• No results found

University of Groningen Redox-behavior and reactivity of formazanate ligands Mondol, Ranajit

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Redox-behavior and reactivity of formazanate ligands Mondol, Ranajit"

Copied!
25
0
0

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

Hele tekst

(1)

Redox-behavior and reactivity of formazanate ligands

Mondol, Ranajit

DOI:

10.33612/diss.107969043

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: 2019

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Mondol, R. (2019). Redox-behavior and reactivity of formazanate ligands: Boron and aluminum chemistry. University of Groningen. https://doi.org/10.33612/diss.107969043

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)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 13PDF page: 13PDF page: 13PDF page: 13

Chapter 1

(3)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 14PDF page: 14PDF page: 14PDF page: 14 2

1.1 General introduction

Elementary steps involved in the catalytic cycles for organic transformations are often two-electron processes (e.g., oxidative addition and reductive elimination reactions). These reactions are predominantly carried out by the noble metals (e.g., Pd, Pt, Rh, Ru, etc.) as they have (n) and (n+2) oxidation states that are close in energy and therefore can be accessed in a reversible manner. But, the use of noble metals is not sustainable as they are rare, expensive and often toxic, and above all the resources of noble metals are limited.1,2 That is why there is a

considerable interest in the community to move from noble metal catalysts to ones based on earth-abundant (hence cheaper) and less toxic elements (base metal or main group elements). But this presents several challenges.

For example, base metals (first row transition metals) usually change their oxidation states by one-electron due to their relatively stable (n) and (n+1) oxidation states, which limits their utility in the aforementioned two-electron processes. The utilization of main group complexes as catalysts is even much more challenging. In contrast to the noble metal complexes, where close-lying filled and empty d-orbitals interact synergistically to make/break bonds via (reversible) two-electron changes in the metal oxidation state (reductive elimination / oxidative addition), main group complexes have large separation between filled (s) and empty (p) orbitals does generally not allow for (reversible) redox-reactions with these compounds. Now, the question is how to impose noble metal characteristics to complexes with base metals or main group elements so that these could perform above mentioned two-electron processes. A possible answer to this question is the incorporation of redox-active ligands in base metal3–8 or main

group9–11 complexes. Redox-active ligands can donate or accept electrons to or from substrates

during chemical transformations, and could provide access to two-electron redox-reactions by stabilizing coordination compounds in unusual (formal) oxidation states. Thus, base metal or main group complexes having redox-active ligands may perform catalytic reactions that are not possible with conventional ligands.

1.2 Redox-active ligands

Conventionally, transition metal catalysts change their oxidation state during the catalytic cycle, and the coordinated ligands are not involved in the redox reactions. These coordinated ligands give stabilization to the metal complex during catalysis by acting as a supporting ligand, and also, these ligands allow to tune properties of catalysts by changing the steric and electronic properties via ligand substituent effect. In contrast, when the ligand actively participates in the redox-reaction by accepting or donating redox-equivalents into its backbone, the ligand is

(4)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 15PDF page: 15PDF page: 15PDF page: 15 3

defined as a ‘redox non-innocent’ ligand.12–14 Particularly, when the oxidation states of metal

and ligand in the complex are clearly defined, then the ligand is identified as ‘redox-active’. Due to the presence of energetically accessible frontier orbitals, redox non-innocent/redox-active ligands can accept or donate electrons during chemical redox transformations, and allow the metal to avoid going through unstable oxidation states, and thus, enables challenging chemical transformations.

The following section describes how natural and non-natural catalysts enable challenging chemical transformations with base metal complexes by utilizing the redox-active ligands.

1.2.1 Utilization of redox-active ligands by Nature

Nature is a pioneer in the use of redox-active ligands in organic transformations and catalysis. Nature uses base metals in the active sites of several metalloenzymes (e.g., Fe in P-450,15 Cu

in galactose oxidase,16 etc.) in which either the redox-active organic moiety is in very close

proximity to, or directly bound to the active metal center. In these metalloenzymes, both the active site and the redox-active moiety actively participate as redox-equivalents in the organic transformation. This synergistic effect between metal and ligand allow challenging multi-electron redox reactions. For example, galactose oxidase oxidase16 oxidizes primary alcohols

to aldehyde, which is a two-electron process.

(Inactive) - e -OH2 CuII O (581H)N (496H)N (495Y)O (Active) RCH2OH - H 2O2 RCHO S (C228) (T272) OH2 CuII O (581H)N (496H)N (495Y)O (C228)S (T272) O CuII O (581H)N (496H)N (495Y)O (C228)S (T272) CuI O (581H)N (496H)N S (C228) (T272) CuII O (581H)N (496H)N S (C228) (T272) R OH H2O R HH R H O O2 H2O2 Galactose Oxidase H net reaction: + O2 H (495Y)O O R HH H (495Y)O H

(5)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 16PDF page: 16PDF page: 16PDF page: 16 4

During substrate oxidation, one electron is stored on the Cu(II) active site by reduction to Cu(I) and a second electron is taken up by the tyrosine, redox-active ligand scaffold (Scheme 1.1). Thus, copper maintains an energetically favorable Cu(II)/Cu(I) redox-cycle (without going to the energetically unfavorable Cu(II)/Cu(0) redox-cycle).

Another example where Nature uses redox-active ligands for catalytic transformation is cytochrome P450s, which perform monooxygenation and have been found in all domains of life.17 The active site of cytochrome P450 contains a heme cofactor which is constituted by a

Fe+2 center and a porphyrin ligand (Chart 1.1).18 In 2010, Green and co-workers captured and

characterized a highly reactive species, an intermediate of the catalytic cycle of hydrocarbon to alcohol conversion, named as P450 compound I (P450-I).15 The P450-I contains an

iron(IV)-oxo species and a singly oxidized porphyrin radical ligand (Chart 1.1). The P450-I oxidizes hydrocarbons to alcohols, which is a two-electron process. During this oxidation process, one electron is stored on the iron center (reduces Fe(IV) to Fe(III)) and another electron is taken up by the ligand to convert the porphyrin radical cation into the parent porphyrin.

Chart 1.1 Active site of cytochrome p450 (left) and compound I (right)

N N N N FeIV HOOC HOOC Cytochrome P450 O N N N N FeIII HOOC HOOC P450 Compound I

1.2.2 Bio-inspired and artificial redox-active ligands

Taking inspiration from these enzymatic systems, several efforts have been made for the development of other classes of redox-active ligands. For examples, oxygen-based catecholate (OO) ligands7,19–27 and its nitrogen (NN)27–32-, nitrogen-oxygen (NO)7,26,27,33,34-, sulfur (SS)35– 37- derivatives, and the porphyrin38–44 type ligands are well known (Chart 1.2). In recent years,

several pincer type tridentate redox-active ligands [i.e., (ONO), (NNN), (SNS)- types] have been developed.27,45–51 Also, two new nitrogen-rich pyridine-based redox-active ligands, such

as pyridine imino ligands (IP) and pyridine diimino ligands (PDI), have gained much attention in recent years.4,7,9–11,26,52–61 Structures and various oxidation states of IP and PDI

(6)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 17PDF page: 17PDF page: 17PDF page: 17 5

ligands are shown in Chart 1.2. Redox-active properties of these newly synthesized redox-active ligands have been utilized in catalyst development with base metals bearing these redox-active ligands.3–8,62,63

Chart 1.2 General structure (top, left), and different oxidation states (middle) of the catecholate

type ligands, porphyrin type ligands (top, right), various oxidation states of pyridine imino ligands (IP) (bottom, left) and various oxidation states of pyridine diimino ligands (PDI) (bottom, right) O -O -O -N -N -N -R R S -S -N N N N M R R R R R X -X -X -X X X - e- - e -L L2- L

-a) cat b) o-O,N c) o-N,N d) o-S,S

X = O, N, S, P N N R R NR R N N R R - e- N N R R - e- N N R R - e- - e -N N R R NR R N N R R NR R IP IP IP

2-PDI PDI PDI 2-+ e- + e

-+ e- + e- + e

-+ e

-In the next paragraphs, among the several examples one representative example is presented to show the progress of catalyst development with the base metals bearing redox-active ligands.

1.2.2.1 Terminal alkene hydrosilylation catalyzed by iron complexes

containing redox-active ligands

The Chirik group has been involved in the investigation of redox-active properties of nitrogen based tridentate pyridine(diimine) ligands (PDI).4,14,64,65 They have utilized redox-active

properties of PDI ligands coordinated to several base metals, which mediated unusual catalytic chemical transformations. Examples include iron catalyzed hydrogenation and hydrosilylation of olefins,6,64 iron and cobalt catalyzed [2+2] cycloaddition of dienes,4,58,66,67 iron catalyzed

hydrogenative cyclization of enynes and diynes,3 cobalt catalyzed asymmetric alkene

(7)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 18PDF page: 18PDF page: 18PDF page: 18 6

metal catalysts can perform multi-electron processes by utilizing redox-active ligands. Here is an example of catalytic hydrosilylation of alkenes catalyzed by a [Fe(PDI)(N2)]2(μ-N2) complex,

which has a FeII center coordinated by a triplet diradical [PDI]2- ligand and dinitrogen molecules

(Scheme 1.2).6 In the first step of this catalytic cycle, the alkene is coordinated to the vacant

site at iron. In the second step, the Si-H bond of Et3Si-H is added via oxidative addition to the

iron center. In this step, two electrons are supplied by the reduced ligand backbone, not by the iron center, and thus the iron center maintains its stable Fe(II) oxidation state without going through the unstable Fe(IV) oxidation state. Then, the migratory insertion of hydride into the internal position of the coordinated alkene occurs, and is followed by reductive elimination to form the product and regenerate the active catalyst. During the reductive elimination process, two electrons are stored on the ligand backbone, not on the metal center, and thus again, the iron center maintains its stable Fe(II) oxidation state without going through the unstable Fe(0) oxidation state. Thus, throughout the catalytic cycle the iron center maintains its stable Fe(II) oxidation state and all the redox-processes are enabled by the redox-active ligand.

N N N Ar Ar Fe N2 N2 N N N Ar Ar FeII N N N Ar Ar FeII N N N Ar Ar FeII Ar = 2,6-(Me)2-C6H3 C5H11 C5H11 C5H11 Et3Si Et3Si H C5H11 N N N Ar Ar FeII Et3Si Et3Si-H C5H11 Alkene Hydrosilylation net reaction: C5H11 + Et3Si-H C5H11 Et3Si cat [Fe] neat, 23° C cat [Fe] Scheme 1.2 Proposed catalytic cycle for the Fe-catalyzed hydrosilylation of alkene

1.2.2.2 Redox reactions in main group complexes

(8)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 19PDF page: 19PDF page: 19PDF page: 19 7

show the recent advancement towards catalyst development based on main group elements. The reactivity of main group compounds and their application in catalysis is mostly based on the availability of an empty orbital (i.e., Lewis acidic character) or a lone pair of electrons (Lewis base). In main group complexes, the large separation between filled (s) and empty (p) orbitals does not generally allow for (reversible) redox-reactions with these complexes. In spite of this drawback, in the past decades, new strategies have been developed to stabilize main group complexes in low-valent states, in particular for the heavier congeners, and this has led to the development of reactions that can be carried out under mild conditions and are reminiscent of transition metal chemistry (e.g., small-molecule activation).70 Low-valent

compounds of carbon (N-heterocyclic carbenes, NHCs) have gained prominence as ligands in coordination chemistry and as organocatalysts,71 because of their singlet electronic ground state

that results from a large HOMO-LUMO gap. In 2007, Bertrand and co-workers reported seminal work on the reactivity of related alkyl amino carbenes (AACs), which were shown to cleave strong bonds in H2 and NH3 at a single carbon center (Chart 1.3 a).72 The increased

reactivity of AACs in comparison to the (unreactive) NHCs was rationalized based on the significant lowering of the LUMO energy level to bring it close in energy to the HOMO (a hallmark of the d-orbitals in transition metal compounds). The resulting combination of the nucleophilic and electrophilic characters at the carbene C-atom allows (formal) oxidative addition of H2 and NH3. Following this discovery, activation of small molecules with low-valent

main-group compounds has been studied extensively.70,73–77 For example, the oxidative

addition of NH3 to stannylene followed by reductive elimination to form B-N and B-H bonds

was reported by Aldridge and co-workers (Chart 1.3 b), opening the possibility to develop catalytic transformations based on a Sn(II)/Sn(IV) redox cycle.78 A recently developed

Si(II)/Si(IV) redox-cycle reported by Inoue and co-workers is worthy to mention.79 In this study,

they showed that a highly reactive acyclic iminosilylsilylene underwent intramolecular insertion into a C=C bond to form silacycloheptatriene, which converted back to Si(II) at elevated temperatures (Chart 1.3 c).79 As an example of related chemistry in group 15,

Radosevich and co-workers developed P(III)/P(V) redox cycles to cleave O-H and N-H bonds in a reversible manner (Chart 1.3 d).80 In recent years, oxidative addition of strong bonds to

Roesky’s β-diketiminate aluminum(I) species, (DippNC(Me)CHC(Me)NDipp)Al (Dipp = 2,6-diisopropylphenyl, Chart 1e)81 has been explored in detail.74,77 The reaction between the

aluminum(I) complex and the corresponding dihydride was shown to form an equilibrium between the starting materials and the mixed dimer (Chart 1.3 f), indicating that (formal)

(9)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 20PDF page: 20PDF page: 20PDF page: 20 8

oxidative addition / reductive elimination of the Al-H bonds is reversible in this system.82 In

2019, Aldridge and co-workers showed the reversible oxidative addition of C-C bond of benzene to a single aluminum center of isolated monomeric aluminyl compound [K(2.2.2-

Chart 1. 3 Redox reactions in the main group complexes

B Sn BN N N N R R R R BN N R R NH2 BN N R R H reduced Sn species NAl N Dipp Dipp N Al N Dipp Dipp + H H N Al N Dipp Dipp H NAl N Dipp Dipp H NMe N MeN P NMe N MeN P OR H (b) (e) (d) RT ∆ + N Dipp Bertrand (a) Aldridge Nikonov Radosevich B Sn BN N N N R R R R H NH2 NH3 NH3 N Dipp H2N H RO-H R = Dipp O tBu tBu N N Al Dipp Dipp O tBu tBu N N Al Dipp Dipp [K(2,2,2-Crypt)] [K(2,2,2-Crypt)] Benzene, room temp, 48 h Benzene-d6, 80 °C Benzene, 80 °C O tBu tBu N N Al Dipp Dipp [K(2,2,2-Crypt)] D D D D D D (f) Aldridge N Al N N Ph Ph Dipp Dipp THF Cl 2H+/1e -1e -H2 L L N Al N N Ph Ph Dipp Dipp L L H H Cl L = 4-Dimethylaminopyridine (DMAP) Berben N N (g) N NDipp Dipp NSiBr3 2 KSiTMS3 -2 KBr -TMS3SiBr N N C N Si (TMS)3Si C N N N (c) -78°C, toluene C Si C (TMS)3Si Si(II) Si(IV) Inoue

(10)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 21PDF page: 21PDF page: 21PDF page: 21 9

crypt)][(NON)Al] (Chart 1.3 g) at room temperature, which further adds a significant contribution to the field of catalyst development based on main group compounds.73,83 In this

context, small molecule activation and catalysis by frustrated Lewis pairs (FLPs) is an another sub-field of main group chemistry, which has gained much attention since its inception in 2006 and a staggering progress has been achieved. Several recent reviews describe the advances in FLP chemistry,84–94 and the reader is referred to these articles for an in-depth discussion of this

chemistry.

Despite the advances in stabilization of low-valent main group species and their use to cleave strong bonds via oxidative addition, the reverse reaction (i.e., reductive elimination) still presents considerable challenges in main group chemistry due to the stability of the highest oxidation state in particular for the lighter (2nd and 3rd row) elements. As a consequence,

catalytic turnover involving the oxidative addition / reductive elimination sequences are rarely observed in main group chemistry. In this respect, similar to base metal complexes, the incorporation of redox-active ligands in main group complexes may allow to perform chemical transformations catalytically. In this context, the Berben group has reported a series of reduced Al(III) and Ga(III) complexes with redox-active imino pyridine (IP) and diimino pyridine (PDI) ligands.9–11,52–55,60,61 The Berben group has utilized the metal-ligand cooperative property and

the redox-active property of IP and PDI ligands for several chemical transformations, ranging from dehydrogenative coupling of aniline and benzyl amine, catalytic conversion of CO2 into

MgCO3 or CaCO3, conversion of formic acid into CO2 and H2, electrocatalytic hydrogen and

ammonia production etc.10,11,53,54 From these above mentioned examples, ligand-centered

electrocatalytic hydrogen production is one of the most promising achievements towards catalyst development based on main group complexes bearing redox-active ligands. For this catalytic hydrogen evolution reaction, the locus of protonation and reduction is the redox-active PDI ligand, and the metal center (Al) is not involved (Chart 1.3 f).10

1.3 Formazanate ligands

1.3.1 General features of formazans

Formazanates are monoanionic (deprotonated form of formazan) bidentate N-chelating ligands, which are nitrogen-rich (having NNCNN backbone) analogues of β-diketiminate ligands (also known as NacNac ligands). The coordination chemistry of β-diketiminate ligands has been explored extensively during the past decades. Due to easy accessibility and high

(11)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 22PDF page: 22PDF page: 22PDF page: 22 10

stereoelectronic tunability, a variety of metal complexes of the β-diketiminate ligands have been synthesized, characterized, and several of these metal complexes have found applications as catalysts.95–103 Apart from being used as supporting ligands, in the last few years, their

redox-active properties have also been investigated and well established.104

Formazan has a long history of being used as dye molecules.105,106,107 In modern chemical

research, it has been used in the biochemical research field.108,109 Formazan has been used as a

precursor of thermally stable verdazyl radicals (Chart 1.4).110 Due to the presence of four

nitrogen atoms in the backbone, the verdazyl radical has a low-energy SOMO, which has π anti-bonding character between N-N bonds (the SOMO is shown in Chart 1.4). The unpaired electron of the verdazyl radical is delocalized over all four nitrogen atoms of this heterocyclic ring, which gives exceptional stability to these verdazyl radicals.111,112 As verdazyl radicals have

a low-energy SOMO and a high-energy HOMO, they can be reversibly oxidized to cations and reduced to anions. Due to the exceptional stability of verdazyl radicals to air and moisture, these radicals have found several applications, such as in the preparation of molecular magnets,113 as

ESR spin labels114 and inhibitors for polymerization,115 or mediators in living radical

polymerizations.116

Chart 1.4 Verdazyl radical (left) and SOMO of verdazyl radical (right)

In spite of the structural similarity of formazanate ligands with β-diketiminate ligands, only in recent years attention has been given to the exploration of their coordination chemistry. The presence of two extra nitrogen atoms in the formazanate ligand backbone compared to the structurally related β-diketiminate ligand, introduces a significant amount of flexibility in the coordination chemistry of formazanate ligands. Due to the increased flexibility in the geometry of formazanate ligands, these ligands not only form common 6-membered chelate rings in their metal complexes, but also 5-membered as well as 4-membered chelate rings are accessible (Scheme 1.3).117,118 N N N N R5 R1 R3 R6 Verdazyl radical R5 R1 R3 R6 SOMO

(12)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 23PDF page: 23PDF page: 23PDF page: 23 11 N N R N N N N Zn N N N N p-tol Mes C6F5 C6F5 Mes p-tol N N N N Zn N N N N p-tol Mes C6F5 C6F5 Mes p-tol N N M N R Ar M = Na, K N Ar N N Ar Ar M or

Scheme 1.3 Various coordination modes of the formazanate ligands117,118

1.3.2 Redox-active properties of formazanate ligands

In 2007, the first example of ligand-based reduction of a formazanate ligand was reported by Hicks and co-workers in the formazanate boron diacetate complex LB(OAc)2 (Scheme 1.4).119

Cyclic voltammetry experiment on LB(OAc)2 shows a quasi-reversible redox event at -0.84 V

vs Fc0/+1, which indicates the possibility to access the reduced complex of LB(OAc)2. They have

reduced LB(OAc)2 with cobaltocene to afford the reduced product [LB(OAc)2][Cp2Co]. This

reduced product was characterized by EPR and UV-vis absorption spectroscopic techniques. This study not only reveals the redox-active property of formazanate ligands, but also indicates the easy accessibility of the reduced product of LB(OAc)2 by using mild reducing agents

(reduction potential of LB(OAc)2 = -0.84 V vs Fc0/+1).

N N NHN Ph Ph p-tol N N N N Ph Ph p-tol B(OAc)3 B AcO OAc Cp2Co N N N N Ph Ph p-tol B AcO OAc Cp2Co

Scheme 1.4 Ligand-based reduction in a formazanate boron complex (reported by Hicks and

co-workers)119

Inspired by the work from Hicks, Gilroy and co-workers have prepared a variety of boron complexes with redox-active formazanate ligands.120–131 They have extensively studied the

optical and electrochemical properties of these newly generated materials. All of these complexes absorb strongly in the visible region due to the π-π* transition within the ligand frameworks.120,121,123–127 Cyclic voltammetry experiments reveal that most of the formazanate

boron complexes show two sequential one-electron (quasi)reversible redox events that are ligand-centered.120,121,124–127 The first redox event corresponds to the formation of radical

(13)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 24PDF page: 24PDF page: 24PDF page: 24 12

the optical (e.g., absorption and emission maxima, fluorescence, etc.) and the electrochemical (e.g., reduction potentials) properties of these (formazanate)boron complexes could be modulated via the systematic variation of substituents as well as via the extension in π conjugation within the ligand framework.120,121,124–127 Very recently, they have expanded the

study on formazanate complexes by incorporating heavier analogues of group 13 elements (e.g., aluminum)132 and higher congeners of group 14 elements (e.g., Si, Ge and Sn)133. The

formazanate complexes of aluminum and group 14 elements also disclose redox-active properties of formazanate ligands. Furthermore, their studies not only reveal the tunable optical and electrochemical properties of the (formazanate)boron complexes but also some of these complexes have photoluminescent properties121,124,125 which has led to applications as

cell-imaging agents,125,128 electrochemiluminescence emitters,126,134 multifunctional polymers,131,135

and precursors to a variety of BN heterocycles.136

Concurrent with the Gilroy group, our group also started a research program to explore the coordination chemistry and the redox properties of formazanate ligands. In 2014, our group has studied the ligand-based redox chemistry of bis(formazanate)zinc complexes.137 Cyclic

voltammetry experiments indicate four sequential one-electron redox waves for bis(formazanate)zinc complexes (Figure 1.1). The first and second redoxevents occurred at -1.54 and -1.84 V vs Fc0/+1, respectively. These two redox-events correspond to the

ligand-centered one-electron reduction of each ligand, which leads to the formation of radical anions and radical dianions, respectively, which were isolated and structurally characterized. The stability of the reduced products of bis(formazanate)zinc complexes are due to their “metalaverdazyl”-type structure.137 Further redox-events that are observed at more negative

potentials (< -2.5 V vs Fc0/+) are due to the two-electron reduction of each ligand (Figure

1.1).117,137 N N N N Ph Ph p-tol L -N N N N Ph Ph p-tol 2 L 2-N N N N Ph Ph p-tol L 3-e -I/I' and II/II' e -III/III' and IV/IV' I' I II III IV IV' III' II'

(14)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 25PDF page: 25PDF page: 25PDF page: 25 13

In the same year, reduction chemistry of the mono(formazanate)boron difluoride complexes (LBF2) was investigated.138 Cyclic voltammetry of the complex LBF2 shows two sequential

(quasi)reversible one-electron reduction waves at -0.98 and -2.06 V Fc0/+1, respectively.138 This

indicates the possibility of two-electron reduction in a single formazanate ligand. Though the one-electron reduction product [LBF2]- was isolated and characterized, the isolation of the

two-electron reduced product [LBF2]2- was unsuccessful due to the elimination of NaF, which led to

the formation of N-heterocyclic boron carbenoid species, which react further and become incorporated in a series of unusual BN heterocycles.139

In our group, including the above examples, several other formazanate boron and zinc complexes have been prepared and their optical properties (e.g., absorption and emission spectra) and redox properties have been studied.117,137–140 In 2016, our group has described the

synthesis of a series of (formazanate)boron complexes and their ligand-centered one-electron reduced products. The neutral complexes are fluorescent, whereas the reduced products do not show any fluorescent properties.140 This study shows the possibility to use (formazanate)boron

complexes as redox-switchable dyes. Recently, our group also expanded the coordination chemistry and redox chemistry of formazanate complexes by employing iron and palladium metals.141–144 Due to the π-acceptor properties of formazanate ligands, the

bis(formazanate)iron complex obtains an unusual electronic structure and shows switchable magnetism (spin-crossover).141

In recent years, the group of Holland has also reported ligand-based reductions in several low-coordinate iron complexes with the formazanate ligand.145,146 In 2016, Sundermeyer and

co-workers have described the synthesis and characterization of series of formazanate complexes incorporating heavier analogues of group 13 elements (LMMe2, M = Al, Ga and In).147

In the last decade, several transition metal (Ru,148 Co,149 Ni,150–152 Pd,153 Pt,154–156 Cu,157 and

other158) complexes with redox-active formazanate ligands have been reported. Despite the fact

that a wide range of reported formazanate complexes is known, only few complexes are active catalysts for chemical transformations, which includes H2O2 decomposition159 and ethylene

oligomerization.153,150,160 But, in both these chemical transformations formazanates act as

‘innocent’ supporting ligands without engaging in redox-activity. Thus, in order to utilize the redox-active properties of formazanate ligands in chemical reactivity, there is a need to further enhance the understanding of bonding and electronic structural features of neutral and reduced formazanate complexes.

(15)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 26PDF page: 26PDF page: 26PDF page: 26 14

1.4 Overview of the thesis

The overall aim of the research presented in this thesis is the exploration of the redox-active properties and the study of ligand-centered reactivity of formazanate complexes. In order to achieve this goal, we have synthesized boron and aluminum complexes with formazanate ligands and their redox chemistry has been studied. For the first time, two-electron reduced mono(formazanate) boron and aluminum complexes were isolated and fully characterized, and subsequently, ligand-based reactivity with these reduced complexes is explored.

In chapter 2, the synthesis and characterization of the redox-series of boron diphenyl

complexes with formazanate ligands (i.e., [LBPh2]0/-1/-2) is presented. First, we have established

the redox-active features of the formazanate ligands in mono(formazanate) boron diphenyl complexes (LBPh2) by cyclic voltammetry. For the first time, ligand-based two-electron

reduced products of formazanate complexes ([LBPh2]-2) are isolated. The redox-series LBPh2 0/-/2- is fully characterized by single crystal structure determination and spectroscopic techniques.

The experimental data are corroborated by a computational study (density functional theory). In chapter 3, we have explored the ligand-based reactivity of the two-electron reduced

formazanate boron diphenyl complex (i.e., [LBPh2]2-). The reactions of nucleophilic [LBPh2]2-

with the electrophiles BnBr and H2O lead to the formation of benzylated and

ligand-protonated products, respectively, which are anionic boron analogues of leucoverdazyls. N-C and N-H bond homolysis of these systems is studied by the exchange NMR spectroscopy and kinetic experiments. This study reveals that not only formazanate ligands can store two electrons and an electrophile (E+ = Bn+, H+) on their backbones, but these can react as source

of E• radicals. This type of reactivity is very important in energy storage applications such as

hydrogen evolution.

In chapter 4, the synthesis of a series of aluminum complexes with redox-active formazanate

ligands is described. The cyclic voltammograms of mono(formazanate)aluminum diphenyl and diiodide and bis(formazanate)aluminum chloride complexes are presented. Chemical reduction of all the newly synthesized complexes shows that ligand-based redox reactions in these compounds are accessible. The two-electron reduction of mono(formazanate)aluminum diiodide was performed in order to investigated the possibility to obtain a formazanate aluminum(I) carbenoid species. A computational study further sheds light on the electronic structure and the stability of sub-valent formazanate aluminum(I) carbenoid species.

In chapter 5, the differences in structure, bonding and reactivity between two-electron reduced

(16)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 27PDF page: 27PDF page: 27PDF page: 27 15

containing a highly reduced, trianionic formazanate-derived ligand is described. Due to the substantial delocalization of excess electron density from the ligand core (i.e., NNCNN) onto the peripheral ligand substituents (N-Ph) via the π-framework, partial double bond character is built up in the N-C(Ph) bonds which leads to the restricted rotation around N-C(Ph) bonds. The barrier of this restricted rotation around N-C(Ph) bonds in the two-electron reduced formazanate aluminum complex has been determined by variable temperature NMR experiments (VT-NMR) and NMR lineshape analysis, and is compared to the boron analogues. The crystallographic data of ligand-benzylated products of [(PhNNC(p-tol)NNPh)ZPh2]2- (Z = B, Al), which are anionic

B and Al analogues of leucoverdazyls, are presented and the main differences are addressed. The kinetics of benzyl transfer from the ligand-benzylated products to TEMPO reveal that the more ionic Al compound has one of the weakest N-C bonds reported so far in this type of inorganic leucoverdazyl analogues.

In chapter 6, a detailed analysis of the observed dynamics present in ligand-benzylated

formazanate boron and aluminium complexes ([BnLZPh2]-, where BnL =

(PhNN(Bn)C(p-tol)NNPh), and Z = B, Al) is presented. The activation parameters for the dynamic process present in these ligand-benzylated products have been determined by variable temperature NMR experiments and subsequent NMR lineshape analyses, which indicates that the observed dynamics in these complexes is due to nitrogen inversion. On the basis of these activation parameters, we have proposed a plausible mechanism for the nitrogen inversion processes present in these ligand-benzylated products. Our study reveals that the counter cation which is bound to the ligand backbone of [BnLBPh2]-/[BnLAlPh2]- moiety, dissociates in a rate

determining step that precedes nitrogen inversion. The effects of counter-cations on the dynamic processes present in these anionic B/Al complexes has been investigated. The comparison of activation parameters for the nitrogen inversion processes between boron and aluminum complexes indicates that the central element has little influence on the nitrogen inversion. In chapter 3, we described sequential ligand-based storage of two electrons (2e-) and one proton

(H+) (i.e., [2e-/H+] equivalent) in the formazanate boron diphenyl compound LBPh2, which led

to the formation of ligand-protonated product [HLBPh2]-. In addition, we showed a net H-atom

transfer from [HLBPh2]- to TEMPO. This net H-atom transfer from [HLBPh2]- to TEMPO could

occur either by hydrogen atom transfer (HAT) or by concerted proton-coupled electron transfer (cPCET) mechanism. In chapter 7, a detailed computational analysis of the mechanism of this

reaction is described by analysis of the changes in intrinsic bond orbitals (IBOs) along the reaction coordinate.

(17)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 28PDF page: 28PDF page: 28PDF page: 28 16

1.5 References

(1) Spargo, P. L. Transition Metals for Organic Synthesis; 2nd Ed; WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim, G. 2004. .

(2) Diederich, F.; Stang, P. J. Metal-catalyzed cross-coupling reactions; 2nd Ed; WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim, G. 2008. .

(3) Sylvester, K. T.; Chirik, P. J. J. Am. Chem. Soc. 2009, 131 (25), 8772–8774.

(4) Chirik, P. J.; Wieghardt, K. Science (80-. ). 2010, 327 (5967), 794–795.

(5) Russell, S. K.; Lobkovsky, E.; Chirik, P. J. J. Am. Chem. Soc. 2011, 133 (23), 8858–8861.

(6) Tondreau, A. M.; Atienza, C. C. H.; Weller, K. J.; Nye, S. A.; Lewis, K. M.; Delis, J. G. P.; Chirik, P. J.

Science (80-. ). 2012, 335 (6068), 567–570.

(7) Lyaskovskyy, V.; de Bruin, B. ACS Catal. 2012, 2 (2), 270–279.

(8) Broere, D. L. J.; Plessius, R.; Van Der Vlugt, J. I. Chem. Soc. Rev. 2015, 44 (19), 6886–6915.

(9) Myers, T. W.; Berben, L. A. Chem. Commun. 2013, 49 (39), 4175–4177.

(10) Thompson, E. J.; Berben, L. A. Angew. Chem., Int. Ed. 2015, 54 (40), 11642–11646.

(11) Sherbow, T. J.; Thompson, E. J.; Arnold, A.; Sayler, R. I.; Britt, R. D.; Berben, L. A. Chem. - A Eur. J.

2019, 25 (2), 454–458.

(12) Chaudhuri, P.; Nazari Verani, C.; Bill, E.; Bothe, E.; Weyhermüller, T.; Wieghardt, K. J. Am. Chem. Soc.

2001, 123 (10), 2213–2223.

(13) Herebian, D.; Bothe, E.; Bill, E.; Weyhermüller, T.; Wieghardt, K. J. Am. Chem. Soc. 2001, 123 (41),

10012–10023.

(14) Chirik, P. J. Inorg. Chem. 2011, 50 (20), 9737–9740.

(15) Rittle, J.; Green, M. T. Science (80-. ). 2010, 330 (6006), 933–937.

(16) Que, L.; Tolman, W. B. Nature 2008, 455 (7211), 333–340.

(17) Lamb, D. C.; Lei, L.; Warrilow, A. G. S.; Lepesheva, G. I.; Mullins, J. G. L.; Waterman, M. R.; Kelly, S. L. J. Virol. 2009, 83 (16), 8266–8269.

(18) Poulos, T. L.; Finzel, B. C.; Howard, A. J. J. Mol. Biol. 1987, 195 (3), 687–700.

(19) Pierpont, C. G. Coord. Chem. Rev. 2001, 216–217, 99–125.

(20) Pierpont, C. Coord. Chem. Rev. 2001, 219–221, 415–433.

(21) Pierpont, C. G. Inorg. Chem. 2011, 50 (20), 9766–9772.

(22) Lippert, C. A.; Amstein, S. A.; David Sherrill, C.; Soper, J. D. J. Am. Chem. Soc. 2010, 132 (11), 3879–

3892.

(23) Tezgerevska, T.; Alley, K. G.; Boskovic, C. Coord. Chem. Rev. 2014, 268, 23–40.

(24) Bittner, M. M.; Lindeman, S. V.; Popescu, C. V.; Fiedler, A. T. Inorg. Chem. 2014, 53 (8), 4047–4061.

(25) Dinda, S.; Genest, A.; Rösch, N. ACS Catal. 2015, 5 (8), 4869–4880.

(26) Luca, O. R.; Crabtree, R. H. Chem. Soc. Rev. 2013, 42 (4), 1440–1459.

(27) van der Vlugt, J. I. Chem. - A Eur. J. 2019, 25 (11), 2651–2662.

(28) Mederos, A.; Dominguez, S.; Hernandez-Molina, R.; Sanchiz, J.; Brito, F. Coord. Chem. Rev. 1999, 193–

195, 913–939.

(29) Kraft, S. J.; Fanwick, P. E.; Bart, S. C. J. Am. Chem. Soc. 2012, 134 (14), 6160–6168.

(18)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 29PDF page: 29PDF page: 29PDF page: 29 17

777.

(31) Rajabimoghadam, K.; Darwish, Y.; Bashir, U.; Pitman, D.; Eichelberger, S.; Siegler, M. A.; Swart, M.; Garcia-Bosch, I. J. Am. Chem. Soc. 2018, 140 (48), 16625–16634.

(32) Bai, Y.; Chen, W.; Li, J.; Cui, C. Coord. Chem. Rev. 2019, 383, 132–154.

(33) Poddel’sky, A. I.; Cherkasov, V. K.; Abakumov, G. A. Coord. Chem. Rev. 2009, 253 (3–4), 291–324.

(34) Coughlin, E. J.; Qiao, Y.; Lapsheva, E.; Zeller, M.; Schelter, E. J.; Bart, S. C. J. Am. Chem. Soc. 2019, 141

(2), 1016–1026.

(35) Petrenko, T.; Ray, K.; Wieghardt, K. E.; Neese, F. J. Am. Chem. Soc. 2006, 128 (13), 4422–4436.

(36) Sproules, S.; Wieghardt, K. Coord. Chem. Rev. 2010, 254 (13–14), 1358–1382.

(37) Garreau-de Bonneval, B.; Moineau-Chane Ching, K. I.; Alary, F.; Bui, T. T.; Valade, L. Coord. Chem. Rev.

2010, 254 (13–14), 1457–1467.

(38) Takahashi, A.; Kurahashi, T.; Fujii, H. Inorg. Chem. 2011, 50 (15), 6922–6928.

(39) Blusch, L. K.; Craigo, K. E.; Martin-Diaconescu, V.; McQuarters, A. B.; Bill, E.; Dechert, S.; Debeer, S.; Lehnert, N.; Meyer, F. J. Am. Chem. Soc. 2013, 135 (37), 13892–13899.

(40) Blusch, L. K.; Mitevski, O.; Martin-Diaconescu, V.; Pröpper, K.; Debeer, S.; Dechert, S.; Meyer, F. Inorg.

Chem. 2014, 53 (15), 7876–7885.

(41) Goswami, M.; Lyaskovskyy, V.; Domingos, S. R.; Buma, W. J.; Woutersen, S.; Troeppner, O.; Ivanović-Burmazović, I.; Lu, H.; Cui, X.; Zhang, X. P.; et al. J. Am. Chem. Soc. 2015, 137 (16), 5468–5479.

(42) Solis, B. H.; Maher, A. G.; Dogutan, D. K.; Nocera, D. G.; Hammes-Schiffer, S. Proc. Natl. Acad. Sci.

2016, 113 (3), 485–492.

(43) Vogel, A.; Dechert, S.; Brückner, C.; Meyer, F. Inorg. Chem. 2017, 56 (4), 2221–2232.

(44) Maher, A. G.; Liu, M.; Nocera, D. G. Inorg. Chem. 2019, 58 (12).

(45) Wong, J. L.; Sánchez, R. H.; Logan, J. G.; Zarkesh, R. A.; Ziller, J. W.; Heyduk, A. F. Chem. Sci. 2013, 4

(4), 1906–1910.

(46) Wong, J. L.; Higgins, R. F.; Bhowmick, I.; Cao, D. X.; Szigethy, G.; Ziller, J. W.; Shores, M. P.; Heyduk, A. F. Chem. Sci. 2016, 7 (2), 1594–1599.

(47) Dauth, A.; Love, J. A. Dalton Trans. 2012, 41 (26), 7782–7791.

(48) Munhá, R. F.; Zarkesh, R. A.; Heyduk, A. F. Inorg. Chem. 2013, 52 (19), 11244–11255.

(49) Wojnar, M. K.; Ziller, J. W.; Heyduk, A. F. Eur. J. Inorg. Chem. 2017, 2017 (47), 5571–5575.

(50) Rosenkoetter, K. E.; Wojnar, M. K.; Charette, B. J.; Ziller, J. W.; Heyduk, A. F. Inorg. Chem. 2018, 57 (16),

9728–9737.

(51) Hollas, A. M.; Ziller, J. W.; Heyduk, A. F. Polyhedron 2018, 143, 111–117.

(52) Myers, T. W.; Holmes, A. L.; Berben, L. A. Inorg. Chem. 2012, 51 (16), 8997–9004.

(53) Berben, L. A. Chem. - A Eur. J. 2015, 21 (7), 2734–2742.

(54) Sherbow, T. J.; Fettinger, J. C.; Berben, L. A. Inorg. Chem. 2017, 56 (15), 8651–8660.

(55) Goldsmith, C. R. Coord. Chem. Rev. 2018, 377, 209–224.

(56) Noss, M. E.; Wieder, N. L.; Carroll, P. J.; Zdilla, M. J.; Berry, D. H. Inorg. Chem. 2018, 57 (12), 7036–

7043.

(57) Römelt, C.; Weyhermüller, T.; Wieghardt, K. Coord. Chem. Rev. 2019, 380, 287–317.

(58) Schmidt, V. A.; Hoyt, J. M.; Margulieux, G. W.; Chirik, P. J. J. Am. Chem. Soc. 2015, 137 (24), 7903–7914.

(19)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 30PDF page: 30PDF page: 30PDF page: 30 18

138 (10), 3314–3324.

(60) Myers, T. W.; Kazem, N.; Stoll, S.; Britt, R. D.; Shanmugam, M.; Berben, L. A. J. Am. Chem. Soc. 2011,

133 (22), 8662–8672.

(61) Cates, C. D.; Myers, T. W.; Berben, L. A. Inorg. Chem. 2012, 51 (21), 11891–11897.

(62) Blackmore, K. J.; Lal, N.; Ziller, J. W.; Heyduk, A. F. J. Am. Chem. Soc. 2008, 130 (9), 2728–2729.

(63) Moxey, G. J.; Ortu, F.; Goldney Sidley, L.; Strandberg, H. N.; Blake, A. J.; Lewis, W.; Kays, D. L. Dalton

Trans. 2014, 43 (12), 4838–4846.

(64) Bart, S. C.; Lobkovsky, E.; Chirik, P. J. J. Am. Chem. Soc. 2004, 126 (42), 13794–13807.

(65) Bouwkamp, M. W.; Lobkovsky, E.; Chirik, P. J. Inorg. Chem. 2006, 45 (1), 2–4.

(66) Russell, S. K.; Lobkovsky, E.; Chirik, P. J. J. Am. Chem. Soc. 2011, 133 (23), 8858–8861.

(67) Hoyt, J. M.; Sylvester, K. T.; Semproni, S. P.; Chirik, P. J. J. Am. Chem. Soc. 2013, 135 (12), 4862–4877.

(68) Monfette, S.; Turner, Z. R.; Semproni, S. P.; Chirik, P. J. J. Am. Chem. Soc. 2012, 134 (10), 4561–4564.

(69) Pony Yu, R.; Hesk, D.; Rivera, N.; Pelczer, I.; Chirik, P. J. Nature 2016, 529 (7585), 195–199.

(70) Power, P. P. Nature 2010, 463 (7278), 171–177.

(71) Hopkinson, M. N.; Richter, C.; Schedler, M.; Glorius, F. Nature 2014, 510 (7506), 485–496.

(72) Frey, G. D.; Lavallo, V.; Donnadieu, B.; Schoeller, W. W.; Bertrand, G. Science (80-. ). 2007, 316 (5823),

439–441.

(73) Hicks, J.; Vasko, P.; Goicoechea, J. M.; Aldridge, S. Nature 2018, 557 (7703), 92–95.

(74) Liu, Y.; Li, J.; Ma, X.; Yang, Z.; Roesky, H. W. Coord. Chem. Rev. 2018, 374, 387–415.

(75) Martin, D.; Soleilhavoup, M.; Bertrand, G. Chem. Sci. 2011, 2 (3), 389–399.

(76) Asay, M.; Jones, C.; Driess, M. Chem. Rev. 2011, 111 (2), 354–396.

(77) Chu, T.; Nikonov, G. I. Chem. Rev. 2018, 118 (7), 3608–3680.

(78) Protchenko, A. V.; Bates, J. I.; Saleh, L. M. A.; Blake, M. P.; Schwarz, A. D.; Kolychev, E. L.; Thompson, A. L.; Jones, C.; Mountford, P.; Aldridge, S. J. Am. Chem. Soc. 2016, 138 (13), 4555–4564.

(79) Wendel, D.; Porzelt, A.; Herz, F. A. D.; Sarkar, D.; Jandl, C.; Inoue, S.; Rieger, B. J. Am. Chem. Soc. 2017,

139 (24), 8134–8137.

(80) Zhao, W.; McCarthy, S. M.; Lai, T. Y.; Yennawar, H. P.; Radosevich, A. T. J. Am. Chem. Soc. 2014, 136

(50), 17634–17644.

(81) Cui, C.; Roesky, H. W.; Schmidt, H.-G.; Noltemeyer, M.; Hao, H.; Cimpoesu, F. Angew. Chemie Int. Ed.

2000, 39 (23), 4274–4276.

(82) Chu, T.; Korobkov, I.; Nikonov, G. I. J. Am. Chem. Soc. 2014, 136 (25), 9195–9202.

(83) Hicks, J.; Vasko, P.; Goicoechea, J. M.; Aldridge, S. J. Am. Chem. Soc. 2019, jacs.9b05925.

(84) Power, P. P. Nature 2010, 463 (7278), 171–177.

(85) Stephan, D. W. Acc. Chem. Res. 2015, 48 (2), 306–316.

(86) Légaré, M. A.; Pranckevicius, C.; Braunschweig, H. Chem. Rev. 2019.

(87) Stephan, D. W.; Erker, G. Angew. Chem., Int. Ed. 2015, 54 (22), 6400–6441.

(88) Stephan, D. W. J. Am. Chem. Soc. 2015, 137 (32), 10018–10032.

(89) Bayne, J. M.; Stephan, D. W. Chem. Soc. Rev. 2016, 45 (4), 765–774.

(90) Melen, R. L. Chem. Soc. Rev. 2016, 45 (4), 775–788.

(91) Ma, Y.; Zhang, S.; Chang, C. R.; Huang, Z. Q.; Ho, J. C.; Qu, Y. Chem. Soc. Rev. 2018, 47 (15), 5541–

(20)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 31PDF page: 31PDF page: 31PDF page: 31 19

(92) Melen, R. L. Science (80-. ). 2019, 363 (6426), 479–484.

(93) Liu, L. L.; Stephan, D. W. Chem. Soc. Rev. 2019, 48 (13), 3454–3463.

(94) Lam, J.; Szkop, K. M.; Mosaferi, E.; Stephan, D. W. Chem. Soc. Rev. 2019, 48 (13), 3592–3612.

(95) Bourget-Merle, L.; Lappert, M. F.; Severn, J. R. Chem. Rev. 2002, 102 (9), 3031–3066.

(96) Asay, M.; Jones, C.; Driess, M. Chem. Rev. 2011, 111 (2), 354–396.

(97) Chu, T.; Nikonov, G. I. Chem. Rev. 2018, 118 (7), 3608–3680.

(98) Tsai, Y. C. Coord. Chem. Rev. 2012, 256 (5–8), 722–758.

(99) Mao, W.; Xiang, L.; Chen, Y. Coord. Chem. Rev. 2017, 346, 77–90.

(100) Liu, Y.; Li, J.; Ma, X.; Yang, Z.; Roesky, H. W. Coord. Chem. Rev. 2018, 374, 387–415.

(101) Chen, C.; Bellows, S. M.; Holland, P. L. Dalton Trans. 2015, 44 (38), 16654–16670.

(102) Hohloch, S.; Kriegel, B. M.; Bergman, R. G.; Arnold, J. Dalton Trans. 2016, 45 (40), 15725–15745.

(103) Webster, R. L. Dalton Trans. 2017, 46 (14), 4483–4498.

(104) Camp, C.; Arnold, J. Dalton Trans. 2016, 45 (37), 14462–14498.

(105) Bamberger, E.; Wheelwright, E. Ber. Dtsch. Chem. Ges. 1892, 25 (2), 3201–3213.

(106) Ullmann’s Encyclopedia of Industrial Chemistry; VCH Verlag GmbH & Co. KGaA, Ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2000. .

(107) Friese, P. Ber. Dtsch. Chem. Ges. 1875, 8 (2), 1078–1080.

(108) Martin, A.; Morcillo, N.; Lemus, D.; Montoro, E.; Da Silva Telles, M. A.; Simboli, N.; Pontino, M.; Porras, T.; León, C.; Velasco, M.; et al. Int. J. Tuberc. Lung Dis. 2005, 9 (8), 901–906.

(109) Moodley, S.; Koorbanally, N. A.; Moodley, T.; Ramjugernath, D.; Pillay, M. J. Microbiol. Methods 2014,

104, 72–78.

(110) Kuhn, R.; Trischmann, H. Angew. Chemie 1963, 75 (6), 294–295.

(111) Chung, G.; Lee, D. Chem. Phys. Lett. 2001, 350 (3–4), 339–344.

(112) Ciofini, I.; Daul, C. A. Coord. Chem. Rev. 2003, 238–239, 187–209.

(113) Koivisto, B. D.; Hicks, R. G. Coord. Chem. Rev. 2005, 249 (23), 2612–2630.

(114) Spin Labeling; Berliner, L. J., Reuben, J., Eds.; Biological Magnetic Resonance; Springer US: Boston, MA, 1989; Vol. 8.

(115) Polymerization, V. 1969, 124 (3000), 211–221.

(116) Chen, E. K. Y.; Teertstra, S. J.; Chan-Seng, D.; Otieno, P. O.; Hicks, R. G.; Georges, M. K. Macromolecules

2007, 40 (24), 8609–8616.

(117) Chang, M.-C.; Roewen, P.; Travieso-Puente, R.; Lutz, M.; Otten, E. Inorg. Chem. 2015, 54 (1), 379–388.

(118) Travieso-Puente, R.; Chang, M.-C.; Otten, E. Dalton Trans. 2014, 43, 18035–18041.

(119) Gilroy, J. B.; Ferguson, M. J.; McDonald, R.; Patrick, B. O.; Hicks, R. G. Chem. Commun. 2007, 0 (2),

126–128.

(120) Barbon, S. M.; Price, J. T.; Reinkeluers, P. A.; Gilroy, J. B. Inorg. Chem. 2014, 53 (19), 10585–10593.

(121) Barbon, S. M.; Reinkeluers, P. A.; Price, J. T.; Staroverov, V. N.; Gilroy, J. B. Chem. - A Eur. J. 2014, 20

(36), 11340–11344.

(122) Maar, R. R.; Zhang, R.; Stephens, D. G.; Ding, Z.; Gilroy, J. B. Angew. Chem., Int. Ed. 2019, 58 (4), 1052–

1056.

(123) Van Belois, A.; Maar, R. R.; Workentin, M. S.; Gilroy, J. B. Inorg. Chem. 2019, 58 (1), 834–843.

(21)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 32PDF page: 32PDF page: 32PDF page: 32 20

(125) Maar, R. R.; Barbon, S. M.; Sharma, N.; Groom, H.; Luyt, L. G.; Gilroy, J. B. Chem. - A Eur. J. 2015, 21

(44), 15589–15599.

(126) Hesari, M.; Barbon, S. M.; Staroverov, V. N.; Ding, Z.; Gilroy, J. B. Chem. Commun. 2015, 51 (18), 3766–

3769.

(127) Barbon, S. M.; Price, J. T.; Yogarajah, U.; Gilroy, J. B. RSC Adv. 2015, 5 (69), 56316–56324.

(128) Barbon, S. M.; Novoa, S.; Bender, D.; Groom, H.; Luyt, L. G.; Gilroy, J. B. Org. Chem. Front. 2017, 4 (2),

178–190.

(129) Barbon, S. M.; Buddingh, J. V.; Maar, R. R.; Gilroy, J. B. Inorg. Chem. 2017, 56 (19), 12003–12011.

(130) Novoa, S.; Gilroy, J. B. Polym. Chem. 2017, 8 (35), 5388–5395.

(131) Dhindsa, J. S.; Maar, R. R.; Barbon, S. M.; Olivia Avilés, M.; Powell, Z. K.; Lagugné-Labarthet, F.; Gilroy, J. B. Chem. Commun. 2018, 54 (50), 6899–6902.

(132) Maar, R. R.; Rabiee Kenaree, A.; Zhang, R.; Tao, Y.; Katzman, B. D.; Staroverov, V. N.; Ding, Z.; Gilroy, J. B. Inorg. Chem. 2017, 56 (20), 12436–12447.

(133) Maar, R. R.; Catingan, S. D.; Staroverov, V. N.; Gilroy, J. B. Angew. Chemie Int. Ed. 2018, 57 (31), 9870–

9874.

(134) Hesari, M.; Barbon, S. M.; Mendes, R. B.; Staroverov, V. N.; Ding, Z.; Gilroy, J. B. J. Phys. Chem. C 2018,

122 (2), 1258–1266.

(135) Novoa, S.; Paquette, J. A.; Barbon, S. M.; Maar, R. R.; Gilroy, J. B. J. Mater. Chem. C 2016, 4 (18), 3987–

3994.

(136) Barbon, S. M.; Staroverov, V. N.; Gilroy, J. B. Angew. Chemie Int. Ed. 2017, 129 (28), 8285–8289.

(137) Chang, M. C.; Dann, T.; Day, D. P.; Lutz, M.; Wildgoose, G. G.; Otten, E. Angew. Chem., Int. Ed. 2014,

53 (16), 4118–4122.

(138) Chang, M.-C.; Otten, E. Chem. Commun. 2014, 50 (56), 7431–7433.

(139) Chang, M. C.; Otten, E. Inorg. Chem. 2015, 54 (17), 8656–8664.

(140) Chang, M. C.; Chantzis, A.; Jacquemin, D.; Otten, E. Dalton Trans. 2016, 45 (23), 9477–9484.

(141) Travieso-Puente, R.; Broekman, J. O. P.; Chang, M. C.; Demeshko, S.; Meyer, F.; Otten, E. J. Am. Chem.

Soc. 2016, 138 (17), 5503–5506.

(142) Kamphuis, A. J.; Milocco, F.; Koiter, L.; Pescarmona, P.; Otten, E. ChemSusChem 2019, No. Ii, 1–8.

(143) Milocco, F.; Demeshko, S.; Meyer, F.; Otten, E. Dalton Trans. 2018, 47 (26).

(144) Milocco, F.; De Vries, F.; Dall’Anese, A.; Rosar, V.; Zangrando, E.; Otten, E.; Milani, B. Dalton Trans.

2018, 47 (41), 14445–14451.

(145) Broere, D. L. J.; Mercado, B. Q.; Lukens, J. T.; Vilbert, A. C.; Banerjee, G.; Lant, H. M. C.; Lee, S. H.; Bill, E.; Sproules, S.; Lancaster, K. M.; et al. Chem. - A Eur. J. 2018, 24 (37), 9417–9425.

(146) Broere, D. L. J.; Mercado, B. Q.; Bill, E.; Lancaster, K. M.; Sproules, S.; Holland, P. L. Inorg. Chem. 2018,

57, acs.inorgchem.8b00226.

(147) Schorn, W.; Grosse-Hagenbrock, D.; Oelkers, B.; Sundermeyer, J. Dalton Trans. 2016, 45 (3), 1201–1207.

(148) Mandal, A.; Schwederski, B.; Fiedler, J.; Kaim, W.; Lahiri, G. K. Inorg. Chem. 2015, 54 (16), 8126–8135.

(149) Protasenko, N. A.; Poddel’sky, A. I.; Bogomyakov, A. S.; Fukin, G. K.; Cherkasov, V. K. Inorg. Chem.

2015, 54 (13), 6078–6080.

(150) Rezinskikh, Z. G.; Pervova, I. G.; Lipunova, G. N.; Maslakova, T. I.; Gorbatenko, Y. A.; Lipunov, I. N.; Sigeikin, G. I. Russ. J. Coord. Chem. 2008, 34 (9), 659–663.

(22)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

Processed on: 21-11-2019 PDF page: 33PDF page: 33PDF page: 33PDF page: 33 21

(151) Zaidman, A. V.; Pervova, I. G.; Vilms, A. I.; Belov, G. P.; Kayumov, R. R.; Slepukhin, P. A.; Lipunov, I. N.

Inorganica Chim. Acta 2011, 367 (1), 29–34.

(152) Zaidman, A. V.; Khasbiullin, I. I.; Belov, G. P.; Pervova, I. G.; Lipunov, I. N. Pet. Chem. 2012, 52 (1), 28–

34.

(153) Siedle, A. R.; Pignolet, L. H. Inorg. Chem. 1980, 19 (7), 2052–2056.

(154) Kabir, E.; Wu, C. H.; Wu, J. I. C.; Teets, T. S. Inorg. Chem. 2016, 55 (2), 956–963.

(155) Mu, G.; Cong, L.; Wen, Z.; Wu, J. I. C.; Kadish, K. M.; Teets, T. S. Inorg. Chem. 2018, 57 (15), 9468–

9477.

(156) Kabir, E.; Patel, D.; Clark, K.; Teets, T. S. Inorg. Chem. 2018, 57 (17), 10906–10917.

(157) Hong, S.; Hiii, L. M. R.; Gupta, A. K.; Naab, B. D.; Gllroy, J. B.; Hicks, R. G.; Cramer, C. J.; Tolman, W. B. Inorg. Chem. 2009, 48 (10), 4514–4523.

(158) Sigeikin, G. I.; Lipunova, G. N.; Pervova, I. G. Russ. Chem. Rev. 2006, 75 (10), 885–900.

(159) Gorbatenko, Y. A.; Pervova, I. G.; Lipunova, G. N.; Maslakova, T. I.; Lipunov, I. N.; Sigeikin, G. I. Russ.

J. Appl. Chem. 2005, 78 (6), 936–939.

(160) Pavlova, I. S.; Pervova, I. G.; Belov, G. P.; Khasbiullin, I. I.; Slepukhin, P. A. Pet. Chem. 2013, 53 (2),

(23)
(24)

537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol 537515-L-sub01-bw-Mondol Processed on: 21-11-2019 Processed on: 21-11-2019 Processed on: 21-11-2019

(25)

Referenties

GERELATEERDE DOCUMENTEN

Aluminum complexes with redox-active formazanate ligand: Synthesis, characterization, and reduction chemistry

We attribute this difference to the fact that in 4a, resonance delocalization from the electron-rich, reduced formazanate backbone occurs into two aromatic groups (N-Ph), whereas

In general, we anticipate that the ability to influence basicity, radical stability and N-H/N-C bond strength of compounds containing the formazanate ligand, either via

The synthesis and characterization data for aluminum complexes with one or two redox-active formazanate ligands reported in this paper provides an entry to studying the reactivity

spectroscopy, X-ray crystallography) and computational studies (Wiberg bond indices, NBO analysis) reveal that the increased ionic character in the Al compounds results in a

detailed analysis of this chemical exchange process by variable temperature NMR experiment and NMR lineshape analysis, which reveals that nitrogen inversion is responsible for the

purple and β IBO golden). b) Changes of IBO of O-centered unpaired electron of TEMPO along IRC (blue) which reveals the participation of this unpaired electron in the formation

This, intrinsically semiconducting material with 1.1 eV bandgap in its monolayer limit makes it a promising candidate for 2D memory type devices [51] and enables a versatile