• No results found

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

N/A
N/A
Protected

Academic year: 2021

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

Copied!
21
0
0

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

Hele tekst

(1)

University of Groningen

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

formazanate complexes

Milocco, Francesca

DOI:

10.33612/diss.160960083

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

it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date:

2021

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

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

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

Copyright

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

Take-down policy

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

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

(2)

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

Processed on: 2-2-2021 PDF page: 203PDF page: 203PDF page: 203PDF page: 203

193

Chapter 8

Palladium alkyl complexes with a

formazanate ligand:

synthesis, structure and reactivity

Palladium(II) complexes with a bidentate, anionic formazanate ligand are described. Attempts to prepare mono(formazanate) palladium alkyl complexes often leads to the homoleptic bis(formazanate) complex, which shows rich electrochemistry due to the redox-active nature of the ligands. Performing salt metathesis between the precursor [Pd(COD)(CH3)Cl] and the potassium salt of

the ligand in the presence of tetrabutylammonium chloride yields a square planar mono(formazanate) palladate complex through coordination of chloride anion. Ligand exchange allows binding of unsaturated molecules and evaluation of the reactivity of the Pd-CH3 fragment. Using this approach,

insertion reactions of CO, isocyanide and methyl acrylate into the Pd-CH3 bond are demonstrated.

This chapter has been published:

F. Milocco, F. de Vries, A. Dall'Anese, V. Rosar,E. Zangrando, E. Otten and B. Milani, Dalton. Trans. 2018, 47, 14445-14451. DOI: 10.1039/c8dt03130d Ligand-based redox event Insertions in Pd-CH3bond -2.5 -2 -1.5 -1 Potential vs Fc+/Fc (V)

(3)

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

Processed on: 2-2-2021 PDF page: 204PDF page: 204PDF page: 204PDF page: 204

194

8

8.1

Introduction

Alkylpalladium complexes with either neutral or monoanionic bidentate ligands are well-known for their ability to catalyze olefin oligomerization/polymerization. In particular, these catalysts show a remarkable tolerance to polar functional groups in comparison to early transition metal catalysts. Therefore, they have been studied extensively for coordination-insertion type copolymerization of ethylene with polar vinyl monomers, leading to polyolefins that incorporate functional groups in the polymer chain.1 Following the pioneering work by Brookhart and co-workers on catalysts containing

sterically demanding neutral D-diimine ligands,2 much research has been performed on these systems

to arrive at a thorough understanding of mechanistic details and structure/activity relationships.3 Our

long-standing interest in this type of catalysis has focused on the design of neutral N-donor ligands to improve catalyst activity/stability and obtain products with improved properties.4 Nickel(II) and

palladium(II) compounds with monoanionic bidentate ligands are also known as active catalysts for olefin polymerization (e.g., with phosphino-sulfonate5 or salicylaldimine ligands),6 but, surprisingly,

complexes with monoanionic nitrogen-based ligands such as E-diiminates have been comparatively little studied in group 10 chemistry.7 Limited success in the synthesis of palladium complexes with the

structurally related E-diiminate ligands has been reported in the literature. Bercaw and co-workers observed insertion of COD (1,5-cis,cis-cyclooctadiene) into the Pd-C bond of the transient species (E-diiminate)Pd(CH3)(COD).8 A competing C-H activation with the solvent (benzene) was also observed to

give the product of insertion into a Pd-Ph bond, while performing the reaction in the presence of acetonitrile afforded the adduct (E-diiminate)Pd(CH3)(CH3CN). Pörschke and co-workers described

attempts to synthesize E-diiminate palladium chloride complexes, which led to formation of palladium black and oxidative coupling of the E-diiminate ligand.9 Subsequently, the Song group has successfully

prepared a variety of N-aryl substituted mono(E-diiminate)palladium complexes and studied their reactivity.10 Complexes containing nitrogen-rich analogues of E-diiminates, such as

1,2,4,5-tetrazapentadienyl ligands (based on an NNCNN backbone, also known as ‘formazanates’),11

have not been studied extensively. Although the latter ligands have been sporadically used in coordination compounds with Ni(II)12 and Pd(II),13 it is only recently that they have gained renewed

interest due to their unusual redox- and optoelectronic properties.14,15,16,17 Aiming to exploit the

tuneable steric and electronic properties of formazanate ligands in palladium chemistry, we report here the synthesis and characterization of formazanate palladium complexes. This includes the first example of a methylpalladium complex with a formazanate ligand, and a study of insertion of unsaturated substrates (CO, isocyanide and methyl acrylate) in the Pd-C bond in this system.

8.2

Synthesis and characterization of palladium formazanate

complexes

8.2.1 Bis(formazanate) palladium complex

Initial attempts to generate mono(formazanate) palladium compounds by protonolysis (with free formazan 1H) or salt metathesis (with the corresponding potassium salt 1K18) using [Pd(COD)(CH3)Cl]

as the source of palladium were unsuccessful. Instead, these reactions invariably led to the formation of the homoleptic bis(formazanate) palladium complex Pd-1a, [Pd(L1)2], as the major species. To

(4)

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

Processed on: 2-2-2021 PDF page: 205PDF page: 205PDF page: 205PDF page: 205

195

confirm the identity of Pd-1a, it was prepared cleanly from the reaction of 2 equiv of 1H with Pd(OAc)2

and characterized by spectroscopy as well as X-ray crystallography. The characterization data are unremarkable, and correspond to those reported by Siedle and co-workers for related compounds.13a

In addition, given the recent interest in the redox-active properties of these and related ligands, we evaluated the electrochemistry of Pd-1a by cyclic voltammetry. The cyclic voltammogram shows two quasi-reversible reductions taking place at − 1.47 and − 1.82 V vs Fc0/+ (Fc = ferrocene), that correspond

to the transformation of Pd-1a to the radical anion Pd-1a− and subsequently to the dianion Pd-1a2−,

respectively (Figure 8.1 a). Moreover, a quasi-reversible oxidation to the radical cation Pd-1a+ is

observed at + 0.41 V vs Fc0/+. DFT calculations show predominant spin density on the ligands for all

open-shell species, confirming ligand-based reductions/oxidations taking place Figure 8.1 b-c).

Figure 8.1. a) Cyclic voltammogram of compound Pd-1a (ca. 1.50 mM solution of Pd-1a in THF; 0.1 M [Bu4N][PF6]

electrolyte; scan rate = 0.5 V·s-1). b) Spin density plot for the anion Pd-1a; c) spin density plot for the cation Pd-1a+

(DFT optimized geometries).

A more detailed examination of the species present during salt metathesis of 1K and [Pd(COD)(CH3)Cl]

was carried out by monitoring the reaction by NMR spectroscopy at low temperature in anhydrous THF-d8 solution. Addition of a cold (− 35 °C) solution of 1K in THF-d8 to [Pd(COD)(CH3)Cl]

(formazanate:Pd ratio of 1:1) resulted in immediate colour change from dark red to blue. In the initial

1H NMR spectrum taken at − 40 °C, no signal of the starting materials is present, and only a trace

amount of the bis(formazanate)palladium complex Pd-1a is observed. Although a complex mixture is obtained, free cyclooctadiene and the palladium compound [Pd(COD)(CH3)2] (in THF-d8, Pd-CH31H

NMR: 0.22 ppm; 13C NMR: 3.15 ppm)19,20 are unambiguously identified in the spectra. Judging by the

multitude of signals observed in the aromatic region, several new formazanate-contaning species are present.21 However, warming the mixture to room temperature again results in formation of the

bis(formazanate)palladium complex Pd-1a as the only formazanate-containing species (Scheme 8.1; see Section 8.5.3).

When the same reaction is performed at 0 °C on a larger scale (with adventitious water present), the dimeric complex [Pd(μ-OH)(L1)]2, Pd-1b, is formed as the major species, together with Pd-1a (Scheme

8.1). As indicated by NMR characterization (singlet at − 3.46 ppm) and X-ray diffraction studies on a single crystal (see Figure 8.2), in Pd-1b the palladium centres are connected by two bridging hydroxo groups. A similar μ-hydroxide complex was reported with nickel(II) and sterically demanding 3-nitro- or 3-cyanoformazanate ligands,12b and a related palladium(II) complex with a E-diiminate ligand is

(5)

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

Processed on: 2-2-2021 PDF page: 206PDF page: 206PDF page: 206PDF page: 206

196

known.10b The origin of the hydroxide groups in Pd-1b is likely an adventitious amount of water present

in the reaction solvent, or introduced during workup.

Scheme 8.1. Salt metathesis between 1K and [Pd(COD)(CH3)Cl].

Figure 8.2. Molecular structure of compounds Pd-1a and Pd-1b showing 50% probability ellipsoids, hydrogen atoms

are omitted for clarity.

8.2.2 Mono(formazanate) palladium complex

Based on these results, we hypothesized that successful synthesis of a mono(formazanate)palladium complex would require stabilization of the 14-electron, coordinatively unsaturated intermediate [Pd(L1)(CH3)] to prevent its conversion to Pd-1a. Thus, we examined the reaction of 1K with

[Pd(COD)(CH3)Cl] in the presence of one equivalent of tetrabutylammonium chloride. After stirring for

2 hours, the solution was filtered, after which diffusion of hexane into the THF solution afforded dark blue crystals of the product [Bu4N][Pd(L1)(CH3)Cl] (Pd-1c) in 83% isolated yield (Scheme 8.1). 1H NMR

spectroscopy showed a characteristic Pd-CH3 resonance at − 0.02 ppm (13C NMR: − 5.0 ppm), and the

observed integration was consistent with a formazanate:Pd-CH3 ratio of 1:1. The NMR spectra of Pd-1c

indicate that the two halves of the formazanate ligand are inequivalent, as expected for a square planar complex with Pd-CH3 and Pd-Cl fragments. An X-ray diffraction study confirmed the proposed

formulation (Figure 8.3 a). The anionic part of the molecule consists of a square planar Pd center bound to a bidentate formazanate, a chloride and methyl ligand. The Pd-N bond length is shortest for the nitrogen trans to the chloride ligand (Pd(1)-N(1) = 2.033(3) Å; Pd(1)-N(2) = 2.117(3) Å), which likely reflects the larger trans-influence of the CH3 group. The metrical parameters within the formazanate

(6)

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

Processed on: 2-2-2021 PDF page: 207PDF page: 207PDF page: 207PDF page: 207

197

THF solution (0.1 M [Bu4N][PF6] as electrolyte) shows that Pd-1c possesses a quasi-reversible reduction

wave at E1/2 = − 1.87 V vs. Fc0/+ (Figure 8.3 b), which is likely a ligand-based event; the 400 mV shift to

more negative potential than observed for Pd-1c is consistent with the overall charge of −1 for palladate complex Pd-1c. Somewhat surprisingly, the quasi-reversible nature indicates that the electrochemically generated radical dianion [Pd(L1)(CH3)Cl]2− does not lose Cl− on the timescale of the

voltammetry experiment.

Figure 8.3. Molecular structure of compound Pd-1c showing 50% probability ellipsoids. Hydrogen atoms and the

Bu4N+ cation are omitted for clarity. b) Cyclic voltammogram of compound Pd-1c (ca. 1.50 mM solution of Pd-1c in

THF; 0.1 M [nBu4N][PF6] electrolyte; scan rate = 0.1 V·s-1). Selected bond lengths (Å) and angles (°): Pd(1)-N(1)

2.033(3); Pd(1)-N(2) 2.117(3); Pd(1)-Cl(1) 2.3055(10); Pd(1)-C(21) 2.062(3); N(1)-N(4) 1.293(4); N(2)-N(3) 1.297(4); N(1)-Pd(1)-N(2) 82.37(12); ‘(N–Pd–N)/(X–Pd–X) 5.28.

The straightforward synthesis of Pd-1c confirms that transmetallation between 1K and [Pd(COD)(CH3)Cl] is facile and indeed produces the target mono(formazanate)palladium compound.

The presence of a fourth ligand to complement the Pd coordination sphere is required to stabilize the neutral fragment [Pd(L1)(CH3)]. This is in agreement with data reported for β-diketiminate nickel

compounds, which are stabilized by binding lutidine as neutral co-ligand or via β-agostic interactions with the alkyl group.22 A similar situation is reported for organometallic palladium complexes having

the phosphinosulfonate anion as ancillary ligand, which have been isolated as neutral lutidine23 or

DMSO adducts,24 or as anionic complexes with chloride bound.25 In our case, the presence of chloride

ions leads to formation of a stable four-coordinate palladate complex. Such palladate species are also of current interest due to their involvement in cross-coupling catalysis.26

8

8.2.3 Attempts of chloride abstraction

With the successful synthesis of Pd-1c in hand, we questioned whether chloride abstraction would allow to detect the (transient) putative three-coordinate compound [Pd(L1)(CH3)] and provide insight

in the pathway of its conversion to Pd-1a (vide supra). Thus, THF-d8 solutions of Pd-1c were reacted at

− 30 °C in three separate experiments with NaBPh4, AgBF4 or B(C6F5)3, and the reactions were

monitored by NMR spectroscopy at − 40 °C. A detailed discussion of these experiments is provided in

Section 8.5.4. Here, we note that in none of these reactions we were able to unambiguously identify

[Pd(L1)(CH3)] as an intermediate. The reaction with NaBPh4 results in formation of Pd-1a as the major

product, presumably as a result of chloride abstraction from Pd-1c. Similarly, B(C6F5)3 reacts with Pd-1c

to form a 2:1 mixture of the borates [B(C6F5)3Cl]- and [B(C6F5)3(CH3)]- due to chloride/methyl

abstraction from Pd-1c; again, Pd-1a is the major formazanate-containing product in solution. Finally, the reaction with AgBF4 does not lead to chloride abstraction, but the data indicate that

-8.00E-06 -6.00E-06 -4.00E-06 -2.00E-06 0.00E+00 2.00E-06 4.00E-06 6.00E-06 -2.5 -2 -1.5 -1 Applied potential (V) vs Fc0/+ a) 2.00 4.00 6.00 -2.00 -4.00 0.00 -6.00 -8.00 Cu rre nt A) b)

(7)

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

Processed on: 2-2-2021 PDF page: 208PDF page: 208PDF page: 208PDF page: 208

198

electron-transfer could occur as evidenced by the formation of ethane and a species that we tentatively assign as the chloride-bridged dimer [Pd(μ-Cl)(L1)]2 (

Scheme 8.2), for which there is precedent in related E-diiminate chemistry.27

Scheme 8.2. Proposed reaction sequence for reaction of Pd-1c with AgBF4 in THF-d8.

8

8.3

Ligand exchange and insertion reactions

Ligand exchange and insertion reactions with the palladate complex Pd-1c were evaluated (Scheme 8.3). Stirring complex Pd-1c with 2.6 equiv of pyridine in dichloromethane solution gave the neutral pyridine adduct [Pd(L1)(CH3)(Py)] (Pd-1d), which was isolated in 46% yield. The 1H NMR spectrum of

Pd-1d shows the singlet of the Pd-CH3 group at 0.05 ppm, and inequivalent formazanate NPh groups,

in agreement with the non-symmetrical chemical environment around palladium.

Addition of excess CO to a THF-d8 solution of compound Pd-1c resulted in clean formation of the CO

insertion product [Bu4N][Pd(L1)(COCH3)Cl] (Pd-1e), which shows NMR resonances at G 2.33 (1H) and

235.5/37.4 ppm (13C) that are characteristic for the palladium acyl group.28 A labelling experiment using 13CO confirmed that the Pd-C resonance at 235.7 ppm in the 13C NMR was enhanced due to the

incorporation the 13C label. Aside from free 13CO, no other signals attributable to 13C incorporation

were observed, indicating that the chloride ligand is retained in the anionic product Pd-1e rather than Cl−/13CO exchange to give a neutral 13C-labeledpalladium carbonyl species. Treatment of Pd-1c with 1

equivalent of 4-methoxyphenyl isocyanide on NMR scale in THF-d8 resulted in a mixture of products

and some unreacted Pd-1c, but addition of a second equivalent of isocyanide led to the formation of a single species according to NMR spectroscopy which was characterized as the isocyanide insertion product [Pd(L1)(C(CH3)=NC6H4OCH3)(CNC6H4OCH3)] (Pd-1f, Scheme 8.3).

Diagnostic NMR data for Pd-1f include a singlet at 2.41 ppm for the iminoacyl-CH3 group, and a 13C

NMR signal at 184.5 ppm for the Pd-bound C=N carbon. Compounds Pd-1e and Pd-1f are both sufficiently stable to allow characterization by NMR spectroscopy, but solutions left at room temperature decompose in the course of several hours to unidentified species (including palladium black).

Also the pyridine adduct Pd-1d inserts CO, albeit more slowly than the anionic chloride Pd-1c, to give compound Pd-1h which shows acyl-CH3 NMR resonances at G 1.80 (1H) and 36.7 ppm (13C). The reaction

of Pd-1d with CO in the presence of additional pyridine (ca. 2.2 equiv) is qualitatively found to be slower than without added pyridine: in the former only 13% conversion of Pd-1d is obtained after 3 hours, while in the latter it reaches > 85% conversion in the same time. We tentatively attribute this to CO

(8)

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

Processed on: 2-2-2021 PDF page: 209PDF page: 209PDF page: 209PDF page: 209

199

insertion being preceded by (dissociative) ligand exchange between pyridine and CO (and similarly, Cl−

and CO in Pd-1c). Although square planar 16-electron complexes typically undergo substitution by an associative mechanism, dissociative pathways are also reported in the literature,29 in particular for

electron-rich complexes with strong donor ligands that raise the energy of the metal-based LUMO (dz2

orbital) and/or with ligands that have a high trans-influence. The (imino)acyl products formed upon insertion are obtained as four-coordinate compounds with the fourth ligand in the palladium coordination sphere being either Cl− (Pd-1e), CN-p-An (Pd-1f), or Py (Pd-1h).

Subsequently, insertion of olefinic substrates was tested. Thus, 5 equiv of styrene were added to a THF-d8 solution of Pd-1c, and the reaction monitored by 1H NMR spectroscopy at room temperature.

No changes were observed at room temperature over the course of 16 hours, after which the NMR tube was filled with 1 bar of CO. Immediate formation of the acyl complex Pd-1e was observed, but subsequent insertion of styrene did not occur. Finally, to promote styrene coordination/insertion, abstraction of the chloride ligand from the four-coordinate complex Pd-1e was performed by the addition of NaBPh4. However, this resulted again in the rapid formation of Pd-1a, and no evidence for

conversion of the olefin was seen. In agreement with these observations, tests of catalytic reactions under CO/styrene copolymerization conditions with compounds Pd-1c/Pd-1d only showed decomposition to Pd-1a and some free ligand; no oligomeric/polymeric products were obtained. Similarly, the addition of 1 bar of ethylene to compound Pd-1d does not lead to observable changes in the NMR spectrum when followed over several days.

Scheme 8.3. Ligand exchange reactions with Pd-1c and insertion of unsaturated substrates in the Pd-CH3 bond.

In contrast, the reaction of Pd-1d with methyl acrylate slowly converts to a new species which is formulated as the 2,1-insertion product [Pd(L1)(CH(Et)CO2CH3)(Py)] (Pd-1h) on the basis of its NMR

spectral features. Specifically, the observation of a set of diastereotopic CH2 resonances in the 1H NMR

spectrum at δ 1.25 and 1.06 ppm, which show scalar coupling to a Pd-bound CH (dd at δ 2.34 ppm) and a CH3 group (t at δ 0.70 ppm), is consistent with the proposed 2,1-insertion. Although cationic

palladium complexes initially insert methyl acrylate with the same regiochemistry, subsequent β-hydride elimination and re-insertion events lead to six-membered chelate rings in which the C=O moiety is bound to Pd.1a, 30 For neutral phosphinosulfonate-ligated palladium compounds, 2,1-insertion

(9)

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

Processed on: 2-2-2021 PDF page: 210PDF page: 210PDF page: 210PDF page: 210

200

of methyl acrylate occurs which allows formation of a stable six-membered chelate ring.31 The

observed insertion of methyl acrylate for Pd-1d contrasts the finding of Novak and co-workers, who demonstrated that neutral palladium complexes with anionic pyrrole-imine ligands react with methyl acrylate via a radical mechanism instead of coordination-insertion.32

Although insertion of CO, isocyanides and olefins is well-documented with palladium(II) organometallics,4a, 33 the vast majority of these are for cationic complexes containing neutral ligands

(e.g., phosphines, bipyridines, D-diimines), and organometallic palladium chemistry with anionic nitrogen ligands is much less developed. Our results present rare examples of insertion chemistry in Pd(II) alkyl complexes with a monoanionic, bidentate nitrogen ligand.8, 34

8

8.4

Conclusion

Protonolysis or salt metathesis reactions to obtain mono(formazanate) palladium complexes lead instead to formation of the homoleptic bis(formazanate)palladium complex (Pd-1a) as a thermodynamic sink. For complex Pd-1a, a redox-series spanning 4 different oxidation states based on ligand-centred redox reactions is observed by cyclic voltammetry, highlighting the rich electrochemistry of formazanate ligands. Our attempts to synthesize mono(formazanate) analogues via salt metathesis starting from the precursor [Pd(COD)(CH3)Cl] were successful in the presence of Cl−

as an additional ligand to complement the (square planar) coordination sphere around Pd, which allowed the synthesis and structural characterization of the methylpalladate complex Pd-1c. Exchange of the chloride ligand afforded the neutral pyridine adduct Pd-1d. Insertion reactions into the Pd-CH3

bond in these mono(formazanate) palladium alkyl compounds were examined, and the products of CO, isocyanide and methyl acrylate insertion were characterized by NMR spectroscopy. In contrast to cationic Pd(II) complexes with a neutral bidentate N-donor ligand, which in general are highly reactive towards olefins, insertion of ethylene or styrene did not occur. The sluggish olefin insertion kinetics, together with decomposition to the bis(formazanate) complex Pd-1a, are responsible for the lack of catalytic oligomerization/polymerization reactivity with these compounds.

(10)

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

Processed on: 2-2-2021 PDF page: 211PDF page: 211PDF page: 211PDF page: 211

201

8

8.5

Experimental Section

Synthetic procedure of compounds Pd-1a and Pd-1b, X-Ray crystallographic data, 1D and 2D 1H, 19F and 13C NMR

characterization spectral data of all compounds, catalytic reactions are reported in the ESI, see DOI: 10.1039/c8dt03130d.

8.5.1 General Considerations

Compounds 1H15a and 1K18 were synthesized according to literature procedures. [Pd(CH3)Cl(COD)] was

either prepared according to the literature procedure,35 or obtained commercially (99%, Strem

Chemicals). HCl (37%, Fluka) and 1,5-cis,cis-cyclooctadiene (Fluka) (used for the synthesis of [Pd(CH3)Cl(COD)]), and tetrabutylammonium chloride (99%, Sigma-Aldrich), 4-methoxyphenyl

isocyanide (97%, Sigma-Aldrich) and [Pd(CH3COO)2] (Engelhard Italy) were used as received.

8.5.2 Synthesis of the complexes

[Bu4N][Pd(L1)(CH3)Cl] (Pd-1c).

Tetrabutylammonium chloride (54.2 mg, 1 eq, 0.195 mmol) was added to a pale yellow solution of [Pd(COD)(CH3)Cl] (50.0 mg, 1 eq, 0.195 mmol) in THF (15 mL). 1K (96.0 mg, 0.5 eq, 0.0975 mmol) was

slowly added as a solid to the reaction mixture while stirring. Immediately the reaction mixture turned blue-turquoise. After stirring for 2h the reaction mixture was filtered and slow diffusion of hexane into the THF solution at room temperature afforded 114.9 mg of Pd-1c as blue crystals suitable for X-Ray diffraction (0.161 mmol 83%). 1H NMR (500 MHz, THF-d8, 25 °C): δ 8.41 (d, 2H, Ph o-CH), 8.05 (d, 2H,

Ph o-CH), 7.99 (d, 2H p-Tol o-CH), 7.22-7.16 (m, 6H Ph and Ph' m-CH and p-Tol m-CH), 7.00-6.95 (m, 2H Ph and Ph' p-CH), 3.19 (m, 8H, NBu4+ CH2), 2.36 (s, 3H, p-Tol CH3), 1.49 (m, 8H, NBu4+ CH2), 1.28 (m, 8H,

NBu4+ CH2) 0.88 (t, 12H, NBu4+ CH3) 0.02 (Pd-CH3). 13C NMR (101 MHz, THF-d8, 25 °C) δ 156.0 (Ph'

ipso-C), 155.1 (Ph ipso-C), 149.3 (NCN), 138.7 (p-Tol ipso-C), 136.1 (p-Tol ipso-CCH3), 129.6 (p-Tol m-C),

128.4 (Ph and Ph' m-C), 125.8 (Ph and Ph' p-C), 125.6 (Ph o-C), 125.5 (Ph' o-C), 125.0 (p-Tol o-C), 59.3 (NBu4+ CH2), 24.8 (NBu4+ CH2), 21.5 (p-Tol p-C), 20.7 (NBu4+ CH2), 14.2 (NBu4+ CH3), − 5.0 (Pd- CH3). Anal.

calcd for C37H56N5ClPd: C 62.35, H 7.92, N 9.83; found: C 62.19, H 7.93, N 9.79. IR: νmax = 1276 cm-1

(stretch C-N), 1188 cm-1 (stretch N-N).

[Pd(L1)(CH3)(py)] (Pd-1d).

Complex Pd-1c (99.7 mg, 0.14 mmol) was dissolved in 3.5 mL of distilled DCM under argon atmosphere, to give a turquoise coloured solution. A solution of pyridine (30 μL, 29.5 mg, 0.37 mmol) in DCM (2 mL) was added to the palladium complex under stirring, which gave an immediate color change to a deep blue solution. The reaction mixture was stirred for 3 hours, after which it was filtered over Celite. The filtrate was taken to dryness and redissolved in 2 mL of DCM, to which 14 mL hexane was added. Overnight, a solid precipitated which was filtered off and the filtrate was taken to dryness to afford

Pd-1d as a dark blue solid (33.0 mg, 0.064 mmol, 46%). 1H NMR (500 MHz, CD2Cl2, 25 °C): δ 8.52 (d, 2H,

pyridine CH(2,6)), 8.02 (d, 2H, Ph o-CH), 8.01 (d, 2H, p-Tol o-CH), 7.74 (d, 2H, Ph' o-CH), 7.60 (t, 1H, pyridine CH(4)), 7.39 (t, 2H, Ph m-CH), 7.26 (d, 2H, p-Tol m-CH), 7.19 (t, 1H, Ph p-CH), 7.13 (t, 2H, pyridine CH(3,5)), 7.09 (t, 2H, Ph m-CH), 6.93 (t, 1H, Ph p-CH), 2.42 (s, 3H, p-Tol p-CH3), 0.04 (s, 3H,

Pd-CH3). 1H-13C HSQC NMR (500 MHz, CD2Cl2, 25 °C): δ 152.86 (pyridine CH(2,6)), 137.46 (pyridine

CH(4)), 129.21 (p-Tol m-CH), 128.41 (Ph' m-CH), 128.32 (Ph m-CH), 126.06 (Ph p-CH), 125.84 (Ph' p-CH), 125.25 (pyridine CH(3,5)), 124.93 (p-Tol o-CH & Ph o-CH), 123.72 (Ph' o-CH), 21.20 (p-Tol p-CH3), -0.35

(11)

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

Processed on: 2-2-2021 PDF page: 212PDF page: 212PDF page: 212PDF page: 212

202

8

8.5.3 In situ NMR reactivity

In situ NMR reactivity of 1K + [Pd(COD)(CH3)Cl] in THF-d8

In a glovebox, a cold (− 30 ˚C) solution of 1K (0.5 eq, 11.2 mg, 1.13·10-2 mmol) in THF-d8 was added to

a cold (− 30 ˚C) solution of [Pd(COD)(CH3)Cl] (1 eq, 6.0 mg, 2.26·10-2 mmol) in THF-d8 in a Young’s NMR

tube leading to a blue reaction mixture. The sample was quickly frozen in liquid nitrogen and subsequently inserted into an NMR probe pre-cooled to − 40 °C. The reaction was followed for 1 day at − 40 °C. The first 1H NMR spectrum (Figure S38) shows that the starting materials are completely

converted and only a trace amount of Pd-1a is formed, which does not further increase in intensity. The aromatic region has several unidentified resonances overlapped. Species that was possible to identified are free cyclooctadiene and the palladium compound [Pd(COD)(CH3)2] (in THF-d8, PdMe 1H

NMR: 0.22 ppm; 13C NMR: 3.15 ppm).19 Moreover, there is a new 1H NMRresonance at − 0.05 ppm

(13C: − 5.7 ppm), which reasonably is due to a Pd-Me moiety of the putative compound [(1)2PdMe]-.

The reaction was warmed up to + 25 °C and followed for 6 days. After 24 h at room temperature the mixture is fully converted to Pd-1a. In the 1H NMR spectrum the resonances of Pd-1a, free COD and

[Pd(COD)(CH3)2] account for > 90% of the total signal intensity (Figure S38). At room temperature the

[Pd(COD)(CH3)2] decomposed over time, and formation of a palladium mirror and methane (1H NMR:

0.19 ppm in THF-d8) and ethane (1H NMR: 0.85 ppm in THF-d8) was observed. After 6 days, compound

Pd-1a is the major species with ca. 10% of unknown byproduct(s).

Figure 8.4. 1H NMR spectra of 1K + [Pd(COD)(CH

3)Cl] in distilled THF-d8: a) immediately after mixing at − 40 °C; b)

after standing at room temperature for 1 day.

In situ NMR reactivity of 1H + [Pd(COD)(CH3)Cl] in THF-d8

In a glovebox, 1H (1 eq, 3.8 mg, 1.2·10-2 mmol) was added as a solid to a THF-d8 solution of

[Pd(COD)(CH3)Cl] (1 eq, 3,2 mg, 1.2·10-2 mmol) in a J. Young’s NMR tube. The first 1H NMR spectrum

shows the starting materials as major species with the addition of some new minor resonances. The reaction was followed for 1 week at + 25 °C. After 2 days the resonances of compound Pd-1a start

Pd(COD)(CH3)2 Pd-1c THF THF Pd(COD)(CH3)2 COD Pd-1a Pd-1a a) b)

(12)

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

Processed on: 2-2-2021 PDF page: 213PDF page: 213PDF page: 213PDF page: 213

203

appearing, but the main species are still the starting material. After 1 week at room temperature, the reaction mixture was kept for 3 days at + 60 °C, which led to a brown solution and precipitation of a blue amorphous solid (uncharacterized). The 1H NMR spectrum shows that [Pd(COD)(CH3)Cl] has fully

reacted, and the major soluble (diamagnetic) species is compound Pd-1a. In addition, there are also some minor unknown formazanate-containing species.

Figure 8.5. 1H NMR spectrum of 1H + [Pd(COD)(CH3)Cl], (THF-d8, 25 °C, 500 MHz) : a) t = 20 min; b) t = 1 week at +

25 °C, c) after 3 days at + 60 °C.

NMR reaction of Pd-1c with CO in THF-d8 to give compound Pd-1e.

In a glovebox, compound Pd-1c (5.8 mg, 8.1·10-3 mmol) was dissolved in a J. Young’s NMR tube in

THF-d8. The solution was frozen in liquid N2 and the N2 atmosphere inside the J. Young’s NMR tube was

removed at the schlenk line. Subsequently, the NMR tube was allowed to warm to room temperature and then was filled with 1 bar of CO, observing the solution turning to a darker shade of blue. The in

situ formed product (Pd-1e) was characterized via NMR spectroscopy at − 40 °C. The first spectrum

was recorded after 20 min from the addition of CO and shows that the starting material has been completely converted. 1H NMR (500 MHz, THF-d8, − 40 °C): δ 8.37 (d, 2H, Ph or Ph' o-CH), 8.04 (d, 2H,

p-Tol o-CH), 7.97 (d, 2H, Ph or Ph' o-CH), 7.27-7.22 (m, 6H, p-Tol m-CH, Ph and Ph' m-CH), 7.06 (t, 1H,

Ph or Ph' p-CH) 7.01 (t, 1H, Ph or Ph' p-CH), 3.22 (m, 8H, NBu4+ CH2), 2.38 (6H, p-Tol CH3 and Pd-COCH3),

1.37 (m, 8H, NBu4+ CH2), 1.25 (m, 8H, NBu4+ CH2) 0.86 (t, 12H, NBu4+ CH3) ppm. 13C NMR (126 MHz,

THF-d8, − 40 °C): δ 235.8 (Pd-COCH3), 158.1 (Ph or Ph' ipso-C), 154.6 (Ph or Ph' ipso-C), 146.1 (NCN),

138.4 (p-Tol ipso-C), 136.4 (p-Tol ipso-CCH3), 124.8 (Ph or Ph' o-CH), 124.8 (p-Tol o-CH), 124. 7 (d, 2H

Ph or Ph' o-CH), 130.1 (p-Tol m-CH) 128.7 (Ph or Ph' m-CH), 128.5 (Ph or Ph' m-CH), 126.5 (Ph or Ph'

p-CH) 126.2 (Ph or Ph' p-CH), 58.4 (NBu4+ CH2), 37.4 (Pd-COCH3) 24.4 (NBu4+ CH2), 21.6 (p-Tol CH3) 20.6

(NBu4+ CH2), 14.5 (NBu4+ CH3)ppm. 1H THF-d8 1H THF-d8 [Pd(COD)(CH3)Cl] [Pd(COD)(CH3)Cl] 1H free cod THF-d8 THF-d8 free cod Pd-1a Pd-1a CH4 C2H6 a) b) c)

(13)

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

Processed on: 2-2-2021 PDF page: 214PDF page: 214PDF page: 214PDF page: 214

204

NMR reaction of Pd-1c with 4-methoxyphenyl isocyanide in THF-d8 to give compound Pd-1f.

In a glovebox, 4-methoxyphenyl isocyanide (1 eq, 2.0 mg, 1.5·10-2 mmol) was added as a solid to a J.

Young’s NMR tube containing a blue-turquoise THF-d8 solution of Pd-1c (1 eq, 10.7 mg, 1.5·10-2 mmol).

After recording a 1H NMR spectrum at t = 30 min at − 40 °C a second equivalent of isocyanide was

added (2.0 mg, 1.5·10-2 mmol) to the reaction mixture which turned to a darker shade of blue. The in

situ formed product (Pd-1f) was characterized via NMR spectroscopy at − 40 °C. 1H NMR (500 MHz,

THF-d8, − 40 °C): δ 8.18 (d, 2H Ph' o-CH), 7.76 (d, 2H Ph o-CH), 7.71 (d, 2H p-Tol o-CH), 7.47 (t, 2H Ph'

m-CH), 7.35 (t, 2H Ph m-CH), 7.24-7.15 (m, 6H, CNPhOCH3 o-CH, Ph and Ph' p-CH, p-Tol m-CH), 7.00 (d,

2H, CNPhOCH3 m-CH), 6.83 (d, 2H, CH3CNPhOCH3 o-CH), 6.55 (d, 2H, CH3CNPhOCH3 m-CH), 3.81 (s, 3H,

CNPhOCH3), 3.20 (s, 3H, CH3CNPhOCH3), 2.48 (s, 3H, CH3CNPhOCH3), 2.39 (s, 3H, p-Tol CH3). 13C NMR

(126 MHz, THF-d8, − 40 °C): δ 184.5 (CH3CNPhOCH3), 162.1 (CNPhOCH3 ipso-COCH3), 157.0

(CH3CNPhOCH3 ipso-COCH3), 156.8 (Ph' ipso-C), 156.4 (Ph ipso-C), 146.8 (NCN), 146.6 (CH3CNPhOCH3

ipso-CNC), 141.3 (CNPhOCH3),136.9 (p-Tol ipso-CCH3), 136.6 (p-Tol ipso-C), 129.8 (Ph' m-CH), 129.6

(p-Tol m-CH), 129.0 (CNPhOCH3 o-CH), 128.8 (Ph m-CH), 127.4 (Ph p-CH), 127.3 (Ph' p-CH), 125.5

(p-Tol o-CH), 125.4 (Ph o-CH), 124.6 (Ph' o-CH), 121.7 (CH3CNPhOCH3 o-CH), 119.5 (CNPhOCH3

ipso-CNC), 115.9 (CNPhOCH3 m-CH), 113.7 (CH3CNPhOCH3 m-CH), 56.4 (CNPhOCH3), 54.7

(CH3CNPhOCH3), 35.9 (CH3CNPhOCH3), 21.6 (p-Tol CH3).

NMR reaction of Pd-1d with CO in CD2Cl2 to give compound Pd-1g.

Compound Pd-1d (3.86 mg, 7.51 μmol) was placed in a NMR tube and dissolved in 0.75 mL CD2Cl2

resulting in a turquoise solution. CO was bubbled through the solution for 10 minutes which resulted in a slight darkening of the solution to a dark blue color. The reaction was followed over the course of several hours, during which the reaction proceeded and after 200 minutes the reaction had almost reached completion (95% conversion). The in situ formed product (Pd-1g) was characterized via NMR spectroscopy. 1H NMR (400 MHz, CD2Cl2, 25 °C): δ 8.71 (d, 2H, pyridine CH(2,6)), 8.08 (d, 2H, p-Tol

o-CH), 8.03 (d, 2H, Ph o-CH), 7.59 (t, 1H, pyridine CH(4)), 7.57 (d, 2H, Ph' o-CH), 7.38 (t, 2H, Ph m-CH),

7.30 (d, 2H, p-Tol m-CH), 7.19 (t, 1H, Ph p-CH), 7.14 (t, 2H, pyridine CH(3,5)), 7.05 (t, 2H, Ph' m-CH), 6.91 (t, 1H, Ph' p-CH), 2.44 (s, 3H, p-Tol p-CH3), 1.81 (m, 3H, Pd-COCH3) ppm. 1H-13C HSQC NMR

(400MHz, CD2Cl2, 25 °C): δ 151.73 (pyridine CH(2,6)), 137.61 (pyridine CH(4)), 128.97 (p-Tol m-CH),

128.28 (Ph m-CH), 128.08 (Ph' m-CH), 126.34 (Ph p-CH), 125.70 (Ph' p-CH), 124.73 (p-Tol o-CH), 124.33 (Ph o-CH), 124.56 (pyridine CH(3,5)), 122.77 (Ph' o-CH), 36.72 (Pd-COCH3), 20.82 (p-Tol p-CH3) ppm.

NMR reaction of Pd-1d with methyl acrylate in CD2Cl2 to give compound Pd-1h.

Compound Pd-1d (3.80 mg, 7.4 μmol) was placed in a NMR tube and dissolved in 0.75 mL CD2Cl2, to

which 1.35 μL (1.29 mg, 15 μmol) methyl acrylate was added. The reaction was followed over the course of several hours, during which the reaction proceeded very slowly. After 102 hours, the reaction had reached 87% conversion, and the product was characterized by NMR spectroscopy. Attempts to isolate the product by recrystallization were unsuccessful. 1H NMR (500 MHz, CD2Cl2, 25 °C): δ 8.51 (d,

2H, pyridine CH(2,6)), 8.19 (d, 2H, Ph o-CH), 7.99 (d, 2H, p-Tol o-CH), 7.66 (d, 2H, Ph' o-CH), 7.56 (t, 1H, pyridine CH(4)), 7.45 (t, 2H, Ph m-CH), 7.27 (d, 2H, p-Tol m-CH), 7.24 (t, 1H, Ph p-CH), 7.13 (t, 2H, pyridine CH(3,5)), 7.11 (t, 2H, Ph m-CH), 6.93 (t, 1H, Ph p-CH), 3.36 (s, 3H, -OCH3), 2.42 (s, 3H, p-Tol

p-CH3), 2.34 (dd, 1H, CH), 1.25 (m, 1H, CH2a), 1.06 (m, 1H, CH2b), 0.70 (t, 3H, CH3). 1H-13C HSQC NMR

(500 MHz, CD2Cl2, 25 °C): δ 152.30 (pyridine CH(2,6)), 137.34 (pyridine CH(4)), 128.90 (p-Tol m-CH),

128.21 (Ph m-CH), 128.10 (Ph' m-CH), 126.09 (Ph p-CH), 125.67 (Ph' p-CH), 125.03 (Ph o-CH), 124.36 (p-Tol o-CH), 124.34 (pyridine CH(3,5)), 123.19 (Ph' o-CH), 49.67 (-OCH3), 31.27 (CαH), 24.03 (CβH2),

(14)

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

Processed on: 2-2-2021 PDF page: 215PDF page: 215PDF page: 215PDF page: 215

205

NMR reactivity of Pd-1c with styrene, CO and NaBPh4 in THF-d8.

NM

NM NM

NMRRR RRrerererer acacacacatitititivivivivivtytytytyoooof fff fPdPdPdPd--1c--1c1c1c wiwiwiwiwthtthththsssstytytytyyrerererennenenene, ,,,COCOCOCOaaaandndndndNNNNNaBaBaBaBPhPhPhPh4 4444ininininnTTTTHFHFHFHFF----ddddddddddd8888. ...

In a glovebox, compound Pd-1c (3.6 mg, 5.1·10-3 mmol) was dissolved in a J. Young’s NMR tube in

THF-d8. To the solution was added 4 equivalents of styrene (2.9 μL, 2.6 mg, 25.3·10-3 mmol)and the

reaction was monitored via 1H NMR spectroscopy. No changes were observed over the course of 29

hours. The solution was frozen in liquid N2 and the J. Young’s NMR tube was evacuated at the Schlenk

line. Subsequently, the NMR tube was allowed to warm to room temperature and then was filled with 1 bar of CO, resulting in a color change to a darker shade of blue. The addition of CO led to the clean formation of compound Pd-1e which was not influenced by the presence of styrene. Reaction of the in situ formed compound Pd-1e with styrene was followed by NMR spectroscopy at − 40 °C for 30 minutes, followed by another 30 minutes at RT, which did not show any insertion of styrene. Subsequently, 1 equivalent of NaBPh4 (1.8 mg, 5.3·10-3 mmol) was added to the J. Young’s NMR tube

to abstract the chloride ligand from Pd-1e, after which the CO atmosphere was restored. The 1H NMR

spectrum taken immediately after showed decomposition of Pd-1e to the homoleptic compound

Pd-1e and formation of palladium black was observed; no indication of olefin insertion was observed

in the 1H NMR spectra.

8

8.5.4 Attempted chloride abstraction from Pd-1c

Attempts to study the formation and decomposition of the putative three-coordinate intermediate [Pd(L1)Me] were carried out by treating Pd-1c with reagents that could abstract Cl. In all cases, a

solution of Pd-1c was prepared in THF-d8 and cooled to − 30 °C in the glovebox freezer in a

Teflon-sealed NMR tube. Subsequently, the chloride abstraction reagent was added and the sample was quickly frozen in liquid N2 before being inserted into the cold NMR probe (at − 40 °C).

Reaction of Pd-1c with AgBF4

Treatment of Pd-1c with AgBF4 proceeds rapidly at − 40 °C. A 1H NMR spectrum taken immediately

upon mixing at − 40 °C shows that a Pd-Me resonance is no longer observed. A new singlet at 0.85 ppm appears which is attributed to the formation of ethane (Figure 8.6). A plausible explanation for these observations is that AgBF4 reacts with Pd-1c by electron transfer (rather than Cl− abstraction) producing

Ag0 and the neutral radical [Pd(L1)MeCl], which subsequently results in homolysis of the Pd-Me bond

to form CH3• and the (dimeric) complex [Pd(μ-Cl)(L1)]2 (

Scheme 8.2), for which there is precedent in related E-diiminate chemistry.27 The presence of

additional, very broad signals in the in situ 1H NMR spectra is indicative of (unknown) paramagnetic

impurities, consistent with a homolytic bond cleavage (radical) pathway. A solid sample of [Pd(μ-Cl)(L1)]2 was obtained by precipitation, but NMR spectroscopy showed it was contaminated with

tetrabutylammonium salts. The amount of Bu4N+ is variable between batches (based on 1H NMR

integration, see Figure 8.7), suggesting that the [Pd(L1)]-containing fragment is not ionic, but rather the neutral complex [Pd(μ-Cl)(L1)]2.

Experimental procedure for in situ NMR reaction of Pd-1c with AgBF4 in THF-d8.

In glovebox, to a cold (− 30 ˚C) solution of Pd-1c (5.5 mg, 8.0·10-3 mmol) in THF-d8, in a Young’s NMR

tube, AgBF4 (excess) was added as a solid. The sample was quickly frozen in liquid nitrogen and

subsequently inserted into an NMR probe pre-cooled to − 40 °C. The reaction was followed for 1 day at − 40 °C and then was warmed up to + 25 °C. The volatiles were removed and the solid was redissolved in THF-d8 to confirm the disappearance of the peak of ethane at 0.85 ppm. Filtration of a THF through

(15)

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

Processed on: 2-2-2021 PDF page: 216PDF page: 216PDF page: 216PDF page: 216

206

[Pd(μ-Cl)(L1)]2, as a mixture with a Bu4N+ salt (the ratio between [Pd(μ-Cl)(L1)]2 and Bu4N+ is dependent

on the batch, see Figure 8.7).

[Pd(μ-Cl)(L1)]2. 1H NMR (400 MHz, THF-d8, 25 °C): δ 8.40 (d, 8H, Ph o-CH), 7.94 (d, 4H p-Tol o-CH), 7.53

(t, 8H Ph m-CH), 7.37 (t, 4H Ph p-CH), 7.28(d, 4H p-Tol m-CH) 2.40 (s, 6H, p-Tol CH3) ppm.

Figure 8.6. 1H NMR spectra of Pd-1c (THF-d8, 500 MHz): a) at – 40 °C, b) + AgBF4 at – 40 °C, c) + AgBF4 at + 25 °C.

Figure 8.7. 1H NMR spectrum of the isolated solid, [Pd(μ-Cl)(L1)]

2, from reaction of Pd-1c with AgBF4 (THF-d8, 400

MHz; a) and b) spectra show different batches highlighting the varying amount of Bu4N+ present).

Pd-1c THF-d8 THF-d8 Pd-1c Pd-CH3 NBu4+ Pd-1c [Pd(μ-Cl)(L1)]2 THF-d8 THF-d8 [Pd(μ-Cl)(L1)]2 C2H6 Pd-1c a) b) c) [Pd(μ-Cl)(L1)]2 THF-d8 THF-d8 [Pd(μ-Cl)(L1)]2 NBu4+ a) b)

(16)

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

Processed on: 2-2-2021 PDF page: 217PDF page: 217PDF page: 217PDF page: 217

207

Reaction of Pd-1c with NaBPh4

The analogous reaction of Pd-1c with NaBPh4 does not proceed at − 40 °C, but increasing the

temperature gave a complicated mixture in which Pd-1a is ultimately the major product in solution (Scheme 8.4, Figure 8.8). However, also the formation of a palladium mirror is observed, and a white solid is precipitated (NaCl).

Scheme 8.4. Proposed reaction sequence for in situ NMR reactivity of Pd-1c + NaBPh4 in THF-d8.

Experimental procedure for in situ NMR reaction of Pd-1c with NaBPh4 in THF-d8.

In a glovebox, to a cold (− 30 ˚C) solution of Pd-1c (1 eq, 5.5 mg, 8.0·10-3 mmol) in THF-d8, in a Young’s

NMR tube, NaBPh4 (1 eq, 2.7 mg, 8.0·10-3 mmol) was added as a solid. The sample was quickly frozen

in liquid nitrogen and subsequently inserted into an NMR probe pre-cooled to − 40 °C. The reaction was followed for 2 h at − 40 °C (no reaction was observed), for 2 days at + 25 °C and then the sample was warmed up to +75 °C for 5 days after which a dark green-brownish solution was observed together with formation of a Pd(0) mirror and precipitation of a white solid.

Figure 8.8. 1H NMR spectra of Pd-1c (THF-d8, 500 MHz): a) at – 40 °C; b) + NaBPh4 at – 40 °C t= 1h 30 min; c) + NaBPh4

at + 25 °C t=1 day; d) + NaBPh4 after heating the sample at + 75 °C for 5 days, spectrum recorded at + 25 °C.

Pd-1c THF-d 8 THF-d8 Pd-1c NBu4+ Pd-1c Pd-1c Pd-CH3 BPh4 -Pd-1a BPh4 -Pd-1a a) b) c) d)

(17)

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

Processed on: 2-2-2021 PDF page: 218PDF page: 218PDF page: 218PDF page: 218

208

Reaction of Pd-1c with B(C6F5)3

Chloride abstraction using the Lewis acid B(C6F5)3 at − 40 °C immediately forms Pd-1a as the major

product in solution (Scheme 8.5). Concomitantly, a significant amount of blue solid is precipitated which was insoluble in common (organic) solvents including MeOH, DMSO, mineral acid and the ionic liquid imidazolium salt. 19F NMR analysis of the brown supernatant solution shows that the borane is

converted to a mixture of chloro- and methylborate ([B(C6F5)3Cl]- and [B(C6F5)3Me]-, assigned by

comparison to literature NMR data)36 in a ratio of ca. 2:1, respectively (Figure 8.9).

Scheme 8.5. Proposed reaction sequence for in situ NMR reactivity of Pd-1c + B(C6F5)3 in THF-d8.

In situ NMR reactivity of Pd-1c with B(C6F5)3 in THF-d8.

In a glovebox, to a solution of B(C6F5)3 (1 eq, 4.1 mg, 8.0·10-3 mmol) in THF-d8, in a Young’s NMR tube,

Pd-1c (1 eq, 5.5 mg, 8.0·10-3 mmol) was added as a solid, leading to a dark brown reaction mixture and

precipitation of a dark blue solid. The reaction was followed via 1H, 19F, 1B NMR spectroscopy at + 25

°C. The 1H NMR spectrum recorded after 20 min shows that compound Pd-1c has almost completely

reacted, leading to compound Pd-1a as major species in addition to some traces amount of unidentified formazanate species and free ligand. In addition to this, in the aliphatic region the resonances of NBu4+ (1H NMR: 3.21, 1.64, 1.37, 0.97 ppm, in THF-d8) and of methane (1H NMR: 0.19

ppm in THF-d8) and ethane (1H NMR: 0.85 ppm in THF-d8) are present. The 19F and 11B NMR spectra

recorded after 20 min show a mixture of B(C6F5)3, [B(C6F5)3Cl]− and [B(C6F5)3CH3]−.

B(C6F5)3.19F NMR (376 MHz, THF-d8, 25 °C): − 131.8 (6H, d, o-F), − 156.5 (3H, t, p-F), − 163.3 (6H, t, m-F) ppm. 11B NMR (128 MHz, THF-d8, 25 °C): + 4.4 ppm. [B(C6F5)3Cl]−.19F NMR (376 MHz, THF-d8, 25 °C): − 130.4 (6H, m, o-F), − 162.5 (3H, t, p-F), − 166.8 (6H, m, m-F) ppm. 11B NMR (128 MHz, THF-d 8, 25 °C): − 7.6 ppm. [B(C6F5)3CH3]−.1H NMR (400 MHz, THF-d8, 25 °C): 0.5 (br, B-CH3) ppm. 19F NMR (376 MHz, THF-d8, 25 °C): − 130.9 (6H, m, o-F), − 165.2 (3H, t, p-F), − 167.3 (6H, m, m-F) ppm. 11B NMR (128 MHz, THF-d8, 25 °C): − 14.9 ppm.

(18)

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

Processed on: 2-2-2021 PDF page: 219PDF page: 219PDF page: 219PDF page: 219

209

Figure 8.9. 1H NMR spectra (THF-d8, 400 MHz) of: a) Pd-1c, b) Pd-1c + B(C6F5)3 t = 20 min.

Figure 8.10. 19F NMR spectra (THF-d8, 376 MHz) of: a) B(C6F5)3; b) Pd-1c + B(C6F5)3 t = 20 min. Pd-1c THF-d8 THF-d8 Pd-1c NBu4+ Pd-1c Pd-1c Pd-CH3 Pd-1a Pd-1a Pd-1a Pd-1a Pd-1a Pd-1a Pd-1a C2H6 [B(C6F5)3Me] -CH4 THF-d8 THF-d8 a) b) B(C6F5)3 [B(C6F5)3Cl]- [B(C6F5)3Me] -a) b)

(19)

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

Processed on: 2-2-2021 PDF page: 220PDF page: 220PDF page: 220PDF page: 220

210

Figure 8.11. 11B NMR spectra (THF-d

8, 128 MHz) of: a) B(C6F5)3 (THF adduct); b) Pd-1c + B(C6F5)3 t= 20 min.

8

8.5.5 DFT calculations

Computational studies used the Gaussian09 software package.37 Geometry optimizations were carried

out for the bis(formazanate) palladium compound Pd-1a using DFT calculations with B3LYP functional and a 6-31+G(d,p) basis set for all atoms except Pd, for which a LANL2DZ basis set (with ECP) was used. Optimized geometries were verified to be minima on the potential energy surface by frequency calculations. Starting from the geometry of Pd-1a also the geometries of the 1-electron reduced compound Pd-1a− and the 1-electron oxidized compound Pd-1a+ were optimized (as doublet states

using unrestricted DFT). Visualization of the molecular orbitals and spin density distributions was performed using Chemcraft 1.7 or Gaussview 5.0.

8.5.6 UV-Vis absorption spectroscopy

Figure 8.12. a) UV-Vis absorption spectra of compounds Pd-1a, Pd-1c, and Pd-1d in CH2Cl2. b) Physical appearance

of the CH2Cl2 solutions of compounds Pd-1a, Pd-1c, and Pd-1d.

B(C6F5)3 [B(C6F5)3Me] -[B(C6F5)3Cl] -a) b) -1000 1000 3000 5000 7000 9000 11000 13000 15000 380 580 780 ε (M -1·c m -1) Wavelength (nm) PdL2 [Pd(L)(CH3)Cl][NBu4] Pd(L)(CH3)(py) Pd-1a Pd-1c Pd-1d Pd-1a Pd-1c Pd-1d a) b)

(20)

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

Processed on: 2-2-2021 PDF page: 221PDF page: 221PDF page: 221PDF page: 221

211

8

8.6

References

1. (a) A. Nakamura, S. Ito and K. Nozaki, Chem. Rev., 2009, 109, 5215-5244; (b) B. P. Carrow and K. Nozaki, Macromolecules, 2014, 47, 2541-2555.

2. L. K. Johnson, C. M. Killian and M. Brookhart, J. Am. Chem. Soc., 1995, 117, 6414-6415.

3. (a) S. D. Ittel, L. K. Johnson and M. Brookhart, Chem. Rev., 2000, 100, 1169-1204; (b) Z. Guan and C. S. Popeney, in Metal Catalysts in Olefin Polymerization, ed. Z. Guan, Springer Berlin Heidelberg, Berlin, Heidelberg, 2009, pp. 179-220; (c) L. Guo, S. Dai, X. Sui and C. Chen, ACS Catalysis, 2016, 6, 428-441; (d) L. Guo, W. Liu and C. Chen, Mater. Chem Front., 2017, 1, 2487-2494.

4. (a) J. Durand and B. Milani, Coord. Chem. Rev., 2006, 250, 542-560; (b) A. Meduri, T. Montini, F. Ragaini, P. Fornasiero, E. Zangrando and B. Milani, ChemCatChem, 2013, 5, 1170-1183; (c) F. Amoroso, E. Zangrando, C. Carfagna, C. Müller, D. Vogt, M. Hagar, F. Ragaini and B. Milani, Dalton Trans., 2013, 42, 14583-14602; (d) V. Rosar, A. Meduri, T. Montini, F. Fini, C. Carfagna, P. Fornasiero, G. Balducci, E. Zangrando and B. Milani, ChemCatChem, 2014, 6, 2403-2418; (e) V. Rosar, T. Montini, G. Balducci, E. Zangrando, P. Fornasiero and B. Milani, ChemCatChem, 2017, 9, 3402-3411.

5. A. Nakamura, T. M. J. Anselment, J. Claverie, B. Goodall, R. F. Jordan, S. Mecking, B. Rieger, A. Sen, P. W. N. M. van Leeuwen and K. Nozaki, Acc. Chem. Res, 2013, 46, 1438-1449.

6. T. R. Younkin, E. F. Connor, J. I. Henderson, S. K. Friedrich, R. H. Grubbs and D. A. Bansleben, Science,

2000, 287, 460-462.

7. D. Zhu and P. H. M. Budzelaar, Dalton Trans., 2013, 42, 11343-11354.

8. B.-L. Lin, K. X. Bhattacharyya, J. A. Labinger and J. E. Bercaw, Organometallics, 2009, 28, 4400-4405. 9. X. Tian, R. Goddard and K.-R. Pörschke, Organometallics, 2006, 25, 5854-5862.

10. (a) A. Hadzovic, J. Janetzko and D. Song, Dalton Trans., 2008, 3279-3281; (b) A. Hadzovic and D. Song, Inorg. Chem., 2008, 47, 12010-12017; (c) V. T. Annibale, L. M. Lund and D. Song, Chem. Commun., 2010, 46, 8261-8263; (d) V. T. Annibale, R. Tan, J. Janetzko, L. M. Lund and D. Song, Inorg. Chim. Acta, 2012, 380, 308-321.

11. (a) A. W. Nineham, Chem. Rev., 1955, 55, 355-483; (b) G. I. Sigeikin, G. N. Lipunova and I. G. Pervova, Russ. Chem. Rev., 2006, 75, 885-900.

12. (a) N. Frolova, S. Vatsadze, V. Zavodnik, R. Rakhimov and N. Zyk, Russ. Chem. Bull., 2006, 55, 1810-1818; (b) J. B. Gilroy, B. O. Patrick, R. McDonald and R. G. Hicks, Inorg. Chem., 2008, 47, 1287-1294. 13. (a) A. R. Siedle and L. H. Pignolet, Inorg. Chem., 1980, 19, 2052-2056; (b) J. B. Gilroy, M. J. Ferguson, R.

McDonald and R. G. Hicks, Inorg. Chim. Acta, 2008, 361, 3388-3393; (c) A. V. Zaĭdman, I. G. Pervova, Z. G. Rezinskikh, I. N. Lipunov and P. A. Slepukhin, Cryst. Reports, 2010, 55, 424-427.

14. (a) M.-C. Chang, T. Dann, D. P. Day, M. Lutz, G. G. Wildgoose and E. Otten, Angew. Chem. Int. Ed., 2014, 53, 4118-4122; (b) M. C. Chang and E. Otten, Chem. Commun., 2014, 50, 7431-7433; (c) M.-C. Chang and E. Otten, Inorg. Chem., 2015, 54, 8656-8664; (d) 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; (e) M. C. Chang, A. Chantzis, D. Jacquemin and E. Otten, Dalton Trans., 2016, 45, 9477-9484; (f) R. Mondol, D. A. Snoeken, M.-C. Chang and E. Otten, Chem. Commun., 2017, 53, 513-516.

15. (a) J. B. Gilroy, M. J. Ferguson, R. McDonald, B. O. Patrick and R. G. Hicks, Chem. Commun., 2007, 126-128; (b) S. M. Barbon, J. T. Price, P. A. Reinkeluers and J. B. Gilroy, Inorg. Chem., 2014, 53, 10585-10593; (c) R. R. Maar, S. M. Barbon, N. Sharma, H. Groom, L. G. Luyt and J. B. Gilroy, Chem. Eur. J., 2015, 21, 15589-15599; (d) M. Hesari, S. M. Barbon, V. N. Staroverov, Z. Ding and J. B. Gilroy, Chem. Commun.,

2015, 51, 3766-3769; (e) S. M. Barbon, V. N. Staroverov and J. B. Gilroy, Angew. Chem. Int. Ed., 2017,

56, 8173-8177.

16. E. Kabir, C.-H. Wu, J. I. C. Wu and T. S. Teets, Inorg. Chem., 2016, 55, 956-963.

17. D. L. J. Broere, B. Q. Mercado, J. T. Lukens, A. C. Vilbert, G. Banerjee, H. M. C. Lant, S. H. Lee, E. Bill, S. Sproules, K. M. Lancaster and P. L. Holland, Chem. Eur. J., 2018, 24, 9417-9425.

18. R. Travieso-Puente, M.-C. Chang and E. Otten, Dalton Trans., 2014, 43, 18035-18041. 19. unpublished work.

20. (a) M. Rudler-Chauvin and H. Rudler, J. Organomet. Chem., 1977, 134, 115-119; (b) S. R. Foley, R. A. Stockland, H. Shen and R. F. Jordan, J. Am. Chem. Soc., 2003, 125, 4350-4361.

21. A more detailed analysis of the spectral data is provided in the Supporting Information.

22. (a) E. Kogut, A. Zeller, T. H. Warren and T. Strassner, J. Am. Chem. Soc., 2004, 126, 11984-11994; (b) H. L. Wiencko, E. Kogut and T. H. Warren, Inorg. Chim. Acta, 2003, 345, 199-208.

(21)

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

Processed on: 2-2-2021 PDF page: 222PDF page: 222PDF page: 222PDF page: 222

212

24. D. Guironnet, P. Roesle, T. Rünzi, I. Göttker-Schnetmann and S. Mecking, J. Am. Chem. Soc., 2009, 131, 422-423.

25. T. Kochi, K. Yoshimura and K. Nozaki, Dalton Trans., 2006, 25-27.

26. (a) C. Amatore and A. Jutand, Acc. Chem. Res, 2000, 33, 314-321; (b) M. Kolter, K. Böck, K. Karaghiosoff and K. Koszinowski, Angew. Chem. Int. Ed., 2017, 56, 13244-13248.

27. A. Hadzovic and D. Song, Organometallics, 2008, 27, 1290-1298. 28. H. Shen and R. F. Jordan, Organometallics, 2003, 22, 1878-1887.

29. (a) U. Frey, L. Helm, A. E. Merbach and R. Romeo, J. Am. Chem. Soc., 1989, 111, 8161-8165; (b) R. Romeo, A. Grassi and L. Monsu Scolaro, Inorg. Chem., 1992, 31, 4383-4390; (c) J. A. Casares, S. Coco, P. Espinet and Y.-S. Lin, Organometallics, 1995, 14, 3058-3067; (d) A. L. Casado, J. A. Casares and P. Espinet, Inorg. Chem., 1998, 37, 4154-4156; (e) O. F. Wendt, R. J. Deeth and L. I. Elding, Inorg. Chem.,

2000, 39, 5271-5276; (f) M. R. Plutino, L. M. Scolaro, A. Albinati and R. Romeo, J. Am. Chem. Soc., 2004,

126, 6470-6484; (g) S. Akita, R. Nakano, S. Ito and K. Nozaki, Organometallics, 2018, 37, 2286-2296. 30. S. Mecking, L. K. Johnson, L. Wang and M. Brookhart, J. Am. Chem. Soc., 1998, 120, 888-899.

31. D. Guironnet, L. Caporaso, B. Neuwald, I. Göttker-Schnetmann, L. Cavallo and S. Mecking, J. Am. Chem. Soc., 2010, 132, 4418-4426.

32. G. Tian, H. W. Boone and B. M. Novak, Macromolecules, 2001, 34, 7656-7663.

33. (a) T. Vlaar, E. Ruijter, B. U. W. Maes and R. V. A. Orru, Angew. Chem. Int. Ed., 2013, 52, 7084-7097; (b) G. Qiu, Q. Ding and J. Wu, Chem. Soc. Rev., 2013, 42, 5257-5269; (c) V. P. Boyarskiy, N. A. Bokach, K. V. Luzyanin and V. Y. Kukushkin, Chem. Rev., 2015, 115, 2698-2779.

34. J. A. Olson, T. A. P. Paulose, P. Wennek, J. W. Quail and S. R. Foley, Inorg. Chem. Commun., 2008, 11, 1297-1299.

35. R. E. Rulke, J. M. Ernsting, A. L. Spek, C. J. Elsevier, P. W. N. M. van Leeuwen and K. Vrieze, Inorg. Chem.,

1993, 32, 5769-5778.

36. G. Bellachioma, G. Cardaci, E. Foresti, A. Macchioni, P. Sabatino and C. Zuccaccia, Inorg. Chim. Acta,

2003, 353, 245-252.

37. 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. J. Bearpark, J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, N. J. 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 09, Revision D.01, Gaussian, Inc., Wallingford, CT, USA, 2009.

Referenties

GERELATEERDE DOCUMENTEN

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

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

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

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

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

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

Single crystals suitable for X-ray crystallography of 1-ZnMe-5-ZnMe were obtained by slow diffusion of hexane into a solution of the corresponding LZnMe complex in toluene Figure

Figure 7.4 Left: magnetic susceptibility measurements for 1-Fe in the solid state and in solution THF and toluene; right: variable temperature UV/Vis spectra for 1-Fe in toluene in