• 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!
31
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: 117PDF page: 117PDF page: 117PDF page: 117

Chapter 5

Structure and bonding in reduced boron

and aluminium complexes with

formazanate ligands

Group 13 complexes of the type [(PhNNC(p-tol)NNPh)ZPh2]2- (Z = B, Al) containing a highly

reduced, trianionic formazanate-derived ligand were studied and the differences in structure, bonding and reactivity between the B and Al compounds was investigated. The increased ionic character in the bonding of the Al complex is evident from enhanced charge delocalization onto the peripheral ligand substituents (N-Ph) via the π-framework, as shown by the rotation barrier around the N-C(Ph) bond. The electron-rich nature of these compounds allows facile benzylation at the ligand, and the structures of the products were analysed by X-ray crystallography. The products are inorganic analogues of 1-alkylated 1,2,3,4-tetrahydro-1,2,4,5-tetrazines (‘leucoverdazyls’). The six-membered heterocyclic cores of the B and Al compounds are shown to be different, having envelope- and boat-type conformations, respectively. Homolysis of the N-C(benzyl) bond in these compounds was studied by NMR spectroscopy under conditions that trap the organic radical as TEMPO-Bn. Analysis of the reaction kinetics affords activation parameters that approximate the N-C(benzyl) bond strength. The ionic Al compound has one of the weakest N-C bonds reported so far in this type of inorganic leucoverdazyl analogues.

This chapter has been published:

Ranajit Mondol and Edwin Otten* “Structure and bonding in reduced boron and aluminium complexes with formazanate ligands” Dalton Trans., 2019, 48, 13981-13988 (DOI:

(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: 118PDF page: 118PDF page: 118PDF page: 118

104

5.1 Introduction

The direct involvement of the ligands in redox-reactions by coordination complexes bearing non-innocent ligands is an emerging research area.1–3 In this context, there is a growing interest

in designing new type of redox-active (non-innocent) ligands. In 2007, Hicks and coworkers observed reversible ligand-based reductions in a formazanate boron acetate compound.4 This

demonstrated for the first time that the formazanate ligand can be considered a redox-active analogue of the well-known β-diketiminate ligands.5–8 Taking inspiration from this work, our

group has explored the coordination chemistry, redox-behavior and reactivity of complexes with formazanate ligands.9–16 Along with our work, the Gilroy group17–26 and others27–29 have

been involved in the synthesis and application of the new molecular complexes with formazanate ligands. Previously, we showed that the ligands in boron and aluminum compounds with formazanate ligands could be sequentially reduced by 1- and 2-electron at moderate reduction potentials (Scheme 5.1).30,31 The 2-electron reduced formazanate boron

compound (12-) subsequently reacted with electrophiles (E+) such as benzyl bromide (BnBr)

and water (H2O) to form ligand-benzylated and -protonated products (Bn1- and H1- in Scheme

5.1).32 In these compounds, the formazanate ligands is modified by ‘storage’ of [2e-/E+], which

could be converted to Bn• and H radicals by the homolytic cleavage of the N-C(Bn) and N-H

bonds, respectively (Scheme 5.1).32 These reactions occur readily, because the boron-containing

radical that is generated (1-•) is

N N N N Ar Ph p-tol Z Ph Ph N N N N Ar Ph p-tol Z Ph Ph N N N N Ar Ph p-tol Z Ph Ph 1e -1e -2 (1/1Mes/2) 2-(1/1Mes/2) N N N N Ph Ph p-tol B Ph Ph E [Bn1][Na] (E = CH2Ph) [H1][Na](E = H) E-X = PhCH2-Br H-OH E-X, THF -NaX TEMPO . TEMPO-E Z = B; Ar = Ph, 1 Z = B; Ar = Mes, 1Mes Z = Al; Ar = Ph, 2 ligand-based strorage of 2e -ligand-based strorage of 2e-/E+

Scheme 5.1 Ligand-based storage of [2e-/E+], and subsequent conversion to E (Bn/H) radicals

(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: 119PDF page: 119PDF page: 119PDF page: 119

105 relatively stable due the presence of a low-energy SOMO that is delocalized over all four N-atoms in the ligand backbone.30,32 It is anticipated that the basicity (nucleophilicity), radical

stability and N-C/N-H bond strength of compounds bearing functionalized formazanate ligands could be tuned either via ligand substituent effects or by incorporation of a different central element (main group or transition metal) in the formazanate chelate ring.

In order to investigate the effects of changing the central element (boron) in 1 to the more

electropositive aluminum, here we provide a detailed comparison of formazanate B and Al complexes with an identical ligand. The comparison includes an analysis of resonance delocalization in the two-electron reduced formazanate aluminum diphenyl compound (22-) via

dynamic NMR spectroscopy. The synthesis of ligand-benzylated products Bn2- is described, and

crystallographic and spectroscopic characterization data is provided. Furthermore, the kinetics of homolytic N-C(benzyl) cleavage is studied.

5.2 The hindered rotation around the N-C(Ph) bonds in two-electron reduced

formazanate aluminium diphenyl compound (2

2-

)

The two-electron reduced formazanate aluminium diphenyl compound [PhNNC(p-tol)NNPh]AlPh22- (22-) was synthesized as its disodium salt according to the previously

published procedure.31 The product 22-, which has an electron-rich, formally trianionic

formazanate ligand, is highly air-sensitive, but stable at room temperature under inert conditions. The 1H NMR spectrum of 22- in THF-d8 was shown to be temperature-dependent: at 233 K, the

spectrum contains 5 inequivalent resonances due to the N-Ph groups. Increasing the temperature results in line broadening and ultimately coalescence into three distinct signals as expected for the ortho, meta and para-positions of a Ph group (Figure 5.1). These features are indicative of

hindered rotation around the N-C(Ph) bond, which leads to inequivalent chemical environments for the two ortho- and meta-H positions, with exchange rates that are on the order of the NMR timescale.33,34 Lineshape analysis was carried out for the pairs of exchanging resonances in the

temperature range 233-303 K, which gave the activation parameters for the exchange process in 22- as ΔH = 54.1 ± 1.7 kJ/mol and ΔS = -2.5 ± 6.0 J/mol/K (see SI for details). The activation

enthalpy reflects a substantial amount of N-C(Ph) π-bonding character being broken upon going to the transition state, whereas there is little difference in entropy. A comparison with the related boron-containing compounds (12-) shows that the values are similar to those in the asymmetric

derivative [MesNNC(p-tol)NNPh]BPh22- ([1Mes]2-; ∆H‡ = 57.4 ± 1.8 kJ·mol-1 and ∆S‡ = 1 ±

6 J·mol-1·K-1).30 On the other hand, the B-analogue 12-, which has the same (symmetrical)

(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: 120PDF page: 120PDF page: 120PDF page: 120

106

timescale:§ no line-broadening is observed down to 233 K.30 We interpret the difference

between the B and Al complexes with an identical formazanate ligand (12- and 22-) as a

reflection of the difference in bonding within the boron and aluminium heterocycles. In particular, the largely ionic character in the Al-N bonds in comparison to the more covalent B-N interaction is responsible for increased accumulation of negative charge within the B-NB-NCB-NB-N framework in the Al compound. The increase in N-C(Ph) π-bonding due to resonance delocalization of the negative charge into the N-Ph group is also observed in the solid state structures obtained by X-ray crystallography: the N-C(Ph) bonds in 22- are 1.371(2)/1.375(3)

Å,31 whereas those in 12- are marginally larger at 1.379(3)/1.385(3) Å.30

Figure 5.1 1H NMR spectra of 22- in THF-d8 at various temperatures.

To further probe the bonding differences between formazanate boron and aluminium complexes, a computational study was carried out at the DFT level (B3LYP functional and 6-311+G(d,p) basis set). The geometries were optimized in the gas phase starting from the crystallographic coordinates, with the Na(THF)x cations removed (the computational results for 12-calc were

described previously).30 The metrical parameters of the optimized structures of 12-calc and 22-calc

are in good agreement with the experimental structures, albeit that the N-N bonds are somewhat shorter in the DFT models (see Table S2). Visual inspection of the frontier orbitals shows that

(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: 121PDF page: 121PDF page: 121PDF page: 121

107 the HOMO is primarily π-antibonding between the N-atoms in the formazanate ligand, and in addition, evidences the presence of π-bonding character in the N-C(Ph) fragment (Figure S5.12). Although the differences are small, the calculated Wiberg bond index35 for the N-C(Ph) bonds

is larger in the Al compound (22-calc: 1.24) than in the B compound (12-calc: 1.22), corroborating the trend for the strength of π-bonding that was obtained from the NMR study. In addition, a higher Wiberg bond index is found for the B-N bonds in 12- (0.68) in comparison to the Al-N

bonds in 22- (0.40) and the natural charges indicate that the formazanate ligand bears a

significantly higher negative charge in the Al complex (-2.40 e) than in the B analogue (-1.99 e).

Computational evaluation of the barrier to rotation around the N-C(Ph) bond in 22-calc was

carried out by scanning the dihedral angle between the Ph ring and the ligand backbone. As expected, a maximum was found at a dihedral angle of ca. 90°, and at this geometry a transition state optimization was carried out to arrive at a saddle point (22-TS, Nimag = 1). The computed

barrier for N-C(Ph) bond rotation was found to be 55.9 kJ·mol-1, which is in good agreement

with the experimental value (~ 54.9 kJ·mol-1 at 298 K). A comparison between the ground and

transition state shows that π-bonding is lost in the N-Ph ring that is rotated (Wiberg bond index = 1.04), whereas a small increase is observed in the other N-C(Ph) bond (1.26). This is also reflected in the different N-C(Ph) bond lengths that are calculated for 22-TS (1.411 Å and 1.365 Å; cf. 1.370 Å in 22-calc).

5.3 Reactivity study of two-electron reduced formazanate aluminium

diphenyl compound (2

2-

) with benzyl bromide

5.3.1 Synthesis of ligand-benzylated product

Bn

2

-

, and its characterization

by NMR spectroscopy

We subsequently evaluated the reactivity of compound 22- towards the electrophile benzyl

bromide. Treatment of an orange THF-d8 solution of 22- with 1 equivalent of BnBr resulted in

immediate colour change from orange to yellowish-green. The 1H NMR spectrum of the

reaction mixture at room temperature (400MHz, THF-d8) reveals diagnostic resonances for the

diastereotopic protonsof the benzyl-CH2 group at δ 4.29 and 4.45 ppm with a geminal coupling

constant of 2JHH = 12.5 Hz. Also, the 1H NMR spectrum shows two inequivalent resonances for

the para-protons of the N-Ph rings at δ 6.09 and 6.34 ppm due to a descent in symmetry from the C2v-symmetric precursor 22-. The NMR data of the reaction product are similar to that of

(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: 122PDF page: 122PDF page: 122PDF page: 122

108

in the ligand backbone. Consequently, we formulate the product as the anionic, ligand-benzylated complex [PhNN(Bz)C(p-tol)NNPh]AlPh2- (Bn2-, Scheme 5.2). On a preparative

scale, the reaction between 22- and BnBr allowed isolation of the sodium salt [Bn2][Na]as a

solid material in 52 % yield.

N N N N Ph Ph p-tol Al Ph Ph 2 2-RT, THF PhCH2Br N N N N Ph Ph p-tol Al Ph Ph [Bn2][Na] - NaBr Na(THF)3 (THF)3Na Ph (THF)3Na RT, THF Bu4NBr N N N N Ph Ph p-tol Z Ph Ph [Bn1][NBu4] [Bn2][NBu4] - NaBr Ph N N N N Ph Ph p-tol Z Ph Ph Ph (THF)3Na Bu4N Z = B, [Bn1][Na] Z = Al, [Bn2][Na]

Scheme 5.2 Synthesis of ligand-benzylated products [Bn2][Na], [Bn2][NBu4] and [Bn1][NBu4]

5.3.2 Crystallographic characterization data of ligand-benzylated products

Bn

1

-

and

Bn

2

-

Crystals of [Bn2][Na] were obtained by slow diffusion of hexane into the THF solution at -30 °C.

Although the previously reported boron analogue [Bn1][Na] also crystallizes under these

conditions, these crystals quickly melt at room temperature, thus thwarting a structure determination.32 Treatment of both these sodium salts with Bu4NBr resulted in cation exchange

and formation of the tetrabutyl ammonium salts [Bn1][NBu4] and [Bn2][NBu4] in yields of ca.

75% (Scheme 5.2). Gratifyingly, crystals of [Bn1][NBu4] had a much higher melting point, and

X-ray structure determinations were successfully completed for both B- and Al-containing products. The molecular structures of [Bn1][NBu4] and [Bn2][Na] shows that in both of these

compounds, the benzyl group is attached to one of the internal N-atoms to retain the 6-membered chelate ring structures of the dianionic precursors (Figure 5.2). A closer inspection of the central six-membered heterocyclic rings in these compounds reveals some noteworthy differences. Compound [Bn1][NBu4] adopts an envelope conformation that is similar to the

precursor 12-,30 with the formazanate backbone atoms (NNCNN) nearly coplanar and the B

atom residing 0.568 Å above that plane. In contrast, the six-membered ring in the Al compound [Bn2][Na] is found to be present in a boat conformation in which the N1, Al1, N3 and C7 are

coplanar, whereas and N2 and N4-atoms are displaced out from that plane by 0.502 and 0.507 Å, respectively (Figure 5.2). This is markedly different from the nearly planar conformation in the precursor 22-.31 A Cremer-Pople puckering analysis36 was carried out using PLATON,37

which allowed a more quantitative distinction between the divergent geometrical parameters within the core of [Bn1][NBu4] vs. [Bn2][Na]. In this analysis, a combination of the puckering

(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: 123PDF page: 123PDF page: 123PDF page: 123

109 sphere. The following puckering parameters are obtained for these compounds: for [Bn1][NBu4]

(Q = 0.44 Å; θ = 57.0°; φ = 332.2°) and [Bn2][Na] (Q = 0.58 Å; θ = 89.8°; φ = 300.2°). Of these

parameters, the θ angles in six-membered rings distinguish between chair, envelope and boat conformations, whereas φ defines pseudorotation pathways that transverse twisted conformations as well. The angles θ found for both compounds are indicative of either an envelope (idealized: 54.7°; [Bn1][NBu4]: 57.0°) or boat conformation (idealized: 90°; [Bn2][Na]:

89.8°).

Figure 5.2 (a) Molecular structures of [Bn2][Na] (top, left) and [Bn1][NBu4] (top, right), and

showing 50% probability ellipsoids. Hydrogen atoms for [Bn2][Na] and [Bn1][NBu4] are

omitted for clarity. THF molecules (except for the O atoms bonded to Na) for [Bn2][Na] are

omitted for clarity. (b) Showing six-membered chelate rings for [Bn2][Na] (bottom, left) and for

[Bn1][NBu4] (bottom, right).

In the B and Al dianions (12- and 22-) the bonding in the six-membered core is delocalized, as

shown by equivalent N-N and C-N bond lengths, whereas ligand benzylation leads to a more localized bonding picture. For example, the C-N bonds in the ligand backbone change from 1.332(2)/1.328(2) Å in 22- to 1.437(1)/1.292(1) Å in [Bn2][Na]. Similar changes in bond lengths

occur in the boron analogue (see Tables S5.1/S5.2 for a full comparison of metrical data). In both compounds, the benzylated N-atom is pyramidal as indicated by the sum of angles around N(2) being much smaller than 360° (∑∠(N(2)) = 348.1° and 334.7° for[Bn1][ NBu4] and

distorted boat conformation

(a)

(b)

(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: 124PDF page: 124PDF page: 124PDF page: 124

110

[Bn2][Na], respectively). In [Bn2][Na] the Ph-substituted N-atoms are planar (sp2-hybridized)

with ∑∠(N) = 359.5° for N(1) and 354.5° for N(4). Conversely, in the boron analogue [Bn1][NBu4] the N(4) atom is almost planar (∑∠(N) = 357.5°), while N(1) is pyramidalized

((∑∠(N) = 347.8°). It appears that these differences reflect a larger degree of directionality (covalency) in the B-N bonds in comparison to the more ionic Al congener.

Table 5.1 Selected bond lengths (Å) and bond angles (°) of [Bn1][NBu4] and [Bn2][Na]

[Bn1a][NBu4] [Bn2][Na] [Bn1a][NBu4] [Bn2][Na]

N1-N2 1.428(2) 1.446(1) N1-C1 1.414(2) 1.387(1) N3-N4 1.393(2) 1.402(1) N4-C15 1.392(2) 1.399(1) N2-C7 1.396(2) 1.437(1) N2-C33 1.468(2) 1.486(1) C7-N3 1.289(2) 1.292(1) N1-B1-N4 105.78(1) Al1-N1 1.867(1) N1-Al1-N4 95.24(4) Al1-N4 1.914(1) N1-N2-C7 115.41(1) 112.27(8) B1-N1 1.575(2) N1-N2-C33 113.94(1) 110.33(8) B1-N4 1.584(2) C7-N2-C33 118.79(1) 112.13(8)

5.3.3 The evaluation of ligand-benzylated products

Bn

1

-

and

Bn

2

-

by DFT

calculations

The anions Bn1- and Bn2- were also evaluated by DFT calculations. Optimized geometries of Bn1-calc and Bn2-calc were obtained starting from the crystallographic coordinates (counter cations

were removed) and shown to be in good agreement with the empirical structures. Importantly, the dissimilarities in pyramidalization at the N-atoms of the B and Al compounds is reproduced in the computational models, suggesting that these are intrinsic and not due to packing or ion pairing effects. In both the experimental structures and the DFT models, there is a small but noticeable elongation of the N-C(Bn) bond in the Al compound Bn2- (X-ray: 1.486(1) Å; DFT:

1.492 Å) compared to that in the B congener Bn1- (X-ray: 1.468(2) Å; DFT: 1.472 Å). An NBO

analysis indicated that the bonding within the NNCNN ligand backbone is similar in both B and Al compounds and can be described by a normal σ-bonding framework with an additional

(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: 125PDF page: 125PDF page: 125PDF page: 125

111

Figure 5.3 (A) Selected natural bond orbitals for the σ-framework in the ligand backbone (top);

the N-C π-bond and those around the N-benzyl atom (bottom). (B) Comparison of the B-N (covalent) vs Al-N (ionic) bonding interactions.

localized N=C π-bond (Figure 5.3A). The hybridization of the benzyl-substituted nitrogen atom is intermediate between sp2 and sp3, with a lone pair that has a small but non-negligible amount

of s-character (10% for Bn1-calc; 16% for Bn2-calc). Importantly, the NBO analysis shows the

N-B and N-Al interaction in both compounds to be quite different. For the boron compound Bn1-,

we find two natural bond orbitals that represent relatively covalent B-N σ-bonds (ca. 77% contribution from N; 23% from B). For the Al analogue Bn2-, on the other hand, no 2-center

Al-N bonds are obtained and instead the Al-NBO analysis indicates Al-N-based lone pairs that interact with an empty Al orbital (Figure 5.3). Similarly, the natural charges indicate the Al complex to be much more ionic (NPA charge for Al in Bn2-calc: 1.82; B in Bn1-: 0.74).

5.3.4 Characterization of ligand-benzylated product

Bn

2

-

by UV/Vis

spectroscopy

A comparison of the UV/Vis spectra of the Na+ and Bu4N+ salts of Bn1- and Bn2- shows that the

(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: 126PDF page: 126PDF page: 126PDF page: 126

112

compounds Bn1- and the aluminium compounds Bn2- absorb at 395 and 389 nm, respectively

(Figure S5.1). These absorbance maxima are most likely due to a π-π* transition of the localized N=C bonds present in these compounds, as shown by the crystallographic data (vide supra), and are blue-shifted compared to the delocalized precursors 12- and 22-max = 486 nm).†,31

5.4 Investigation of the cleavage of the N-C(Bn) bond in

Bn

2

-The compounds Bn1- and Bn2- areinorganic, anionic analogues of purely organic 1-alkylated

tetrahydro-1,2,4,5-tetrazines (‘leucoverdazyls’).38 Leucoverdazyls (with N-H bonds) and the

alkylated (N-C) derivatives are known to have weak N-H/N-C bonds because homolytic cleavage generates a stable verdazyl radical.38–41 The stability of this type of radicals extends to

inorganic systems.4,9,10,14,17,26,30,31,42 In order to examine the effect on the N-C(Bn) bond

dissociation energy (BDE) due to the replacement of B to the more electropositive Al in [Bn2][Na], the cleavage of the N-C bond in [Bn2][Na] was investigated. The N-C(Bn) bond

dissociation enthalpy in [Bn2][Na] was obtained experimentally by NMR spectroscopic

monitoring of the kinetics of benzyl radical transfer from [Bn2][Na] to TEMPO (present in

excess) in the temperature range of 65-85 °C (Scheme 5.3). Clean exponential decay of the starting material and concomitant appearance of TEMPO-Bn was observed, which allowed the rate constants to be determined (Figure S5.8). An Eyring analysis afforded the activation parameters △H‡ and △S of 107 ± 4 kJ∙mol-1 and 17 ± 11 J∙mol-1∙K-1, respectively, for the

N N N N Ph Ph p-tol Al Ph Ph Bn N N N N Ph Ph p-tol Al Ph Ph Bn ‡ TEMPO N N N N Ph Ph p-tol Al Ph Ph + N O Bn [Bn2-][Na] 2.- TEMPO-Bn

Scheme 5.3 Benzyl group transfer from [Bn2][Na] to TEMPO

Figure 5.4 Eyring analysis for benzyl transfer from [Bn2][Na] to TEMPO

∆H‡ = 107 ± 4

kJ∙mol-1

(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: 127PDF page: 127PDF page: 127PDF page: 127

113 benzyl transfer (Figure 5.4; see SI for details). Both these values are somewhat smaller than those found for the boron analogue [Bn1][Na] (△H = 121 ± 5 kJ∙mol-1; △S = 77 ± 14 J∙mol -1∙K-1).32 A likely explanation for these differences is the ground-state destabilization of the

N-C(Bn) bond that is indicated by the somewhat larger N-N-C(Bn) distance found by both experiment and theory (vide supra).

To investigate a possible influence of ion-pairing on the N-C(Bn) bond dissociation energy, the kinetics of benzyl transfer were also measured for the tetrabutyl ammonium salts [Bn1][NBu4]

and [Bn2][NBu4] using the same methodology. The resulting rate constants are in good

agreement with those of the sodium salts (see SI for details). Thus, even though in the solid state the sodium cation is bound to the ligand backbone, this weak interaction is likely broken in solution. This is further supported by the observation that the UV/Vis spectra in solution do not depend on the nature of the cation.

These data indicate that N-C(Bn) bond homolysis is modulated by the central element in these heterocyclic leucoverdazyl analogues: on going from relatively covalent, C-based parent structures (i.e., 1-benzyl-substituted 1,2,3,4-tetrahydro-1,2,4,5-tetrazines), the N-C(Bn) bond strength progressively decreases with an increase in the electropositive nature of the central element (i.e., C > B > Al).

5.5 Conclusions

In conclusion, this work addresses the differences in structure and bonding between dianionic formazanate boron (12-) and aluminum (22-) complexes. Experimental (NMR, UV/Vis

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 higher degree of resonance delocalization of the ligand negative charge into the periphery of the ligand (i.e., the N-Ph substituents), which is reflected in the rotation barrier around the N-C(Ph) bond. For both compounds, facile ligand benzylation occurs upon reaction with benzyl bromide to form

Bn1- and Bn2- as anionic analogues of carbon-based leucoverdazyls. X-ray diffraction studies of Bn2- and the boron congener Bn1- are reported and a comparison shows distinct differences in

the solid state structures between these complexes that can be related to a different degree of ionic character in the bonding. The kinetics of benzyl transfer show that the N-C(Bn) bond homolysis is modulated by the nature of central element present in the six-membered heterocyclic rings of Bn1- and Bn2-, with the ionic Al-based compound Bn2- having the weakest

(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: 128PDF page: 128PDF page: 128PDF page: 128

114

5.6 Notes

‡ The additional broadening observed above 283 K is attributed to the presence of a small amount of radical species (the paramagnetic monoanion 2-•) that engages in electron transfer

with 22-.

§ The estimated upper limit of the barrier for N-C(Ph) bond rotation is ca. 40 kJ·mol-1. See ESI

for details.

† The UV/Vis spectrum of 12- reported in the ESI of reference 30 is incorrect, likely due to

decomposition of this highly sensitive compound.

5.7 Experimental section

5.7.1 General considerations

All manipulations were carried out under nitrogen atmosphere using standard glovebox, Schlenk, and vacuum-line techniques. THF and hexane (Aldrich, anhydrous, 99.8%) were passed over columns of Al2O3 (Fluka). The compounds [Bn1][NBu4],[Bn2][Na]and [Bn2][NBu4]

are highly air-sensitive, and the solvents (THF and hexane) used for their preparation and characterization were additionally dried on Na/K alloy and subsequently vacuum transferred and stored under nitrogen. All solvents were degassed prior to use and stored under nitrogen. THF-d8 (Sigma-Aldrich) was vacuum transferred from Na/K alloy and stored under nitrogen.

The compounds [Bn1][Na]32 and 22- (as its disodium salt,

[(PhNNC(p-tol)NNPh)AlPh2][Na2(DME)4])31 were synthesized according to published procedures.NMR

spectra were recorded on Varian Mercury 400 or Inova 500 spectrometers. The 1H and 13C NMR

spectra were referenced internally using the residual solvent resonances and reported in ppm relative to TMS (0 ppm); J is reported in Hz. Assignments of NMR resonances was aided by COSY, NOESY, HSQC and HMBC experiments using standard pulse sequences. UV-Vis spectra were recorded in THF solution (~ 10-3 M) in a quartz cuvette using an AVANTES

AvaSpec-2048. Samples for elemental analyses were sent to Kolbe Microanalytical Laboratory (Mülheim an der Ruhr, Germany). However, despite our best efforts, no satisfactory analysis data could be obtained for these compounds, which is likely due to their highly air-sensitive nature.

(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: 129PDF page: 129PDF page: 129PDF page: 129

115

5.7.2 Compounds synthesis and characterization

Synthesis of [BnLBPh2][NBu4] ([Bn1][NBu4]). Compound [Bn1][Na] (75 mg, 0.092 mmol) was

dissolved in 2 ml of THF in a vial inside the glove box. To this was added 1 equiv of tetrabutyl ammonium bromide (Bu4NBr) and the mixture was stirred overnight. The NaBr that had

precipitated was allowed to settle and the supernatant was transferred to another vial. After concentration under vacuum, a layer of hexane (1 mL) was added on top of the THF solution and the layers were allowed to slowly diffuse at -30 °C. After two days, a crystalline material had precipitated, which was isolated and washed with hexane (3 x 2 mL). Subsequently, drying under vacuum gave compound [Bn1][NBu4] as a light green crystalline material (59 mg, 0.072

mmol, 78 %). 1H NMR (400 MHz, THF-d8,25°C) δ 7.83 (d, J = 8.1 Hz, 2H, p-tol o-H), 7.70

(d, J = 7.1 Hz, 2H, B-Ph(1) o-H), 7.11-7.06 (m, overlapped, 4H, B-Ph(1) (m)-H and p-tol m-H), 7.03-6.98 (m, overlapped, 7H, B-Ph(1) p-H, B-Ph(2) o-H, N(1)Ph o-H and (benzyl)Ph o-H), 6.86 – 6.79 (m, 3H, (benzyl)Ph (m+p)-H), 6.64 – 6.53 (m, 5H, overlapped, 5H, N(1)Ph m-H and B-Ph(2) (m+p)-H), 6.46 (d, J = 4.2 Hz, 4H, N(2)Ph (o+m)-H), 6.17-6.13 (m, 1H, N(2)Ph p-H), 6.09 (t, J = 7.1 Hz, 1H, N(1)Ph p-H), 3.78 (d, J = 15.3 Hz, 1H, benzyl-CH2), 3.67 – 3.59

(m, 4H, THF), 3.44 (d, J = 15.3 Hz, 1H, benzyl-CH2 ), 3.12 (s, 8H, NBu4), 2.29 (s, 3H, p-tol

CH3), 1.58 (s, 8H, NBu4), 1.34-1.30 (m, 8H, NBu4), 0.95 (t, J = 6.8 Hz, 12H, NBu4).11B NMR

(128.3 MHz, THF-d8, 25 °C) δ 1.26 (s). 13C NMR (100 MHz, THF-d8, 25°C) δ 158.47 (N(2)Ph ipso-C), 155.16 (B-Ph(1,2) ipso-C), 154.14 (N(1)Ph ipso-C), 142.91 (NCN), 141.43 ((benzyl)Ph ipsC)), 137.70 (NCN-p-tol ipsC), 137.19 (B-Ph(1) CH), 137.08 (B-Ph(2) o-CH), 135.55 (p-tol-CH3 ipso-C), 129.82 ((benzyl)Ph o-CH), 128.86 (p-tol m-CH), 127.60

((benzyl)Ph p-CH), 127.44 (p-tol o-CH), 126.90 (B-Ph(2) m-CH), 126.43 (B-Ph(1) m-CH), 126.27 (N(2)Ph o-CH), 125.85 (B-Ph(2) p-CH), 125.77 ((benzyl)Ph m-CH), 124.16 (B-Ph(1) p-CH), 123.61 (N(2)Ph m-CH), 123.53 (N(1)Ph m-CH), 118.39 (N(1) Ph o-CH), 116.36 (N(2)Ph p-CH), 113.77 (N(1)Ph p-CH), 59.67 (NBu4), 58.55 (benzyl-CH2), 24.78 (NBu4),

21.37 (p-tol CH3), 20.71 (NBu4), 14.08 (NBu4).

Synthesis of [BnLAlPh2][Na(THF)3] ([Bn2][Na]). Compound 22- (500 mg, 0.543 mmol) was

dissolved in 3 ml of THF in a small vial inside the glove box. To this was added 1 equiv of benzyl bromide, which caused the color to change from orange to yellowish-green. After stirring the mixture for 30 minutes, all the volatiles were removed under reduced pressure and the crude product was washed with hexane (3 x 2ml). Subsequently, drying under vacuum gave compound [Bn2][Na] as an oily green material. The oil was dissolved in a minimal amount of

(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: 130PDF page: 130PDF page: 130PDF page: 130

116

allowed to diffuse slowly at -30 °C. After three days, the light green crystals that had precipitated were isolated and washed with hexane (3 x 2 mL). The crystals were dried under vacuum to give compound [Bn2][Na] in 52 % yield (232 mg, 0.281 mmol). 1H NMR (400 MHz,

THF-d8, 25°C) δ 7.94 (d, J = 7.8 Hz, 2H, p-tol o-H), 7.86 (d, J = 6.4 Hz, 2H, Al-Ph(1) o-H),

7.62 – 7.60 (m, 2H, Al-Ph(2) o-H ), 7.37 (d, J = 8.0 Hz, 2H, N(2)Ph o-H), 7.13-7.06 (overlapped, 4H, (benzyl)Ph H and Al-Ph(1) m-H), 7.04 (overlapped, 3H, (benzyl)Ph p-H and N(1)Ph o-H), 6.97 (d, J = 7.8 Hz, 2H, p-tol m-o-H), 6.90-6.84 (overlapped, 8H, (benzyl)Ph m-H, Al-Ph(1) p-H, Al-Ph(2) (m+p)-H and N(2)Ph m-H), 6.69 (t, J = 7.5 Hz, 2H, N(1)Ph m-H), 6.34 (t, J = 7.0 Hz, 1H, N(2)Ph p-H), 6.09 (t, J = 6.9 Hz, 1H, N(1)Ph p-H), 4.44 (d, J = 12.5 Hz, 1H, benzyl-CH2), 4.29 (d, J = 12.5 Hz, 1H, benzyl-CH2), 3.68-3.59 (m, 4H, THF), 2.26 (s, 3H, p-tol CH3),

1.80-1.75 (m, 4H, THF). 13C NMR (100 MHz, THF-d8, 25°C) δ 157.29 (N(1)Ph ipso-C), 155.70

(N(2)Ph ipso-C), 154.24 (Al(1,2)Ph ipso-C), 146.35 (NCN), 139.97 (benzyl)Ph ipso-C), 139.65 (Al-Ph(1) o-CH), 139.56 (Al-Ph(2) o-CH),138.80 (NCN-p-tol ipso-C), 135.65 (p-tol-CH3

ipso-C), 131.17 (benzyl)Ph o-CH), 128.67 (p-tol m-CH), 128.39 (N(2)Ph m-CH), 128.15 (N(1)Ph m-CH), 127.90 (p-tol p-CH), 127.82 (Al-Ph(2) m-CH), 126.96 (Al-Ph(1) m-CH), 126.67 ((benzyl)Ph m-CH), 126.49 ((benzyl)Ph p-CH), 126.14 (Al-Ph(2) p-CH), 125.87 (Al-Ph(2) p-CH), 116.45 (N(2)Ph o-CH), 115.89 (N(2)Ph p-CH), 114.50 (N(1)Ph o-CH), 113.32 (N(1)Ph p-CH), 68.26 (THF), 58.60 (benzyl-CH2), 26.43 (THF), 21.34 (p-tol CH3).

Synthesis of [BnLAlPh2][NBu4] ([Bn2][NBu4]). Compound [Bn3][Na] (50 mg, 0.062 mmol)

was dissolved in 2 ml of THF in a vial inside the glove box. To this was added 1 equiv of tetrabutyl ammonium bromide (Bu4NBr) and the mixture was stirred overnight. The NaBr

that had precipitated was allowed to settle and the supernatant was transferred to another vial. After concentration under vacuum, a layer of hexane (1 mL) was added on top of the THF solution and the layers were allowed to slowly diffuse at -30 °C. After two days, a crystalline material had precipitated, which was isolated and washed with hexane (3 x 2 mL). Subsequently, drying under vacuum gave compound [Bn][NBu4] as a light green crystalline material (37 mg,

0.045 mmol, 73 %). 1H NMR (400 MHz, THF-d8, 25°C) δ 7.97 (d, J = 8.0 Hz, 2H, p-tol

o-H), 7.81 (d, J = 7.1 Hz, 2H, Al-Ph(1) o-o-H), 7.70-7.60 (m, 2H, Al-Ph(2) o-H ), 7.41 (d, J = 7.8 Hz, 2H, N(2)Ph o-H), 7.13-7.02 (overlapped, 7H, (benzyl)Ph (o+p)-H, Al-Ph(1) m-H and N(1)Ph o-H), 6.99 (d, J = 8.0 Hz, 2H, p-tol m-H), 6.96-6.84 (overlapped, 8H, (benzyl)Ph m-H, Al-Ph(1) p-H, Al-Ph(2) (m+p)-H and N(2)Ph m-H), 6.74 (t, J = 7.8 Hz, 2H, N(1)Ph m-H), 6.37 (t, J = 7.1 Hz, 1H, N(2)Ph p-H), 6.14 (t, J = 7.1 Hz, 1H, N(1)Ph p-H), 4.44 (d, J = 12.4 Hz, 1H, benzyl-CH2), 4.27 (d, J = 12.4 Hz, 1H, benzyl-CH2), 2.95 – 2.81 (m, 8H, NBu4), 2.27

(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: 131PDF page: 131PDF page: 131PDF page: 131

117 J = 7.3 Hz, 12H, NBu4). 13C NMR (100 MHz, THF-d8, 25°C) δ 157.12 (N(1)Ph ipso-C), 155.60

(N(2)Ph ipso-C), 153.82 (Al(1,2)Ph ipso-C), 146.33 (NCN), 139.98 (benzyl)Ph ipso-C), 139.68 (Al-Ph(1) o-CH), 139.61 (Al-Ph(2) o-CH),138.90 (NCN-p-tol ipso-C), 135.93 (p-tol-CH3

ipso-C), 131.27 (benzyl)Ph o-CH), 128.79 (p-tol m-CH), 128.50 (N(2)Ph m-CH), 128.35 (N(1)Ph m-CH), 128.06 (p-tol p-CH), 127.86 (Al-Ph(2) m-CH), 127.03 (Al-Ph(1) m-CH), 126.80 ((benzyl)Ph m-CH), 126.70 ((benzyl)Ph CH), 126.27 (Al-Ph(2) CH), 126.21 (Al-Ph(2) CH), 116.61 (N(2)Ph o-CH), 116.0 (N(2)Ph CH), 114.61 (N(1)Ph o-CH), 113.60 (N(1)Ph p-CH), 59.14 (NBu4), 58.52 (benzyl-CH2), 24.75 (NBu4), 21.36 (p-tol CH3), 20.55 (NBu4), 14.08

(NBu4).

5.7.3 X-ray crystallography

Suitable crystals of compounds [Bn1][NBu4] and [Bn2][Na] were mounted on top of a cryoloop

and transferred into the cold (100 K) nitrogen stream of a Bruker D8 Venture diffractometer. Data collection and reduction was done using the Bruker software suite APEX2.43 Data

collection was carried out at 100 K using Cu radiation (1.54178 Å) (for [Bn1][NBu4]) and Mo

radiation (0.71073 Å) (for [Bn2][Na]). The final unit cell was obtained from the xyz centroids

of 9929 ([Bn1][NBu4]) and 3515 ([Bn2][Na]) reflections after integration. A multiscan

absorption correction was applied, based on the intensities of symmetry-related reflections measured at different angular settings (SADABS).43 The structures were solved by intrinsic

phasing methods using SHELXT,44 and refinement of the structure was performed using SHELXL45. For [Bn1][NBu4], the refinement indicated the presence of smeared out

electron-density due to a disordered THF molecule that could not be modelled in a satisfactory manner. The contribution of this electron-density was removed using the PLATON/SQUEEZE routine,46

giving 4 solvent-accessible volumes in the unit cell, each containing 37 electrons (in agreement with THF). For [Bn2][Na], refinement was frustrated by a disorder problem: two of the THF

molecules bound to the Na+ cation showed unrealistic displacement parameters when refined

freely. For one of these, a two-site occupancy model was applied for all the atoms of the THF molecule and the site occupancy factor was refined (major fraction: 0.82). A SAME instruction was applied for the two disorder components, such that the two disordered THF molecules were restrained to have a similar geometry as that found for the THF molecule without disorder. Some atoms in the disordered THF molecule showed non-positive definite displacement parameters when refined freely, and RIGU/DELU instructions were applied. One of the carbon atoms of the remaining Na-THF fragment was split into two disorder components, which refined to a s.o.f. of 0.73 for the major component. The hydrogen atoms were generated by

(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: 132PDF page: 132PDF page: 132PDF page: 132

118

geometrical considerations, constrained to idealized geometries and allowed to ride on their carrier atoms with an isotropic displacement parameter related to the equivalent displacement parameter of their carrier atoms. Crystal data and details on data collection and refinement are presented in Table 5.1.

Table 5.1 Crystallographic data for [Bn1][NBu4] and [Bn2][Na]

[Bn1][NBu4] [Bn2][Na]

chem formula C55 H70 B N5 C51 H58 Al N4 Na O3

Mr 811.97 824.98

cryst syst Monoclinic Monoclinic

color, habit green, hexagonal yellow, block

size (mm) 0.21 x 0.19 x 0.06 0.36 x 0.33 x 0.33 space group P21/c P21/n a (Å) 11.0674(3) 14.8381(8) b (Å) 16.7937(4) 16.7498(10) c (Å) 27.9584(6) 18.1161(19) α (°) 90 90 β (°) 90.985(1) 96.559(4) γ (°) 90 90 V (Å3) 5195.7(2) 4473.0(6) Z 4 4 ρcalc, g.cm-3 1.038 1.225 Radiation [Å] 1.54178 0.71073 µ(Mo Kα), mm-1 0.102 µ(Cu Kα), mm-1 0.454 F (000) 1760 1760 temp (K) 100(2) 100(2) θ range (°) 3.07 - 75.01 2.958 - 29.686 Data collected (h,k,l) -13:11; -20:19; -33:32; -20:20; -23:23; -25:25; no. of rflns collected 51829 181202

(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: 133PDF page: 133PDF page: 133PDF page: 133

119

no. of indpndt reflns 9493 12588

observed reflns Fo ≥ 2.0 σ (Fo) 8215 10500 R(F) (%) 4.96 3.98 wR(F2) (%) 15.67 10.61 GooF 1.038 1.017 weighting a,b 0.0918, 2.0831 0.0471, 2.3119 params refined 555 598

min, max resid dens -0.343, 0.300 -0.276, 0.367

5.7.4 Computational studies

Calculations were performed with the Gaussian09 program47 using density functional theory

(DFT) in the gas phase. The geometries of the anions were fully optimized (after removing the counter cation from the crystallographically determined structure) using the B3LYP exchange-correlation functional with the 6-311+G(d,p) basis set. Optimizations were performed without (symmetry) constraints, and the resulting structures were confirmed to be minima on the potential energy surface by frequency calculations (number of imaginary frequencies = 0). The stationary points found for 22-calc, Bn1-calc and Bn2-calc closely resemble their crystallographically

determined structures, respectively. For the NBO analysis, single point calculations were performed on the final optimized geometries of 12-calc,3022-calc, Bn1-calc and Bn2-calc by including

keywords for NBO calculation.

To explore the chemical exchange process of 22- due to the rotation around the N-C(Ph) bond,

we scanned the dihedral angle between the two adjacent N atoms in the ligand and the ipso- and ortho-carbons of the Ph group in steps of 10 degrees (all other degrees of freedom were optimized). Starting from the highest energy points along the scan coordinate, the transition state was optimized and confirmed by frequency analysis (number of imaginary frequencies for

22-calc-ts =1).

(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: 134PDF page: 134PDF page: 134PDF page: 134

120

5.8 Supplementary information

5.8.1 UV-Vis spectroscopy

Figure S5.1 UV-vis spectra of [Bn1][Na],32 [Bn1][NBu4], [Bn2][Na] and [Bn2][NBu4] at room

temperature in THF

5.8.2 NMR spectra

Figure S5.2 1H NMR spectra of compound 22- at various temperatures in THF-d8. 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 260 310 360 410 460 510 ε(L it. m ol -1.cm -1x 10000) Wavelength (nm) [Bn1][Na] [Bn1][NBu4] [Bn2][Na] [Bn2][NBu4]

(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: 135PDF page: 135PDF page: 135PDF page: 135

121

5.8.3 Estimation of upper limit of the barrier for N-C(Ph) bond rotation in

1

2-The 1H NMR spectra of 12- do not show decoalescence down to 213 K. Even though the

frequency difference between the exchanging 1H resonances in 12- could thus not be measured

experimentally, an upper limit for the activation free energy can be estimated as follows: The coalescence temperature found for 22- is ca. 283 K with a frequency difference at slow

exchange of ca. 290 Hz; from this the exchange rate and activation free energy at the coalescence temperature can be determined as k = 644 s-1 and ΔG283 = 54 kJ/mol. Assuming

that the frequency difference is similar in 1-2, this means that in this compound the exchange

rate at 213 K is substantially larger than 644 s-1, meaning that the activation free energy at 213

K (ΔG‡213) is < 40 kJ/mol.

5.8.4 Determination of exchange kinetics in compound 2

2-

1H NMR data for compound 22- were collected in the temperature range 233 – 303 K. The

Varian NMR data files were converted to the gNMR49 file format using the gCVT tool included

in the gNMR installation. The chemical shifts of the peaks of interest (those of the N-Ph group) were taken from the experimental spectrum and exchange between pairs (the two ortho-H and meta-H) was modelled; the signal for the para-H was included without exchange. The latter peak was used to estimate the linewidth in the absence of chemical exchange (due to relaxation, inhomogeneity of the magnetic field etc.). Additional line broadening due to chemical exchange was then included, and the agreement between experimental and simulated spectrum was inspected visually. Due to the presence of additional peaks in the region of interest, attempts to perform least-squares fitting of the line shapes were unsuccessful. An estimate of the error in the exchange rate constants was made visually by running simulations with different rate constants and evaluating in which range a satisfactory fit was still obtained. A comparison between experimental spectra and those with ‘best’ fit parameters are shown in Figure S5.4. The rate constants thus obtained were used for constructing an Eyring plot of Ln(k/T) vs. 1/T. The estimated errors were taken into account by giving each data point a weight that was proportional to 1/(σ(k)2). Fitting was performed using Wolfram Mathematica 11.2,50 and

(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: 136PDF page: 136PDF page: 136PDF page: 136

122

Figure S5.3 Eyring plot for the calculation of activation parameters for rotation around the

N-C(Ph) bond in compound 22-. ∆H‡ = 54.1± 1.7 kJ/mol ∆S‡ = -2.5 ± 6.0 J/mol/K

233 K: estimated rate constant k = 2 ± 1 s-1

243 K: estimated rate constant k = 6 ± 2 s-1

273 K: estimated rate constant k = 170 ± 10 s-1 283 K: estimated rate constant k = 500 ± 100 s-1 253 K: estimated rate constant k = 27 ± 3 s-1

263 K: estimated rate constant k = 90 ± 10 s-1

(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: 137PDF page: 137PDF page: 137PDF page: 137

123

Figure S5.4 Comparison of experimental and simulated 1H-NMR spectra of 22- (for each

temperature, top: experimental spectrum, bottom: simulated). Rate constants used for the

simulation are shown for each spectrum, including an estimate of the error.

5.8.5 Analysis of reaction kinetics of [

Bn

1][NBu

4

] + TEMPO, [

Bn

2][Na] +

TEMPO and [

Bn

2][NBu

4

] + TEMPO

A solution of [Bn2][Na] in THF-d8 was prepared in a J. Young’s NMR tube and a few crystals

of bibenzyl were added as internal standard. Subsequently, 20 equiv of TEMPO were added to efficiently trap the benzyl radical generated upon thermolysis. The tube was taken out from the glovebox and inserted into the probe of the NMR spectrometer, which was heated to the desired reaction temperature. The NMR probe was tuned and shimmed before insertion of the sample using another NMR tube with the same amount of THF-d8 at that temperature. After the sample

was inserted, an acquisition array was started that was set up to collect spectra at regular time intervals throughout the course of the reaction (see Figure S8 for example of a few traces of the reaction of [Bn2][Na] with TEMPO (20 equiv)). Integration of the TEMPO-Bn resonance (δ

4.82 ppm in the 1H NMR spectrum) relative to the internal standard bibenzyl allowed analysis

of the rate constant at each temperature by non-linear curve fitting (a+b(1-Exp[-kt]) in Mathematica 11.2.50 To investigate the effects of counter cation, the kinetics of benzyl transfer

to TEMPO from [Bn1][NBu4] and [Bn2][NBu4] were measured using this procedure. For the

kinetics of benzyl transfer to TEMPO from [Bn1][NBu4] and [Bn2][NBu4], the residual proton

resonances from THF-d8 (at 1.73 ppm for [Bn1][NBu4] and at 3.58 ppm for [Bn2][NBu4],

respectively) were used as an internal standard due to the overlapping nature of the bibenzyl resonance (at 2.9 ppm) with the one of the proton resonances from NBu4+ (see Figure S9 and

Figure S10 for example of a few traces of the reaction of [Bn2][NBu4] and [Bn1][NBu4] with

TEMPO (20 equiv), respectively).

303 K: estimated rate constant k = 2000 ± 200 s-1

(23)

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: 138PDF page: 138PDF page: 138PDF page: 138

124

Figure S5.5 Selected 1H-NMR spectra from the 85°C kinetics array for the reaction of [Bn2][Na]

with TEMPO (20 equiv) in THF-d8.

Figure S5.6 Selected 1H-NMR spectra from the 85°C kinetics array for the reaction of

[Bn2][NBu4] with TEMPO (20 equiv) in THF-d8.

(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

Processed on: 21-11-2019 PDF page: 139PDF page: 139PDF page: 139PDF page: 139

125

Figure S5.7 Selected 1H-NMR spectra from the 85 °C kinetics array for the reaction of

[Bn1][NBu4] with TEMPO (20 equiv) in THF-d8.

(25)

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: 140PDF page: 140PDF page: 140PDF page: 140

126

Figure S5.9 Eyring analysis for [Bn2][Na]; ∆H = 107 ± 4 kJ∙mol-1 and ∆S = 17 ± 11 J∙mol -1∙K-1

Figure S5.10 Kinetic traces for the reaction of [Bn2][Na] (Blue) and [Bn2][NBu4] (Yellow) with

TEMPO (20 equiv) in THF-d8 at 75 °C (left) and at 85 °C (right).

Figure S5.11 Kinetic traces for the reaction of [Bn1][Na] (Blue)3 and [Bn1][NBu4] (Yellow) with

TEMPO (20 equiv) in THF-d8 at 65 °C (left) and at 85 °C (right).

340 345 350 355 360 T 0.000 0.005 0.010 0.015 0.020k Eyringplot 0.00280 0.00285 0.00290 0.00295 12.0 11.5 11.0 10.5 10.0 1 T Ln k T linearEyringplot

(26)

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: 141PDF page: 141PDF page: 141PDF page: 141

127

5.8.6 Computational and X-ray crystallographic characterization data

Figure S5.12 HOMOs of 22-calc (left) and the transition state for N-Ph rotation of 22-calc (2 2-calc-ts, right).

Table S5.1 Selected bond lengths (Å) for 12-,8 12-calc,8 [Bn1][NBu4], Bn1-calc, 22-, 22-calc,

[Bn2][Na] and Bn2-calc.

Compo-unds

12- 12-calc [Bn1][NBu4] Bn1-calc 22- 22-calc [Bn2][Na] Bn2-calc

B1-N1 1.578(3) 1.590 1.575(2) 1.585 Al1-N1 1.871(2) 1.896 1.866(9) 1.902 B1-N4 1.583(3) 1.590 1.584(2) 1.602 Al1-N4 1.873(2) 1.897 1.914(1) 1.946 B1-C21 1.637(4) 1.651 1.635(2) 1.645 Al1-C21 2.011(2) 2.03 1.991(1) 2.009 B1-C27 1.626(3) 1.645 1.624(3) 1.644 Al1-C27 2.008(2) 2.03 1.989(1) 2.015 N1-N2 1.428(3) 1.402 1.428(2) 1.426 N1-N2 1.432(2) 1.406 1.446(1) 1.441 N3-N4 1.433(3) 1.402 1.393(2) 1.356 N3-N4 1.432(2) 1.406 1.402(1) 1.364 N2-C7 1.325(3) 1.328 1.396(2) 1.413 N2-C7 1.332(2) 1.333 1.437(1) 1.443 C7-N3 1.326(3) 1.327 1.289(2) 1.293 C7-N3 1.328(2) 1.333 1.292(1) 1.293 N1-C1 1.379(3) 1.372 1.414(2) 1.412 N1-C1 1.375(2) 1.370 1.387(1) 1.386 N4-C15 1.384(3) 1.372 1.392(2) 1.404 N4-C15 1.371(3) 1.370 1.399(1) 1.403 N2-C33 1.468(2) 1.472 N2-C33 1.486(1) 1.493

(27)

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: 142PDF page: 142PDF page: 142PDF page: 142

128

Table S5.2 Selected bond angles (°) for [Bn1][NBu4], Bn1-calc, [Bn2][Na] and Bn2-calc.

Compounds [Bn1][NBu4] Bn1-calc [Bn2][Na] Bn2-calc

N1-B1-N4 105.78(1) 104.58 N1-Al1-N4 95.24(4) 95.02 C21-B1-C27 115.82(1) 112.73 C21-Al1-C27 112.46(5) 112.33 C21-B1-N1 109.5(1) 111.07 C21-Al1-N1 109.15(4) 113.2 C21-B1-N4 110.28(1) 108.93 C21-Al1-N4 111.73(4) 110.16 C27-B1-N1 105.45(1) 108.8 C27-Al1-N1 114.96(4) 112.92 C27-B1-N4 109.44(1) 110.43 C27-Al1-N4 112.14(4) 112.04 N2-C7-N3 127.57(1) 125.17 N2-C7-N3 126.80(9) 125.12 B1-N1-N2 113.06(1) 114.43 Al1-N1-N2 119.07(6) 114.84 B1-N1-C1 124.31(1) 124.8 Al1-N1-C1 127.89(7) 129.65 C1-N1-N2 110.44(1) 111.82 C1-N1-N2 112.55(8) 114.23 N1-N2-C7 115.41(1) 114.88 N1-N2-C7 112.27(8) 111.09 N1-N2-C33 113.94(1) 114.11 N1-N2-C33 110.33(8) 111.74 C7-N2-C33 118.79(1) 116.19 C7-N2-C33 112.13(8) 113.33 N3-N4-B1 118.8(1) 120.96 N3-N4-Al1 117.27(7) 120.34 B1-N4-C15 127.75(1) 126.84 Al1-N4-C15 125.47(7) 124.24 N3-N4-C15 110.99(1) 112.19 N3-N4-C15 111.72(8) 112.44

Table S5.3 Wiberg bond indices for the selected bonds in 12-calc, 12-calc-ts, 22-calc, 22-calc-ts

12-calc 12-calc-ts 22-calc 22-calc-ts

Bonds Wiberg Bond Index

N-C(Ph) 1.2196/1.2196 1.2262/1.0049 1.2412/1.2415 1.2605/1.0365

(28)

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: 143PDF page: 143PDF page: 143PDF page: 143

129

5.9 References

1 V. Lyaskovskyy and B. De Bruin, ACS Catal., 2012, 2, 270–279. 2 O. R. Luca and R. H. Crabtree, Chem. Soc. Rev., 2013, 42, 1440–1459.

3 D. L. J. Broere, R. Plessius and J. I. Van Der Vlugt, Chem. Soc. Rev., 2015, 44, 6886–6915.

4 J. B. Gilroy, M. J. Ferguson, R. McDonald, B. O. Patrick and R. G. Hicks, Chem. Commun., 2007, 0, 126– 128.

5 C. Chen, S. M. Bellows and P. L. Holland, Dalton Trans., 2015, 44, 16654–16670. 6 C. Camp and J. Arnold, Dalton Trans., 2016, 45, 14462–14498.

7 R. L. Webster, Dalton Trans., 2017, 46, 4483–4498.

8 Y. Liu, J. Li, X. Ma, Z. Yang and H. W. Roesky, Coord. Chem. Rev., 2018, 374, 387–415.

9 M. C. Chang, T. Dann, D. P. Day, M. Lutz, G. G. Wildgoose and E. Otten, Angew. Chem.Int. Ed, 2014, 53, 4118–4122.

10 M.-C. Chang and E. Otten, Chem. Commun., 2014, 50, 7431–7433.

11 R. Travieso-Puente, M. C. Chang and E. Otten, Dalton Trans., 2014, 43, 18035–18041.

12 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.

13 M. C. Chang and E. Otten, Organometallics, 2016, 35, 534–542.

14 M. C. Chang, A. Chantzis, D. Jacquemin and E. Otten, Dalton Trans., 2016, 45, 9477–9484.

15 M. C. Chang, P. Roewen, R. Travieso-Puente, M. Lutz and E. Otten, Inorg. Chem., 2015, 54, 379–388. 16 M.-C. Chang and E. Otten, Inorg. Chem., 2015, 54, 8656–8664.

17 S. M. Barbon, J. T. Price, P. A. Reinkeluers and J. B. Gilroy, Inorg. Chem., 2014, 53, 10585–10593. 18 M. Hesari, S. M. Barbon, V. N. Staroverov, Z. Ding and J. B. Gilroy, Chem. Commun., 2015, 51, 3766–3769. 19 S. M. Barbon, J. T. Price, U. Yogarajah and J. B. Gilroy, RSC Adv., 2015, 5, 56316–56324.

20 S. M. Barbon, J. V. Buddingh, R. R. Maar and J. B. Gilroy, Inorg. Chem., 2017, 56, 12003–12011. 21 S. Novoa and J. B. Gilroy, Polym. Chem., 2017, 8, 5388–5395.

22 S. M. Barbon, S. Novoa, D. Bender, H. Groom, L. G. Luyt and J. B. Gilroy, Org. Chem. Front., 2017, 4, 178– 190.

23 S. M. Barbon, V. N. Staroverov and J. B. Gilroy, Angew. Chem. Int. Ed, 2017, 129, 8285–8289.

24 R. R. Maar, A. Rabiee Kenaree, R. Zhang, Y. Tao, B. D. Katzman, V. N. Staroverov, Z. Ding and J. B. Gilroy,

Inorg. Chem., 2017, 56, 12436–12447.

25 J. S. Dhindsa, R. R. Maar, S. M. Barbon, M. Olivia Avilés, Z. K. Powell, F. Lagugné-Labarthet and J. B. Gilroy, Chem. Commun., 2018, 54, 6899–6902.

(29)

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: 144PDF page: 144PDF page: 144PDF page: 144

130

26 A. Van Belois, R. R. Maar, M. S. Workentin and J. B. Gilroy, Inorg. Chem., 2019, 58, 834–843. 27 A. Mandal, B. Schwederski, J. Fiedler, W. Kaim and G. K. Lahiri, Inorg. Chem., 2015, 54, 8126–8135. 28 G. Mu, L. Cong, Z. Wen, J. I. C. Wu, K. M. Kadish and T. S. Teets, Inorg. Chem., 2018, 57, 9468–9477. 29 E. Kabir, D. Patel, K. Clark and T. S. Teets, Inorg. Chem., 2018, 57, 10906–10917.

30 R. Mondol, D. A. Snoeken, M.-C. Chang and E. Otten, Chem. Commun., 2017, 53, 513–516. 31 R. Mondol and E. Otten, Inorg. Chem., 2019, 58, 6344–6355.

32 R. Mondol and E. Otten, Inorg. Chem., 2018, 57, 9720–9727. 33 C. L. Perrin and T. J. Dwyer, Chem. Rev., 1990, 90, 935–967. 34 A. D. Bain, Prog. Nucl. Magn. Reson. Spectrosc., 2003, 43, 63–103. 35 K. B. Wiberg, Tetrahedron, 1968, 24, 1083–1096.

36 D. Cremer and J. A. Pople, J. Am. Chem. Soc., 1975, 97, 1354–1358. 37 A. L. Spek, J. Appl. Crystallogr., 2003, 36, 7–13.

38 I. Polumbrik, O. Ryabokon, Zh. Org. Khim., 1978, 14, 1332.

39 A. Misyura, O. Polumbrik, L. Markovskii, Zh. Org. Khim., 1989, 25, 424.

40 C. W. Johnston, T. R. Schwantje, M. J. Ferguson, R. McDonald and R. G. Hicks, Chem. Commun., 2014, 50, 12542–12544.

41 J. L. Lukkarila, PhD thesis, University of Toronto, 2009.

42 R. R. Maar, S. D. Catingan, V. N. Staroverov and J. B. Gilroy, Angew. Chem. Int. Ed, 2018, 57, 9870–9874. 43 U. 2012. Bruker. APEX2 (v2012.4-3), SAINT (Version 8.18C) and SADABS (Version 2012/1). Bruker AXS Inc., Madison, Wisconsin, .

44 G. M. Sheldrick, Acta Crystallogr. Sect. A Found. Adv., 2015, 71, 3–8. 45 G. M. Sheldrick, Acta Crystallogr. Sect. C Struct. Chem., 2015, 71, 3–8. 46 A. L. Spek, Acta Crystallogr. Sect. C Struct. Chem., 2015, 71, 9–18.

47 Gaussian 09, Revision D.01, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, and D. J. Fox, Gaussian, Inc., Wallingford CT, 2009.

(30)

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: 145PDF page: 145PDF page: 145PDF page: 145

131 48 GaussView, Version 5, Dennington, Roy; Keith, Todd; Millam, John. Semichem Inc., Shawnee Mission, KS, 2009.

49 Budzelaar, P. H. M.; gNMR v. 5.0.6 (2006).

(31)

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

Referenties

GERELATEERDE DOCUMENTEN

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

The synthesis and characterization data for formazanate aluminum complexes presented in this chapter provide scope for further exploration of the reactivity of these compounds

De synthese en karakterisatie van de formazanaat-aluminiumcomplexen in dit hoofdstuk bieden ruimte voor verder onderzoek naar de reactiviteit van deze verbindingen door gebruik

I would like to thanks to the members of Brownie group ( Francesco, Davide Angelone, Duenpen, Juan Chen, Sandeep, Hans Kasper, Jorn, Linda, Tjalling, Luuk, Pat, Shaghayegh) and

Gupta, Ranajit Mondol, Stuart K Langley, Keith S Murray, Gopalan Rajaraman,. and Maheswaran Shanmugam, “Role of the Diamagnetic Zinc(II) Ion in Determining the Electronic Structure

The synthesis and characterization data for formazanate aluminum complexes presented in this chapter provide scope for further exploration of the reactivity of these compounds

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