• No results found

University of Groningen A journey into the coordination chemistry, reactivity and catalysis of iron and palladium formazanate complexes Milocco, Francesca

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen A journey into the coordination chemistry, reactivity and catalysis of iron and palladium formazanate complexes Milocco, Francesca"

Copied!
19
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

A journey into the coordination chemistry, reactivity and catalysis of iron and palladium

formazanate complexes

Milocco, Francesca

DOI:

10.33612/diss.160960083

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from

it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date:

2021

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Milocco, F. (2021). A journey into the coordination chemistry, reactivity and catalysis of iron and palladium

formazanate complexes. University of Groningen. https://doi.org/10.33612/diss.160960083

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 11PDF page: 11PDF page: 11PDF page: 11

1

Chapter 1

(3)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 12PDF page: 12PDF page: 12PDF page: 12

2

1

1.1

Catalytic transformations

A long series of disruptive and creative events have made possible life on Earth as we know it. The stellar nucleosynthesis, the self-replication, the metabolism, the photosynthesis, the transition from a reducing sulfur-rich atmosphere to an oxidizing oxygen-rich atmosphere, just to name a few. All these events, as well as a lot of process and items in our daily life (such as digestion, cars, beers, plastic, bread, detergents, etc.), have one thing in common: catalysis.

Appreciation of the subtle and delicate balance required for our life, can motivate us to find sustainable way of living in order to preserve it. There is an urgent need of re-thinking catalytic transformations and, in this context, fundamental research on novel types of reactivity is essential.

Homogeneous catalytic processes can be classified as reductive, oxidative or redox-neutral processes. In many cases, even if the reaction is formally redox-neutral, two-electron transformations are required to promote bond-forming and bond-breaking events. In particular, oxidative addition and reductive elimination are key steps in several catalytic cycles and these reactions are typical of square-planar complexes with 2nd and 3rd row transition metals. Noble metals are often the first choice in

catalysis thanks to their accessible Mn+ and M(n+2)+ oxidation states. A famous example of this kind of

catalysis is the palladium-catalyzed cross couplings, for which Heck, Negishi and Suzuki won the Nobel Prize in Chemistry in 2010.1

However, if on the one hand noble metals give outstanding results in catalysis, on the other hand their use is accompanied by many drawbacks, such as the fact that they are expensive, rare and toxic. Therefore, the replacement of noble metals with cheap, abundant and less toxic2 first-row transition

metals is highly desirable. The supremacy of noble metals in homogeneous catalysis over base metals comes from the differences in electronic structure. Noble metals generally prefer two-electron redox events, while base metals usually favor one-electron redox events leading to odd-electron pathways that are difficult to control.3 Therefore, one of the biggest challenges in order to employ more

extensively earth-abundant first-row transition elements in homogeneous catalysis is either manage to achieve control and selectivity over odd-numbered redox steps, or find a way to confer nobility on base metals. Moreover, another advantage for the late metals is their high affinity to π-bonds, which is a relevant feature because the chemical industry is mainly oil-based having as feedstocks alkenes, arenes and alkynes. On the other hand, the future direction is towards renewable raw materials and this could favor the employment of base metals.4 Regarding the need of the development of more

sustainable post-fossil energy systems, a lot of effort has been addressed to convert readily accessible and abundant small molecules (e.g. N2, O2, CO2, CH4, H2O, H2) into high-value chemical feedstocks and

fuels. Therefore, the challenge is to find the right conditions to make these mostly stable and inert molecules reactive and to control the reaction pathways that often involve multielectron redox processes coupled with proton transfer.5 To do so, it is fundamental to understand the bonding and

coordination properties of transition metals to such small molecules, by investigating the metal-ligand interaction in different oxidation and spin state.6

1.1.1 CO

2

conversion

While nature is based on circularity as exemplified by the relationship between photosynthesis and cellular respiration, the CO2 cycle has been broken by humankind since the industrial revolution.7 This

(4)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 13PDF page: 13PDF page: 13PDF page: 13

3

ppm,8 with drastic consequences for the climate change.9 Therefore, it is fundamental to reinvent

industrial processes in order to close the loop of that cycle and bring it back to balance.

Carbon capture utilization and storage (CCUS) could play a critical role in reaching net-zero emissions.

10 The most oxidized form of carbon is carbon dioxide and the conversion of this molecule is challenging

due to its high thermodynamic and kinetic stability. Nevertheless, the thermodynamic stability can be overcome by providing an energy input in the form of heat, electric current or radiation or by combining CO2 with high free energy compounds, such as hydrogen, amines, or epoxides, resulting in

an exergonic reaction. The kinetic stability can be tackled by using a suitable catalyst and different catalytic methods have been explored for CO2 conversion: ranging from homogeneous,

heterogeneous, electrochemical and photochemical catalysis.7, 11 For the sake of sustainability it is

essential to consider the choice of source of energy supply (renewable or integrated heat-management systems), source of chemical reagents to couple with CO2; of catalyst (based on earth abundant

metal).7a, 12

Scheme 1.1 depicts some of the possible reactions that use carbon dioxide as a feedstock to synthesize various added-value chemicals.7, 11b, f, 12-13

Scheme 1.1. Examples of CO2 conversion into various chemicals.

Many of the possible transformations involving CO2 require proton-coupled multielectron steps, which

are typically challenging for earth-abundant metals (see Section 1.2), nevertheless promising results have been shown in this field.11d, 14 Examples of CO

2 conversion with non-noble metal are reported in Section 1.4.2.

1.2

Development of non-noble metal catalysts for 2-electron

redox-transformations

One possible approach to replace rare metals with earth abundant ones is to mimic the noble behavior with base elements and this can be achieve by bringing the ligand framework into play. In classical homogeneous catalysis, the so-called ancillary or spectator ligands remain unchanged during the chemical transformation, which take place at the metal center, and their role is to modulate the

(5)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 14PDF page: 14PDF page: 14PDF page: 14

4

electronic and steric properties of the metal complex in order to achieve the desired reactivity. Conversely, actor ligands are those that engage in a chemical change. With the aim of developing novel catalytic reactivity, increasing attention has been addressed to the strategy of changing the role of the ancillary ligands making them actively participate at the transformation in several ways:15

i) cooperative ligands, which are directly involved in the bond activation reaction and they cooperate

with the metal in a synergistic approach called Metal-Ligand Cooperation (MLC).15b, c As a

consequence, they can be chemically modified either at the 1st coordination sphere or at the 2nd

coordination sphere. For example, one research line in our group is focused on the catalytic reactivity of ruthenium pincer (PNP and PNN) complexes, which is based on aromatization/dearomatization of the pyridine fragment of the ligand;16

ii) hemilabile ligands can be employed in order to create coordinative unsaturation at the metal

center;

iii) redox-non-innocent ligands, which can change their oxidation state and can be used to avoid

unstable oxidation steps during chemical redox transformations (see Section 1.4).

Many actor ligands combine more than one of the above strategies and perhaps that may be the origin of the confusion of terminologies adopted throughout the literature to address those ligands, e.g. "cooperative", "non-innocent", "redox-non-innocent", "redox-active" (see Section 1.3).

1

1.2.1 Iron

Iron is the Earth’s most abundant element by mass, fourth most abundant element of the crust and it is an essential trace element in almost all the living organisms (first most abundant transition metal in human body, a4 g).17 Iron abundancy in the Earth is related to its copious production by nuclear fusion

in high-mass stars, where it is the last chemical element to be formed accompanied by release of energy before the rapid and violent collapse of the massive star which scatters the chemical elements (including iron) into space.

This readily available, cheap and relatively nontoxic2 metal is a highly desirable candidate in catalysis.

In addition, its central location in the d-block could make it cover both early and late transition metal characters. Moreover, it is a redox active metal with its formal oxidation states going from –II to +VI, therefore, it can be useful both in reductive and oxidative chemistry.4

On the one hand, iron has made a breakthrough in heterogeneous catalysis (e.g. Haber-Bosh process to convert molecular nitrogen into ammonia and Fisher-Tropsch process to transform carbon monoxide into liquid fuel) and its vital role in biocatalysis is also well-established (e.g. heme group in the hemoglobin for oxygen transport or in the cytochrome P450 for oxygen activation), on the other hand, its employment in homogeneous catalysis has been for long time hampered by the established supremacy of noble metals in the field. Nevertheless, in the last 25 years the number of reports on this topic has significantly increased, showing the potentiality of iron catalysis.4, 17a, 18

As already mentioned in Section 1.1, the use of iron catalysts to mimic noble metals behavior has to deal with competing single electron transfer. Different strategies could be used to alter the inclination toward a certain type of redox action (1e− vs 2e):4

i) modulating the splitting of the d-orbital energy levels;

ii) achieving metal-ligand cooperativity thanks to “non-innocent” ligands;

iii) developing metal-metal cooperativity by the combination of two metal centers.

This thesis focuses on the first two strategies and the following section aims to uncover how they can be seen as two sides of the same coin.

(6)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 15PDF page: 15PDF page: 15PDF page: 15

5

1

1.3

Metal-ligand bonding: the electrons and orbitals dance

The electronic occupancy of the orbitals in a complex is controlled by the splitting (Δ) of the d-orbital energy levels, which can be modulated by the coordination geometry and the nature of the ligands. Commonly, the “normal ligand-field” splittings are two below three for the tetrahedral, three below two for the octahedral and four below one for the square-planar coordination (Figure 1.1).19

Figure 1.1. Most common ligand-field splittings.

In addition, strong field ligands (π-acceptor ligands) lead to larger splitting, whereas weak field ligands (π-donor ligands) result in smaller splitting. This feature is particularly relevant in the octahedral environment where ligand field theory can be used either to favor the electron pairing and therefore the formation of low-spin complexes or to encourage the occupancy of all the levels resulting in high-spin complexes.

When we construct the orbital interaction diagram for a transition metal ion interacting with the surrounding ligands (both for the σ-donor and π-donor character) we typically assume that the ligand orbitals are lower in energy compared to the metal d orbitals (Figure 1.2 left). This assumption is justified by the fact that the ionization potential of transition metal d orbitals is usually substantially smaller than those of typical Lewis base ligands.19

Figure 1.2. Ligand field molecular orbital for σ-bonding (in orange) and π-bonding (light blue) in different regimes:

classical Werner-type (left); covalent (middle) and inverted (right). Free-ion ΔTd t2g eg ΔTd= 4/9 ΔOh Octahedral Tetrahedral e t2 dxy dyz dxz dx2−y2 dz2 dx2−y2 dz2 dxy dyz dxz ΔOh Δ dx2−y2 dz2 dxy dyz dxz Square planar eg a1g b2g b1g

Classical Covalent Inverted

d (M−L)σorπ (M−L)* σorπ d d σorπ σorπ σorπ (M−L)σorπ (M−L)* σorπ (M−L)σorπ (M−L)* σorπ or or or or or or E

(7)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 16PDF page: 16PDF page: 16PDF page: 16

6

However, we can imagine that if we raise the ligand levels close (Figure 1.2 middle) or above (Figure 1.2 right) the metal levels, we are able to alter the molecular orbital splitting and potentially obtain

inverted ligand-field splittings with respect to Figure 1.1 (i.e. three below two for the tetrahedral, two

below three for the octahedral and one below four for the square-planar).19-20 In fact, cases of inverted

ligand fields are being more and more recognized in the chemistry comunitry.19-21

In those “non classical” cases, often one question arises: are the resulting frontier orbitals mainly centered on the metal or on the ligands? The fact that it is not always straightforward to answer to this question causes ambiguity in the oxidation assignment.

Moreover, metal-ligand covalency also affects the pairing energy (i.e. increasing covalency means increasing molecular orbital size and, therefore, diminishing the electron repulsion through the so-called nephelauxetic effect), which together with Δ, determines the ground-state electron occupancy in the orbitals.22 Clearly, the correct description of the spin state of a transition metal

complex is fundamental given that different structures, chemical properties and reactivity are associated with different spin states.22-23

Metal-ligand covalency and ligand-field inversion can lead to bonding schemes where the frontier

orbitals are molecular orbitals that are ligand localized rather than metal-localized. This is the feature of non-innocent/redox active ligands, which possess higher energy HOMO, or alternatively lower-lying LUMO, compared to those of typical innocent/redox inactive ligands and therefore can participate in electron transfer.18c, 20d

Often the terms non-innocent ligand (NIL) and redox-active ligand (RAL) are used interchangeably. Some reports, however, shed light on the differences between these terms.23a, 243a Moreover, it is

usually implicit that the first term, NIL, refers to redox non-innocent ligands.

The definition of the term "innocent" ligand goes back to 1966, when Jørgensen stated that “ligands are innocent when they allow oxidation states of the central atoms to be defined”.25 Conversely,

"non-innocent" ligands imply an uncertainty or ambiguity in the oxidation state assignment.23a, 24, 26

Chirik then pointed out in 201124 the fact that modern experimental and computational methods can

often identify the appropriate oxidation state of the metal. Therefore, the term "redox-active" ligand is more appropriately used in cases in which a well-defined redox process occurs at the ligands. As stated by Hofmann et al., "molecules do what comes naturally; it is we who, in our struggle to understand them, pigeonhole them into rigid categories and eventually run into trouble".19

1

1.4

Redox active ligands

The idea of redox-active ligands comes from Nature, which has developed metalloenzymes based on earth-abundant metal, such as iron and copper, even if these metals preferentially react via one-electron redox steps.27 To overcome the problem that several biological transformations are

multi-electron processes, redox-active ligands have been employed in the enzymatic active sites both to promote two-electron steps and to discourage potentially damaging radical reactions.3b, 24

In a complex with traditional innocent25 spectator ligands the redox event occurs at the metal center

because the oxidation or reduction of the ligands are too energetically demanding (Scheme 1.2 a). However redox-active ligands have energetically accessible levels for reduction or oxidation, therefore the changes in the oxidation state can take place either exclusively at the ligand (Scheme 1.2 b) or in a synergetic manner both at the ligand and the metal (Scheme 1.2 c).3As a result, redox-active ligands

(8)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 17PDF page: 17PDF page: 17PDF page: 17

7

structures, which could hopefully confer to first raw transition metals some noble catalytic properties.3b

Scheme 1.2. a) Innocent ligand; b) Redox-active ligand; c) Synergism between redox-active ligand and redox-active

metal.

One of the most famous examples of synergism between the metal and the ligand is the cytochrome P450, which is featured by a heme group in the active site and it is capable of selectively catalyzing monooxygenase reactions (Scheme 1.3).27The heme is constituted by a porphyrin ligand coordinated

to a Fe2+ center. The activation of O2 requires its two-electron reduction to the peroxo state and

subsequently the O–O bond cleavage takes place generating the so-called “Compound I”, which is a highly reactive intermediate considered responsible for the oxidation of the substrate. In 2010 Green

et al. were able to isolate and characterize Compound I, which resulted to be an iron(IV)oxo species

with a singly oxidized porphyrin radical ligand.28 Compound I promotes the hydroxylation of

hydrocarbons to alcohols for which one oxidant equivalent comes from the Fe(IV) and another from the porphyrin radical ligand, recovering the iron(III)porphyrin complex.

Scheme 1.3. (Right) Catalytic cycle of the hydroxylation of hydrocarbons by Cytochrome P450 and (left) structure

of Compound I.

Fascinatingly, the heme group is also responsible for oxygen transport and storage in the hemoglobin and myoglobin proteins. This shows the elegant fine-tuned character of the heme molecule which is capable to modulate its electronic structure and function and engage either in O2-activation as well as

reversible O2 binding. In this regard, spin crossover properties of the heme are relevant both in the

globins as well as in the cytochrome P450. Moreover, modulation of the extent of π-back bonding into the O-O bond is fundamental to achieve either weak and reversible binding or strong binding, conferring to the heme moiety the dual abilities of O2-transport and activation.17b

In general a redox-active ligand can act in four main ways:3a, 29

a) The oxidation/reduction of the ligand modifies the Lewis acidity/basicity of the metal. b) + 2 e− a) + 2 e− c) + 2 e− Synergism

(9)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 18PDF page: 18PDF page: 18PDF page: 18

8

b) The ligand acts as an electron reservoir preventing the metal to go in unstable oxidation states by delivering or accepting the electron density required from the multielectron process (e.g. Cytochrome P45027).

c) Formation of reactive ligand-radicals, which are actively involved in bond-making and breaking.

d) Activation of the substrate via ligand-to-substrate single electron transfer, allowing odd-electron reactivity.

1

1.4.1 Redox active ligands acting as electron reservoir for catalytic reactions

The term “electron reservoir” has been introduced for the first time in 1979 for 19-electron sandwiches, such as [FeI5-C5H5)(η6-C6Me6)], and it has been defined by the author as “a highly

reduced species which is one half of a totally reversible redox catalytic system in which both the oxidized and reduced forms are stable and isolable”.30 This terminology has then been used to describe

the ability of redox-active ligand to store electrons.3a, 29, 31 Even if redox-active ligands can provide

tremendous opportunities in homogeneous catalysis, this field started to be more systematically explored only in the last 15 years. Examples of redox-active ligands reported in literature are dithiolenes,32 dioxolenes,32 α-diimines,33 bis(imino)pyridines,31a, 34 bis(imino)acenaphtene,35

β-diketiminates,31c formazanate36 (see Section 1.5) and porphyrins,37 (Chart 1.1).

Chart 1.1. Examples of redox-active ligands.

Examples of how redox active ligands are used to donate or accept electrons, enabling new (catalytic) reactivity with base-metals, are shown below.

Chirik and Wieghardt highlighted that the iron complex (iPrPDI)Fe(N2)2 is a suitable precatalyst for

several reactions,3b, 31a, 34afor instance for hydrogenation (Scheme 1.4 a) and hydrosilylation (Scheme

1.4 b) of olefins,38 cyclization of enynes and dynes (Scheme 1.4 c)34band [2π + 2π] cycloaddition of

α,ω-dienes (Scheme 1.4 d).34c

Even if these reactions are overall two-electron processes, the iron(II) oxidation state is maintained and the required electrons are provide by the redox-active ligand, which acts as an electron reservoir.3b, 31aThe complex [Fe(iPrPDI)(N2)2] comes from the reduction of the [Fe(iPrPDI)(X)2] (X = Br, Cl)

(Scheme 1.5) and it was proved to be not the Fe(0) species with two neutral bis(imino)pyridine ligands, as initially described by the authors,38 but a Fe(II) species with the doubly reduced form of the

(10)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 19PDF page: 19PDF page: 19PDF page: 19

9

Scheme 1.4. Example of reactions catalyzed by the complex (iPrPDI)Fe(N

2)2.34b, c, 38

However, in a later review about non-innocent ligands18c the authors ascribed the difficulties in

assigning a definite oxidation state due to the high covalent character of the metal-ligand bond and they proposed that the complex (iPrPDI)Fe(N2)2 is better represented as a mesomeric equilibrium. This

way of describing the system also fits with the concept that this complex can be depicted in the framework of multi-state reactivities (MSR),39 meaning that the intrinsic reactivity of the complex is

related to different spin configurations. Hence, the bis(imino)pyridine ligands can act either as innocent ligand (e.g. Brookhart olefin polymerization catalyst40) or as non-innocent ligand with the

same metal, proving also the adaptability in term of spin state of iron.18c

Scheme 1.5. Reduction of (iPrPDI)Fe(II)(Cl)2 to (iPrPDI)Fe(N2)2 and mesomeric electronic configurations of

(iPrPDI)Fe(N2)2.18c, 31a

Another example of two-electron transformations which takes place without any changes in the oxidation state of the iron center is the reductive elimination of disulfide with an iron complex bearing the pincer redox-active ligand [ONO] (Scheme 1.6 a).41 Heyduk et al. reported a five coordinated

(11)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 20PDF page: 20PDF page: 20PDF page: 20

10

tert-butyldisulfide in the presence of 3 eq of pyridine (Scheme 1.6 b). The resulting complex is still an

Fe(III) species featuring a two-electron reduced ligand [ONOcat]3-.

Scheme 1.6. a) Redox active pincer ligands [ONO·]; b) Reductive elimination of a disulfide from [ONO]Fe complex.41

Considering the two above examples it has to be highlighted that redox-active [ONO] ligand was found to favor iron in its +3 oxidation state,41 while redox-active bis(imino)pyridine ligand favors its +2

oxidation state.3b, 31a Two possible reasons have been suggested:31a first, the harder nature of the

oxygen atoms and amide nitrogen of the [ONO] ligands compared to the one of pyridine and imine nitrogen of the diiminopyridine ligand should favor the iron(III) oxidation state; second, the oxidized forms of the [ONO] ligand (the radical semi-quinonate [ONOsq·]2- and the quinonate [ONOq]-) are

significantly stronger oxidants than the neutral diiminopyridine, therefore, they favor the iron(III) oxidation state.

1

1.4.2

Redox active ligands: a useful tool for CO

2

conversion

Iron porphyrins are known powerful electrocatalysts for the reduction of CO2 to CO42 and typically the

iron(0) complex, [Fe(TPP)]2-, was proposed to be the active species. However, porphyrins can act as

redox active ligands (as shown in Scheme 1.3 for the heme group), and in fact, spectroscopic and computational studies pointed out that [Fe(TPP)]2-, is best described as an intermediate spin iron(II)

center antiferromagnetically coupled to a porphyrin diradical dianion (Scheme 1.7).43

Scheme 1.7. Ligand-based reductions of [Fe(TPP)].

Therefore, the reduction is ligand-centered and the two electrons occupy the porphyrin π*-orbitals. Furthermore, a mechanistic study for the reduction of CO2 to CO was reported for the iron porphyrin

(12)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 21PDF page: 21PDF page: 21PDF page: 21

11

having four trimethyl ammonium groups in para position of the phenyl groups, where it was calculated that the Fe maintains its oxidation state of +2 during the entire catalytic cycle.44

Another example in which a redox active ligand is used to accommodate reducing equivalents, leading to interesting reactivity is the one reported by Peters et al. in 2014.45 A cobalt complex bearing the

redox active pyridyldiimine ligand, [CoIIIN4H(Br)2]+, was tested for electrocatalytic CO2 reduction in wet

MeCN and reduction of CO2 to CO was observed occurs near the formal CoI/0 redox couple (Scheme

1.8).45 Subsequently, the precatalyst [CoN

4H(MeCN)]+ has been isolated by protonation of the “CoI”

amido complex [CoN4], both of which were characterized by X-ray crystallography and DFT, which

elucidated that the pyridyldiimine moiety acts as an electron reservoir. In particular, it has been found that in both complexes the Co is not in the oxidation state +1, but they are low-spin CoII ions

antiferromagnetically coupled to a ligand radical-anion (N4H•- and N4•-). Later a combined

spectroscopic-theoretical study57 showed that CO

2 reacts with [CoN4H(MeCN)]+ forming a η1-C adduct,

[CoN4H(CO2)]+, where the CO2 ligand is only partially charged and moreover it can also be formed

from the complex [CoN4H(MeCN)]2+ (Scheme 1.8 bottom). From this study it is suggested that [CoN4H]

is not necessarily part of the electrocatalytic cycle and that the 2e− species could also be directly

generated from [CoN4H(CO2)]+. Nevertheless, both these two studies highlight the redox

non-innocence of the N4H ligand, which is actively involved in the CO2 activation in cooperation with

the metal center. 11b, 45-46

Scheme 1.8. Electrocatalytic CO2 reduction mediated by the complex [CoN4H(MeCN)]+.11b, 45-46

1

1.5

Formazanate ligands

Formazans are a class of colorful nitrogen-rich compounds containing the characteristic chain of atoms R1-NH-N=CR3-N=N-R5 which have been discovered at the end of the 1800s47 and have been then

reviewed by Nineham in 1955.48 Thanks to their intense color, formazans have found applications as

dyes,49 in chemical biology as redox-based staining agents for cell-viability assays50 and they raised

(13)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 22PDF page: 22PDF page: 22PDF page: 22

12

The backbone of formazans allows structural isomerization (Chart 1.2 a) and the different isomers vary in colors, ranging from the blood red color typical of the “closed” form to the orange and yellow colors of the “open” and “linear” isomers, respectively. The structural isomers of formazans reflects also the possible coordination modes of the deprotonate formazanate ligand to an inorganic element, allowing formation of six-, five- and four-membered chelates (Chart 1.2 b).36 While the coordination chemistry

and the catalytic applications of structural analogues N-donor ligands such as β-diketiminates (Chart 1.3 a) are well established,52 formazanate molecules have started to receive more attention as chelate

ligands only in the last decade36, 53 and examples of catalytic tranformations involving

formazanate-based catalysts still remain extremely rare.54

Chart 1.2. a) Structural isomers of formazans. b) Common coordination modes of formazanate ligands to an

inorganic element (E).

Recently, β-diketiminates, which have been usually considered robust ancillary ligands and have been employed to stabilize both very high and low oxidation state metal centres, were found to be non-innocent ligands that actively participate in numerous redox processes.31c, 52b, 55 However,

β-diketiminate complexes show a limited stability upon changing oxidation state. On the other hand, formazanate ligands are featured by highly enhanced redox-active properties and increased coordinative flexibility, thanks to the presence of two additional nitrogen atoms in the backbone (NNCNN instead of NCCCN).31b, c, 36, 56 Formazanate ligands are able to engage in both reductive and

oxidative redox chemistry due to high lying HOMO and low-lying LUMO orbitals of π-symmetry (see

Chapter 5).36 The stability of the reduced form of formazanate can be related to the known stability of

the organic verdazyl radical.56-57In fact, there is an isolobal relationship between the neutral verdazyl

radicals and the metallaverdazyl radical obtained from the reduction of the formazanate complex (Chart 1.3), where the unpaired electron can be delocalized over four nitrogen atoms.

Chart 1.3. Molecular structure of: a) β-diketiminate and formazanate ligands; b) verdazyl radical and

(14)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 23PDF page: 23PDF page: 23PDF page: 23

13

1

1.5.1 Redox activity of formazanate ligands

The redox-active nature of formazanate ligands was described for the first time only in 2007 by Hicks

et al.56 They reported that the complex [B(L1)(OAc)2] (L1 = PhNNC(p-Tol)NNPh) shows a

quasi-reversible reduction wave in the cyclic voltammetry and they proved that is ligand-based by chemically synthesizing the single-reduced complex [B(L1)(OAc)2][Cp2Co] and characterizing it via EPR and UV-Vis

spectroscopy (Scheme 1.9 a).

Scheme 1.9. Selected examples of isolated single and double- formazanate-centered reduced compounds. 31b, 56, 58

The redox properties of formazanate ligands have been then established in our group by exploring the coordination chemistry of those ligands with the redox inactive metal Zn(II).31b, 59 Taking advantage of

the modular synthesis of the formazanate ligands, which allows the introduction of a variety of different substitution patterns, a set of bis(formazanate)zinc complexes have been synthesized and their redox behavior has been investigated by cyclic voltammetry.59 In all the cases the ability of storing

at least one electron in each formazanate moiety was observed, giving access to the redox series L2Zn0/1−/2−, and in some compounds even the second reduction of the ligand was feasible, extending

the series to L2Zn3−/4−.59 In addition, the single and double reduced compounds were chemically

synthesized (Scheme 1.9 b) and characterized by X-Ray, DFT, EPR and UV-Vis spectroscopy, proving that one or two electrons can be stored in the ligand framework.31b

Furthermore, our group has also studied the coordination chemistry of formazanate to group 13 elements (B and Al)58a, b, 60 and for the first time a compound containing the formazanate moiety in

(15)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 24PDF page: 24PDF page: 24PDF page: 24

14

the trianionic form was structurally characterized first with boron58a and then with aluminum,58b

[E(L1)(Ph)2][Na2] (E = B, Al), (Scheme 1.9 c).

1

1.5.2 Iron formazanate complexes

An interesting case is the one of iron formazanate complexes, where the redox active ligands are combined with a redox active transition metal. Two examples can be considered to illustrate how playing with the ligand-field splitting and pursuing metal-ligand cooperativity through the use of “non-innocent” ligands can be seen as two sides of the same coin (Section 1.3).

In 2016, our group reported an unusual pseudo-tetrahedral bis(formazanate) iron(II) complex, [Fe(L1)2]

(1), which undergoes thermal switching between a low-spin (S = 0) and a high-spin (S = 2) state.61

Typically, if we consider the case of Fe(II) d6 in an octahedral environment, two possibilities can be

depicted: a high spin complex in which ΔO is smaller than the pairing energy and the electrons prefer

to occupy also the higher orbitals, or a low spin complex where the splitting is higher and the electrons are paired in the lower orbitals. However, for a tetrahedral coordination the high spin state is always the favorite one, even with strong-field ligands, due to the decreased splitting of the d-orbitals (Δt =

4/9 ΔO). Octahedral complexes in which these two states are close in energy can exhibit the

phenomenon of spin-crossover (SCO); they can switch between low-spin (LS) and high-spin (HS) states using external stimuli such as temperature, light or pressure.62 The fact that a complex has two

accessible spin states, which typically have completely different structures, reactivity and spin density distributions,23a has a relevant impact on mechanisms, rates and selectivity in organometallic reaction

steps relevant to catalysis, as highlighted by Schröder, Shaik and Schwarz with the two-state reactivity model (TSR).39b

However, four-coordinate complexes that show SCO are very rare. Few examples have been reported in the literature with iron: the three-fold symmetric phosphinimido iron complexes with a tris(carbene)borate ligand by Smith et al.,63 the square-planar bis(imino)pyridines iron imidos by Chirik et al.,64 the pseudo-tetrahedral tris(phosphine)borate phosphiniminato iron complexes by Peters et al.65 and the pseudo-tetrahedral bis(formazanate) iron complex by our group61 (see Chart 2.1, Chapter 2). These works are example of how the “normal ligand-field” splitting can be inverted19 as the result

of forced geometries and alteration of the electronic structure exerted by the ligands. Therefore, the deviation from the classical tetrahedral ligand-field splitting makes possible the switch between low and high spin. Specifically, the unusual occurrence of spin-crossover in the bis(formazanate) iron(II) complex (1) was ascribed to π-back interaction between the iron center and the low-lying formazanate π*-orbital, resulting in significant stabilization of one of the high-energy orbitals (t2).61 This leads to ligand-field inversion and a 2-over-3 splitting, reminiscent of that of six-coordinate octahedral

(16)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 25PDF page: 25PDF page: 25PDF page: 25

15

Figure 1.3. Molecular orbital diagram for [Fe(L1)2] (1) in the LS (S=0) state and the HS (S=2) state compared to the

classical tetrahedral.Adapted from ref 61 (https://pubs.acs.org/doi/10.1021/jacs.6b01552) with permission from the American Chemical Society. Further permissions related to the material excerpted should be directed to the ACS.

Interestingly, chemical one-electron reduction of 1 allows the isolation of the anionic complex [Fe(L1)2][NBu4] and structural, spectroscopic and computational data demonstrated that the singly

reduced compound is best described as a low-spin (S = 1/2) Fe(I) complex featuring closed-shell, monoanionic formazanate ligands. Therefore, despite the redox-active nature of the formazanate ligands the first reduction in this case preferentially occurs at the iron center.

A different situation has been later reported by Holland et al. for a mono(formazanate) iron(II) amide complex [Fe(L1)(N(SiMe3)2)(THF)] and its one-electron reduction product

[Fe(L1)(N(SiMe3)2)(THF)][Na(12-crown-4)2] (Scheme 1.9 d), for which a variety of spectroscopic

techniques suggest formazanate-centered reduction. Multiconfigurational calculations identified a quartet ground state that reproduces the empirical spectroscopic data (Stotal = 3/2). Furthermore, two

configurations were found to be dominant, contributing of approximately 25% each. Therefore, according to the calculations, the ground state is best described as a virtually equally mix of two electronic configurations: a high spin Fe(II) center (SFe = 2) antiferromagnetically coupled to the

unpaired electron located in the formazanate π*-orbital (SL = 1/2) and a high-spin Fe(I) (SFe = 3/2)

without ligand participation.58c

It is important to underline that, despite the different involvement of the formazanate ligand in the reduction chemistry in the two iron complexes reported by Otten61 and Holland,58c in both cases the

peculiar features described for those compounds arise from the presence of a low-lying empty π*-orbital, which in one case is involved in π-backdonation61 and in the other case allows formazanate

participation to the redox event.58c This emphasizes that is feasible for formazanate iron complexes to

achieve a synergism between the redox-active ligand and the redox-active metal and that further fundamental research on the electronic structure of these compounds is needed.

dz2 dx2-y2 dyz dxz dxy π* π* dz2 dx2-y2 dyz dxy dxz Classical tetrahedral d6 LS S = 0 HS S = 2 1 Formazanate π-acceptor ligand e t2

(17)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 26PDF page: 26PDF page: 26PDF page: 26

16

1

1.6

Thesis outline

In this thesis we present an in-depth study of the coordination chemistry of formazanate ligands to iron and palladium and we explore the (catalytic) reactivity of the resulting complex toward small molecules.

In Chapter 2, the synthesis and characterization of five novel bis(formazanate) iron complexes, [FeL2]

capable of undergoing spin-crossover is described. We demonstrate that is possible to tune the SCO properties of those compounds by electronic substituent effects.

In Chapter 3, we extend the study of how the spin-crossover properties of bis(formazanate) iron complexes may be modulated via modification of the ligand via different strategies: electronic effects, steric effects, π-stacking interactions and ligand denticity.

In Chapter 4, the reactivity of bis(formazanate) iron(I) and (II) complexes toward CO and isocyanide is investigated. Novel low spin 6-coordinate complexes, [Fe(L)2(CNAr)2], are synthesized and

characterized.

In Chapter 5, the redox behavior of bis(formazanate) iron complexes was examined via cyclic voltammetry. Oxidative addition on a bis(formazanate) iron(I) complex gave useful insight on the poor stability of these compounds in the Fe(III) state. We then show that isolation of a stable Fe(III) formazanate iron complex is possible with a tridentate NNO ligand, leading to an example of "hidden" non-innocence.

In Chapter 6, we report the synthesis of formazanate ferrate(II) dihalides complexes ([FeLX2][NBu4], X

= Br, Cl) via salt metathesis, which provide a straightforward synthetic pathway toward mono(formazanate) complexes. The halides in the resulting compounds are shown to be labile, opening up possibilities for further reactivity.

In Chapter 7, the aforementioned mono(formazanate) iron(II) dihalide complexes, are used as highly selective single component catalyst for the conversion of CO2 into cyclic carbonates starting from

various terminal and internal epoxides.

In Chapter 8, the synthesis and characterization of a mono(formazanate) alkyl palladate complex, [Pd(L)(CH3)(Cl][NBu4], is reported and ligand substitution and insertion reactions are explored.

1.7

References

1. J. E. Bäckvall, The Royal Swedish Academy of Sciences, 2010, 1-12.

2. K. S. Egorova and V. P. Ananikov, Angew. Chem. Int. Ed., 2016, 55, 12150-12162.

3. (a) D. L. Broere, R. Plessius and J. I. van der Vlugt, Chem. Soc. Rev., 2015, 44, 6886-6915; (b) P. J. Chirik and K. Wieghardt, Science, 2010, 327, 794-795.

4. A. Fürstner, ACS Cent. Sci., 2016, 2, 778-789.

5. (a) B. Milani, G. Licini, E. Clot and M. Albrecht, Dalton Trans., 2016, 45, 14419-14420; (b) C. C. Lu and K. Meyer, Eur. J. Inorg. Chem., 2013, 2013, 3731-3732; (c) F. Meyer and W. B. Tolman, Inorg. Chem., 2015,

54, 5039.

6. D. Benito-Garagorri, I. Lagoja, L. F. Veiros and K. A. Kirchner, Dalton Trans., 2011, 40, 4778-4792. 7. (a) K. M. Yu, I. Curcic, J. Gabriel and S. C. Tsang, ChemSusChem, 2008, 1, 893-899; (b) M. Aresta, A.

Dibenedetto and A. Angelini, Chem. Rev., 2014, 114, 1709-1742.

8. https://www.climate.gov/news-features/understanding-climate/climate-change-atmospheric-carbon-dioxide.

9. https://www.ipcc.ch/sr15/.

10. Energy Technology Perspectives 2020, IEA, Paris, 2020.

11. (a) A. J. Kamphuis, F. Picchioni and P. P. Pescarmona, Green Chem., 2019, 21, 406-448; (b) R. Francke, B. Schille and M. Roemelt, Chem. Rev., 2018, 118, 4631-4701; (c) E. Boutin, L. Merakeb, B. Ma, B. Boudy,

(18)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 27PDF page: 27PDF page: 27PDF page: 27

17

M. Wang, J. Bonin, E. Anxolabehere-Mallart and M. Robert, Chem. Soc. Rev., 2020; (d) H. Takeda, C. Cometto, O. Ishitani and M. Robert, ACS Catalysis, 2017, 7, 70-88; (e) C. Costentin, M. Robert and J. M. Saveant, Chem. Soc. Rev., 2013, 42, 2423-2436; (f) E. E. Benson, C. P. Kubiak, A. J. Sathrum and J. M. Smieja, Chem. Soc. Rev., 2009, 38, 89-99.

12. (a) A. W. Kleij, M. North and A. Urakawa, ChemSusChem, 2017, 10, 1036-1038; (b) J. Artz, T. E. Muller, K. Thenert, J. Kleinekorte, R. Meys, A. Sternberg, A. Bardow and W. Leitner, Chem. Rev., 2018, 118, 434-504.

13. Q. Liu, L. Wu, R. Jackstell and M. Beller, Nat. Commun., 2015, 6, 5933.

14. (a) N. Hazari, N. Iwasawa and K. H. Hopmann, Organometallics, 2020, 39, 1457-1460; (b) K. E. Dalle, J. Warnan, J. J. Leung, B. Reuillard, I. S. Karmel and E. Reisner, Chem. Rev., 2019, 119, 2752-2875. 15. (a) V. T. Annibale and D. Song, RSC Adv., 2013, 3; (b) J. R. Khusnutdinova and D. Milstein, Angew. Chem.

Int. Ed., 2015, 54, 12236-12273; (c) H. Grutzmacher, Angew. Chem. Int. Ed., 2008, 47, 1814-1818; (d) J.

I. van der Vlugt, Eur. J. Inorg. Chem., 2012, 2012, 363-375.

16. (a) J. Bootsma, B. Guo, J. G. de Vries and E. Otten, Organometallics, 2020, 39, 544-555; (b) B. Guo, J. G. de Vries and E. Otten, Chem. Sci., 2019, 10, 10647-10652; (c) B. Guo, D. S. Zijlstra, J. G. de Vries and E. Otten, ChemCatChem, 2018, 10, 2868-2872.

17. (a) C. Bolm, Nat. Chem., 2009, 1, 420; (b) K. P. Kepp, Coord. Chem. Rev., 2017, 344, 363-374.

18. (a) C. Bolm, J. Legros, J. Le Paih and L. Zani, Chem. Rev., 2004, 104, 6217-6254; (b) I. Bauer and H. J. Knolker, Chem. Rev., 2015, 115, 3170-3387; (c) S. Blanchard, E. Derat, M. Desage-El Murr, L. Fensterbank, M. Malacria and V. Mouriès-Mansuy, Eur. J. Inorg. Chem., 2012, 2012, 376-389.

19. R. Hoffmann, S. Alvarez, C. Mealli, A. Falceto, T. J. Cahill, 3rd, T. Zeng and G. Manca, Chem. Rev., 2016,

116, 8173-8192.

20. (a) R. K. Szilagyi, B. S. Lim, T. Glaser, R. H. Holm, B. Hedman, K. O. Hodgson and E. I. Solomon, J. Am.

Chem. Soc., 2003, 125, 9158-9169; (b) K. M. Carsch, I. M. DiMucci, D. A. Iovan, A. Li, S. L. Zheng, C. J.

Titus, S. J. Lee, K. D. Irwin, D. Nordlund, K. M. Lancaster and T. A. Betley, Science, 2019, 365, 1138-1143; (c) I. M. DiMucci, J. T. Lukens, S. Chatterjee, K. M. Carsch, C. J. Titus, S. J. Lee, D. Nordlund, T. A. Betley, S. N. MacMillan and K. M. Lancaster, J. Am. Chem. Soc., 2019, 141, 18508-18520; (d) J. T. Lukens, I. M. DiMucci, T. Kurogi, D. J. Mindiola and K. M. Lancaster, Chem. Sci., 2019, 10, 5044-5055.

21. (a) J. S. Steen, G. Knizia and J. Klein, Angew. Chem. Int. Ed., 2019, 58, 13133-13139; (b) S. Mukherjee, D. E. Torres and E. Jakubikova, Chem. Sci., 2017, 8, 8115-8126.

22. S. Alvarez and J. Cirera, Angew. Chem. Int. Ed., 2006, 45, 3012-3020.

23. (a) O. R. Luca and R. H. Crabtree, Chem. Soc. Rev., 2013, 42, 1440-1459; (b) R. Poli, Chem. Rev., 1996,

96, 2135-2204; (c) R. Poli and J. N. Harvey, Chem. Soc. Rev., 2003, 32, 1-8.

24. P. J. Chirik, Inorg. Chem., 2011, 50, 9737-9740. 25. C. K. Jørgensen, Coord. Chem. Rev., 1966, 1, 164-178. 26. W. Kaim, Inorg. Chem., 2011, 50, 9752-9765.

27. L. Que, Jr. and W. B. Tolman, Nature, 2008, 455, 333-340. 28. J. Rittle and M. T. Green, Science, 2010, 330, 933-937. 29. V. Lyaskovskyy and B. de Bruin, ACS Catal., 2012, 2, 270-279.

30. D. Astruc, J. R. Hamon, G. Althoff, E. Roman, P. Batail, P. Michaud, J. P. Mariot, F. Varret and D. Cozak,

J. Am. Chem. Soc., 1979, 101, 5445-5447.

31. (a) S. C. Bart, K. Chlopek, E. Bill, M. W. Bouwkamp, E. Lobkovsky, F. Neese, K. Wieghardt and P. J. Chirik,

J. Am. Chem. Soc., 2006, 128, 13901-13912; (b) M.-C. Chang, T. Dann, D. P. Day, M. Lutz, G. G. Wildgoose

and E. Otten, Angew. Chem. Int. Ed., 2014, 53, 4118-4122; (c) C. Camp and J. Arnold, Dalton Trans.,

2016, 45, 14462-14498.

32. M. D. Ward and J. A. McCleverty, J. Chem. Soc., Dalton Trans., 2002, 275-288.

33. H. Tsurugi, T. Saito, H. Tanahashi, J. Arnold and K. Mashima, J. Am. Chem. Soc., 2011, 133, 18673-18683. 34. (a) S. C. Bart, E. Lobkovsky, E. Bill, K. Wieghardt and P. J. Chirik, Inorg. Chem., 2007, 46, 7055-7063; (b) K. T. Sylvester and P. J. Chirik, J. Am. Chem. Soc., 2009, 131, 8772-8774; (c) M. W. Bouwkamp, A. C. Bowman, E. Lobkovsky and P. J. Chirik, J. Am. Chem. Soc., 2006, 128, 13340-13341.

35. M. J. Supej, A. Volkov, L. Darko, R. A. West, J. M. Darmon, C. E. Schulz, K. A. Wheeler and H. M. Hoyt,

Polyhedron, 2016, 114, 403-414.

36. J. B. Gilroy and E. Otten, Chem. Soc. Rev., 2020.

37. L. K. Blusch, K. E. Craigo, V. Martin-Diaconescu, A. B. McQuarters, E. Bill, S. Dechert, S. DeBeer, N. Lehnert and F. Meyer, J. Am. Chem. Soc., 2013, 135, 13892-13899.

38. S. C. Bart, E. Lobkovsky and P. J. Chirik, J. Am. Chem. Soc., 2004, 126, 13794-13807.

39. (a) S. Shaik, H. Hirao and D. Kumar, Acc. Chem. Res, 2007, 40, 532-542; (b) D. Schroder, S. Shaik and H. Schwarz, Acc. Chem. Res., 2000, 33, 139-145.

(19)

554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco 554702-L-bw-Milocco Processed on: 2-2-2021 Processed on: 2-2-2021 Processed on: 2-2-2021

Processed on: 2-2-2021 PDF page: 28PDF page: 28PDF page: 28PDF page: 28

18

40. (a) B. L. Small and M. Brookhart, J. Am. Chem. Soc., 1998, 120, 7143-7144; (b) G. J. P. Britovsek, V. C. Gibson, S. J. McTavish, G. A. Solan, A. J. P. White, D. J. Williams, G. J. P. Britovsek, B. S. Kimberley and P. J. Maddox, Chem. Commun., 1998, 849-850.

41. J. L. Wong, R. H. Sánchez, J. G. Logan, R. A. Zarkesh, J. W. Ziller and A. F. Heyduk, Chem. Sci., 2013, 4. 42. (a) I. Bhugun, D. Lexa and J.-M. Savéant, J. Am. Chem. Soc., 1996, 118, 1769-1776; (b) M. Hammouche,

D. Lexa, J. M. Savéant and M. Momenteau, J. Electroanal. Chem. Interfacial Electrochem., 1988, 249, 347-351; (c) C. Costentin, G. Passard, M. Robert and J. M. Saveant, J. Am. Chem. Soc., 2014, 136, 11821-11829; (d) C. Costentin, S. Drouet, G. Passard, M. Robert and J. M. Saveant, J. Am. Chem. Soc., 2013,

135, 9023-9031; (e) C. Costentin, G. Passard, M. Robert and J. M. Saveant, Proc Natl Acad Sci U S A,

2014, 111, 14990-14994.

43. (a) C. Romelt, J. Song, M. Tarrago, J. A. Rees, M. van Gastel, T. Weyhermuller, S. DeBeer, E. Bill, F. Neese and S. Ye, Inorg. Chem., 2017, 56, 4746-4751; (b) B. Mondal and S. Ye, Coord. Chem. Rev., 2020, 405. 44. Y.-Q. Zhang, J.-Y. Chen, P. E. M. Siegbahn and R.-Z. Liao, ACS Catalysis, 2020, 10, 6332-6345. 45. D. C. Lacy, C. C. McCrory and J. C. Peters, Inorg. Chem., 2014, 53, 4980-4988.

46. M. Zhang, M. El-Roz, H. Frei, J. L. Mendoza-Cortes, M. Head-Gordon, D. C. Lacy and J. C. Peters, J. Phys.

Chem. C, 2015, 119, 4645-4654.

47. (a) E. Bamberger and J. Müller, Ber. Dtsch. Chem. Ges., 1894, 27, 147–155; (b) H. V. Pechmann, Berichte

der deutschen chemischen Gesellschaft, 1894, 27, 320-322; (c) H. v. Pechmann, Berichte der deutschen chemischen Gesellschaft, 1892, 25, 3175-3190.

48. A. W. Nineham, Chem. Rev., 1955, 55, 355–483.

49. (a) Y. Zhang, Dyes and Pigments, 1995, 29, 57-63; (b) S. A. Khan, S. Shahid, S. Kanwal and G. Hussain,

Dyes and Pigments, 2018, 148, 31-43; (c) K. Bauer, D. Garbe and H. Surberg, 2002.

50. (a) E. Grela, J. Kozlowska and A. Grabowiecka, Acta Histochem., 2018, 120, 303–311; (b) J. C. Stockert, R. W. Horobin, L. L. Colombo and A. Blázquez-Castro, Acta Histochem., 2018, 120, 159–167.

51. A. S. Shawali and N. A. Samy, J. Adv. Res., 2015, 6, 241-254.

52. (a) L. Bourget-Merle, M. F. Lappert and J. R. Severn, Chem. Rev., 2002, 102, 3031–3065; (b) R. L. Webster, Dalton Trans., 2017, 46, 4483-4498.

53. G. N. Lipunova, T. G. Fedorchenko and O. N. Chupakhin, Russ. J. Gen. Chem., 2019, 89, 1225–1245. 54. (a) A. Schafer, C. Huber and R. Ahlrichs, J. Chem. Phys., 1994, 100, 5829-5835; (b) I. S. Pavlova, I. G.

Pervova, G. P. Belov, I. I. Khasbiullin and P. A. Slepukhin, Pet. Chem., 2013, 53, 127-133; (c) A. V. Zaidman, I. G. Pervova, A. I. Vilms, G. P. Belov, R. R. Kayumov, P. A. Slepukhin and I. N. Lipunov, Inorg.

Chim. Acta, 2011, 367, 29-34; (d) N. M. R. Martins, K. T. Mahmudov, M. F. C. Guedes da Silva, L. M. D.

R. S. Martins and A. J. L. Pombeiro, New J. Chem., 2016, 40, 10071-10083; (e) A. Iqbal, M. G. Moloney, H. L. Siddiqui and A. L. Thompson, Tetrahedron Lett., 2009, 50, 4523-4525; (f) S. Hong, L. M. Hill, A. K. Gupta, B. D. Naab, J. B. Gilroy, R. G. Hicks, C. J. Cramer and W. B. Tolman, Inorg. Chem., 2009, 48, 4514-4523; (g) S. J. Hong, A. K. Gupta and W. B. Tolman, Inorg. Chem., 2009, 48, 6323−6325.

55. M. M. Khusniyarov, E. Bill, T. Weyhermuller, E. Bothe and K. Wieghardt, Angew. Chem. Int. Ed., 2011,

50, 1652-1655.

56. J. B. Gilroy, M. J. Ferguson, R. McDonald, B. O. Patrick and R. G. Hicks, Chem. Commun., 2007, 126-128. 57. M. C. Chang and E. Otten, Chem. Commun., 2014, 50, 7431-7433.

58. (a) R. Mondol, D. A. Snoeken, M.-C. Chang and E. Otten, Chem. Commun., 2017, 53, 513-516; (b) R. Mondol and E. Otten, Inorg. Chem., 2019, 58, 6344-6355; (c) D. L. J. Broere, B. Q. Mercado, J. T. Lukens, A. C. Vilbert, G. Banerjee, H. M. C. Lant, S. H. Lee, E. Bill, S. Sproules, K. M. Lancaster and P. L. Holland,

Chem. Eur. J., 2018, 24, 9417–9425.

59. M.-C. Chang, P. Roewen, R. Travieso-Puente, M. Lutz and E. Otten, Inorg. Chem., 2015, 54, 379-388. 60. (a) R. Mondol and E. Otten, Inorg. Chem., 2018, 57, 9720-9727; (b) R. Mondol and E. Otten, Dalton

Trans., 2019, 48, DOI: 10.1039/C1039DT02831E.

61. R. Travieso-Puente, J. O. P. Broekman, M.-C. Chang, S. Demeshko, F. Meyer and E. Otten, J. Am. Chem.

Soc., 2016, 138, 5503-5506.

62. (a) P. Gütlich and H. A. Goodwin, eds., Spin Crossover in Transition Metal Compounds I-III, Springer,

2004; (b) M. Halcrow, Crystals, 2016, 6, 58.

63. (a) J. J. Scepaniak, T. D. Harris, C. S. Vogel, J. Sutter, K. Meyer and J. M. Smith, J. Am. Chem. Soc., 2011,

133, 3824-3827; (b) H.-J. Lin, D. Siretanu, D. A. Dickie, D. Subedi, J. J. Scepaniak, D. Mitcov, R. Clérac

and J. M. Smith, J. Am. Chem. Soc., 2014, 136, 13326-13332.

64. A. C. Bowman, C. Milsmann, E. Bill, Z. R. Turner, E. Lobkovsky, S. DeBeer, K. Wieghardt and P. J. Chirik,

J. Am. Chem. Soc., 2011, 133, 17353-17369.

Referenties

GERELATEERDE DOCUMENTEN

Thus, a series of bisformazanatemagnesium complexes n-Mg, their corresponding monoTHF adducts n-MgTHF and bisformazanatecalcium complexes n-Ca complexes are presented and compared

Scheme 4.4 Synthesis of tetrazepine derivatives 4.1 and 4.2 via nucleophilic aromatic substitution using KH as base.. These findings contribute to a new synthetic route

For azobenzene, rotation and inversion mechanisms are discussed in the literature as possible thermal reaction paths from Z to E.36,37 Transition state calculations for 4.3

The thermodynamic parameters (∆H and ∆S) that describe the spin state equilibrium in solution were determined by fitting the temperature-dependence of three independent sets

Throughout this thesis, we study the coordination chemistry of the redox active formazanate ligands to iron and palladium and we explore the (catalytic) reactivity of the

Deze heteroleptische, tetraëdrische complexen hebben een heel andere elektronische structuur in vergelijking met de homoleptische verbinding 1, geïllustreerd door de hoge

Thanks to the technical staff for all your precious help: Oetze (thank you for nor retiring before the end of my PhD!), Pieter and Johan (for all the chats during my long

A journey into the coordination chemistry, reactivity and catalysis of iron and palladium formazanate complexes..