119-126.
136. Lu, Y.-T., et al., Trapped interfacial redox introduces reversibility in the oxygen reduction reaction
in a non-aqueous Ca2+ electrolyte. Chemical Science, 2021. 12(25): p. 8909-8919.

70
137. Anantharaj, S. and S. Noda, Appropriate Use of Electrochemical Impedance Spectroscopy in
Water Splitting Electrocatalysis. ChemElectroChem, 2020. 7(10): p. 2297-2308.
138. Lazanas, A.C. and M.I. Prodromidis, Electrochemical Impedance Spectroscopy─A Tutorial. ACS
Measurement Science Au, 2023. 3(3): p. 162-193.
139. Sheng, W., H.A. Gasteiger, and Y. Shao-Horn, Hydrogen Oxidation and Evolution Reaction Kinetics
on Platinum: Acid vs Alkaline Electrolytes. Journal of The Electrochemical Society, 2010. 157(11):
p. B1529.
140. Haid, R.W., et al., Exploration of the electrical double-layer structure: Influence of electrolyte
components on the double-layer capacitance and potential of maximum entropy. Current
Opinion in Electrochemistry, 2022. 32: p. 100882.
141. Wang, Y., et al., Encapsulation and Stability Testing of Perovskite Solar Cells for Real Life
Applications. ACS Materials Au, 2022. 2(3): p. 215-236.

71

You might also like