You are on page 1of 8

Dalton

Transactions
View Article Online
PAPER View Journal | View Issue

Coordinatively- and electronically-unsaturated


Published on 02 November 2017. Downloaded by Universitat de Barcelona on 8/27/2020 5:18:33 PM.

Cite this: Dalton Trans., 2017, 46,


square planar cobalt(III) complexes of a pyridine
16228 dianionic pincer ligand†
Faidh Hana,a Alan J. Loughb and Gino G. Lavoie *a

A series of low-valent Co(III) square planar complexes supported by a dianionic pincer ligand bis(aryl-
amido)pyridine ([NNN]2−) was synthesized, including the first structurally characterized square planar
hydroxide complex ([NNN]Co(OH), 4) of a 3d transition metal in the +3 oxidation state. The magnetic pro-
perties of the Co(III) complexes were highly dependent on the charge of the complex and on the coordi-
nation environment of the metal. The diamagnetic cationic complex {[NNN]Co(py)}BF4 (2) can be con-
verted to neutral paramagnetic complexes [NNN]Co(OR) (R = Ph 3a, iPr 3b, and H 4) by a simple substi-
tution of the ancillary pyridine ligand in 2. The double bond character between the metal and the anilido
Received 29th September 2017, nitrogen was evident from short Co–N2,3 bond lengths in the solid-state structures, and further sup-
Accepted 2nd November 2017
ported by density functional theory calculations. Complex 2 showed well-behaved redox processes at
DOI: 10.1039/c7dt03674d −0.93 and +0.70 V assigned to CoII/III and [NNN]2−/1− redox couples, respectively. In contrast, both com-
rsc.li/dalton plexes 3a and 4 showed some irreversibility in redox processes on either cobalt or ligands.

Introduction Lu),4–11 with only one report on Fe, Co, and Ni complexes.12
An analogous ligand with an isopropylene linkage between the
Bis(amido)pyridine A (Fig. 1) is a class of pincer ligands based central pyridine ring and the anilide groups was reported and
on a pyridine backbone that is flanked by two anilide groups discovered serendipitously by Wang13 while studying the tri-
through an alkyl linker, first utilized by McConville in 1995 as methylaluminum complexes of bis(imino)pyridine, which
a rigid dianionic tridentate ligand to support Ti,1 Zr,2 and Ta3 undergoes methyl insertions to give B (Fig. 1). This family of
complexes. Interestingly, the coordination chemistry of bis ligands A and B is highly rigid, coordinating to the metal
(amido)pyridine has been limited to early-transition metals centers in a meridional fashion, and allows for easy variations
(Sc, Y and Re) and to lanthanides/actinides (La, Th, Nd, and of its steric and electronic properties. These bis(amido)pyri-
dine ligands are formally 6-electron σ-donors, with additional
π-electron density donation available from the anilide nitrogen
atoms. This electron-donating ability of the ligand offers the
opportunity to stabilize electron-poor metal centers in coordi-
natively-unsaturated complexes.
We became interested in studying the coordination chem-
istry of these ligands to cobalt(III) complexes for potential use
Fig. 1 Structures of the dianionic pincer ligands A and B. in olefin polymerization. [NNN]CoR presents interesting simi-
larities to the bis(imino)pyridine cobalt(II) systems, which are
highly active in olefin polymerization. [NNN]CoR closely
a
Department of Chemistry, York University, 4700 Keele Street, Toronto, Ontario, resembles the catalytically-active species obtained by the acti-
M3J 1P3, Canada. E-mail: glavoie@yorku.ca
b vation of bis(imino)pyridine Co(II)Cl2 with aluminoxanes,
X-Ray Laboratory, Department of Chemistry, University of Toronto, 80 St George
Street, Toronto, Ontario, M5S 1A1, Canada without being positively charged, thus decreasing the propen-
† Electronic supplementary information (ESI) available: Crystallographic data for sity of functional groups present in monomers such as acry-
2, 3a and 4 in CIF and other electronic format, including tables of crystal data lates from impeding and even poisoning catalysis.14,15 Gibson
and structure refinement, bond lengths, angles, atomic coordinates and equi- further suggested that the propagating species could in fact
valent isotropic displacement parameters, and anisotropic displacement para-
involve a Co(III) dialkyl zwitterionic complex.16 Organocobalt(III)
meters. NMR spectra for 2, 3a, 3b and 4. DFT optimized coordinates and mole-
cular orbital diagrams for 2, 3a and 4. CCDC 1571906–1571908. For ESI and crys- systems have also been shown to be active in olefin polymeriz-
tallographic data in CIF or other electronic format see DOI: 10.1039/c7dt03674d ation, even in the absence of a cocatalyst.17 [NNN]CoR com-

16228 | Dalton Trans., 2017, 46, 16228–16235 This journal is © The Royal Society of Chemistry 2017
View Article Online

Dalton Transactions Paper

plexes are also excellent candidates in controlled radical


polymerization,18 and in olefin hydrogenation, a transform-
ation that continues to be of great interest thanks to recent
developments in the use of earth-abundant metal-based
catalysts.19
We thus herein report the synthesis of cobalt(III) cationic
and neutral square planar complexes of the tridentate [NNN]2−
ligand, including the first isolated and structurally-character-
Published on 02 November 2017. Downloaded by Universitat de Barcelona on 8/27/2020 5:18:33 PM.

ized 3d metal hydroxide complex. To minimize the possible


undesired participation of the ligand in β-hydride elimination,
our initial study focused on the use of ligand class B. Features
from the solid-state structures of the isolated complexes are Scheme 2 Synthesis of complexes 3 and 4.
presented, along with their electrochemical behavior and
results from density functional theory calculations.

relatively sharp resonances (average ν1/2 (Hz) for all methyl


Results and discussion resonances: 6.7 (3a) and 31 (3b)), and chemical shifts ranging
from 26 to −10 ppm for 3a, and from 47 to −11 ppm for 3b.
Preparation and reactivity of cobalt(III) complexes
Single crystals of 3a suitable for X-ray diffraction analysis were
The synthesis of complex 1 was adopted from that reported by obtained by liquid diffusion of pentane into a dichloro-
Summerscales and Gordon,12 where an alkyl lithium base was methane solution of 3a at low temperatures. The crystals of
added to a cold suspension of H2[NNN] and Co(II)Cl2py4. The complex 3b could not be isolated due to reactivity towards
1
H NMR spectrum of the paramagnetic complex 1 showed four adventitious water. While analysis of the reaction mixture by
(20.6, 17.1, 1.1 and −4.4 ppm) very broad resonances. Complex 1
H NMR spectroscopy (see the ESI†) supported the formation
1 was readily oxidized by AgBF4 to give the blue diamagnetic of 3b, attempts to isolate the product through standard wash-
coordinatively-unsaturated square planar Co(III) complex 2 ings and extractions led to the formation of the paramagnetic
(Scheme 1).30 The resonances in the 1H and 13C NMR spectra hydroxo complex 4, which was independently prepared by the
of 2 were unambiguously assigned by 2D correlation experi- reaction of 2 with either sodium or cesium hydroxide.
ments. The fluoride nucleus in the tetrafluoroboride countera- Complex 4 could also be generated by protonolysis of 3a
nion resonates at −152.5 ppm in the 19F NMR spectrum. with water. The UV-visible spectrum of 4 resembled that of the
Complex 2 is insoluble in water but soluble and stable in Co(III) complexes 2 and 3a. The resonances observed in the
protic (methanol) and polar organic (dichloromethane) sol- relatively sharp (average ν1/2 for all methyl resonances: 13 Hz)
vents. The complex is relatively stable in air, with minimal paramagnetically-shifted (δ 48 to −11 ppm) 1H NMR spectrum
decomposition observed after prolonged exposure of up to 4 of 4 are consistent with the expected structure. The hydroxy
weeks. resonance was however not observed. The X-ray diffraction ana-
The synthesis of the corresponding neutral alkyl and lysis of complex 4 corroborated the structure and the coordi-
hydride complexes was explored by reactions of 2 with metal nation of the hydroxy ligand to the metal. The hydrogen atom
alkyl and hydride reagents, such as AlMe3, CH2CHCH2MgCl, was successfully located on the differential electron density
CH3MgBr, LiCH2Si(CH3)3, and LiHB(CH2CH3)3. These reac- map, eliminating the need to use the calculated position in a
tions gave a reduced product, presumably through homolytic riding model to confirm the molecular structure (vide infra).
cleavage of the metal–hydrogen or metal–carbon bond in the A search of the Cambridge Crystallographic Data Center
new transiently-formed complex, resulting in a red para- (CCDC) database revealed that complex 4 represents the first
magnetic complex, consistent with 1 based on spectroscopic structurally characterized cobalt(III) hydroxide complex and
data.31–34 This was further supported by the independent syn- the first square planar first-row transition metal in the +3
thesis of complex 1 by the chemical reduction of 2 with KC8. oxidation state with a terminal hydroxide.
Reactions of 2 with sodium phenoxide and isopropoxide to The spin-only magnetic moment for 3a and 4, respectively,
produce coordination complexes [NNN]Co(III)OR (R = Ph 3a was determined in THF-d8 at room temperature by the Evans
and iPr 3b) were explored (Scheme 2). Complexes 3a and 3b method and found to be of 2.4 and 2.7µB, consistent with two
displayed a paramagnetically-shifted 1H NMR spectrum with unpaired electrons. Complex 3b could not be isolated in a pure
enough form to allow for magnetic susceptibility measure-
ments. The paramagnetism of 3 based on Co(II) oxyl radical
species is ruled out based on crystallographic35–37 and
spectroscopic38–40 data. Complexes 3a and 4 (and most likely
3b) thus adopt a low-lying triplet ground state (S = 1, as pre-
dicted by DFT calculations, vide infra) at room temperature,
Scheme 1 Synthesis of complex 1 and 2. commonly observed in square planar Co(III) complexes.41–43

This journal is © The Royal Society of Chemistry 2017 Dalton Trans., 2017, 46, 16228–16235 | 16229
View Article Online

Paper Dalton Transactions

X-ray studies
X-ray diffraction analysis on single crystals was used to deter-
mine the solid-state structure for complexes 2, 3a and 4. All
complexes adopt a slightly-distorted square planar geometry in
the solid state, with four-coordinate geometry index (τ4) values
of 0.09, 0.10, and 0.11, respectively (Fig. 2).44 The cobalt is
positioned within 0.03 (2, 4) to 0.09 Å (3a) from the mean
plane formed by the three nitrogen atoms of the [NNN]2−
Published on 02 November 2017. Downloaded by Universitat de Barcelona on 8/27/2020 5:18:33 PM.

ligand. The sum of the angles of the amido nitrogen atoms


ranges from 359.0 to 360.0°, highlighting the trigonal planar geo-
metry and the sp2 hybridization of N2,3, which allows for the
enhanced stability of these otherwise electronically-unsaturated
metal centers through the donation of π-electron density to
empty d-orbitals on the metal (vide infra). The cobalt amide bond
lengths (Co–N) in 2 are unusually short, with an average of
1.792 Å. These bond lengths are in fact the shortest cobalt(III)–
amide bonds reported in the CCDC database. Co–NR2 single
bond lengths normally range from 1.87 to 1.92 Å,45–50 while
CovNR double bond lengths range from 1.66 to 1.69 Å.51–53
The coordinating amido groups enforce a much shorter
Co–N1[NNN] pyridine bond length in 2 (1.829(3) Å), compared to
Co–N4pyridine (1.938(3) Å) (Table 1). The corresponding Co–N1
bond in 3a remains unchanged at 1.829(2) Å but elongates to
1.8573(16) Å in 4. Likewise, the Co–NR2 bonds in 2 and 3a are
statistically equivalent, but they are elongated in 4. The Co–O1
and O1–C28 bond lengths in 3a are 1.8984(17) Å and 1.329(3)
Å, respectively, well within the literature values of Co-bonded
anionic phenoxide ligands.37,39,40,54,55 The phenoxide is
almost perpendicular (85°) to the mean plane formed by N1,
N2 and N3, with a C28–O1–Co–N2 torsional angle of 81°. The
Co–O1 bond length (1.7945(15) Å) in complex 4 is shorter than
that reported in other 3d transition metal complexes contain-
ing a terminal hydroxide ligand.56 The planes formed by xylyl
groups in 2, 3a, and 4 are 80–86° to the plane formed by the
[NNN]2− ligand.

Cyclic voltammetry
Cyclic voltammetry of the cationic complex 2 in dichloro-
methane shows a reversible one-electron CoII/III couple at
−0.93 V, and a ligand-based ([NNN]2−/[NNN]1−) quasi-revers-
ible one-electron couple at 0.70 V (Fig. 3). The metal-based
redox couple in the neutral complex 3a occurs at a higher
negative potential (−1.47 V), with two oxidation peaks (Epa) at
−0.02 V and 0.89 V, centered on the organic fragments.
Complex 4 shows an irreversible reduction of Co(III) at −1.82 V
and a [NNN]2− ligand-based reversible one-electron redox Fig. 2 ORTEP representations of 2 (top), 3a (middle), and 4 (bottom)
process at 0.00 V. DFT calculations (vide infra) predict that the (50% probability level). Most hydrogen atoms, the BF4 counteranion in 2
redox events at Epa 0.80 (2) and 0.05 V (4) occur on the [NNN]2− and solvent molecules in 3a and 4 were removed for clarity. Selected
ligand itself, while the irreversible anodic peaks for 3a involve bond lengths and angles are presented in Table 1.

both the phenoxide (Epa −0.02 V) and [NNN]2− (Epa 0.89 V)


ligands. Thus, the potentials for the oxidation of the [NNN]2−
ligand in 2 ([NNN]Co( py)1+) and [3a]+ ([NNN]Co(OPh)1+) are Density functional theory calculations
within 0.09 V of each other and, as expected based on the Density functional theory (DFT) calculations were performed
greater charge of the complexes, approximately 0.80 V higher on complexes 2, 3a and 4. The geometry of these complexes
than that for 4 ([NNN]Co(OH)). was optimized at the B3LYP level of theory using the LANL2DZ

16230 | Dalton Trans., 2017, 46, 16228–16235 This journal is © The Royal Society of Chemistry 2017
View Article Online

Dalton Transactions Paper

Table 1 Selected bond lengths (Å) and bond angles (°) for complexes 2,
3a and 4

Complexes 2 3a 4

Co–N1 1.829(3) 1.829(2) 1.8573(16)


Co–N2 1.788(3) 1.790(2) 1.8363(16)
Co–N3 1.796(2) 1.796(2) 1.8392(16)
Co–N4 1.938(3)
Co–O1 1.8984(17) 1.7945(15)
Published on 02 November 2017. Downloaded by Universitat de Barcelona on 8/27/2020 5:18:33 PM.

N1–Co–N2 84.86(13) 84.93(10) 84.16(7)


N1–Co–N3 84.58(12) 84.73(10) 83.64(7) Fig. 4 π-(HOMO−10) and π*-(LUMO) molecular orbitals for the Co–
N2–Co–N3 169.42(13) 168.52(10) 167.71(7) N2,3 bonds in complex 2 (isovalue of 0.04 e− Å−3).
N1–Co–N4 178.33(13)
N1–Co–O1 177.03(9) 177.17(7)
Co–O1–C28 120.86(16)
solid-state structure of 2 (Fig. 4). The [NNN]2− contributes at
least 85% of the electron density for the highest occupied
molecular orbitals HOMO to HOMO−4, with the electron
density in the HOMO to the HOMO−2 solely localized on the
xylyl groups, predicting possible oxidation of the ligand as
observed in the cyclic voltammetry.
The composition of the two lowest singly unoccupied mole-
cular orbitals (LSUMO and LSUMO+1) in complex 3a is split
approximately equally between the cobalt atom (55%) and the
[NNN]2− ligand (45%). The LSUMO+1 is a π*-orbital arising
from dxz,pz-interactions between cobalt (dxz) and nitrogen ( pz),
with the corresponding (also antibonding) singly occupied
molecular orbital HSOMO−3 (q = 0.81). The short Co–N bond
Fig. 3 Cyclic voltammograms of 2 (dotted), 3a (dashed), and 4 (solid lengths observed in the solid-state structure of 3a are further
line) in dichloromethane containing 0.1 M Bu4N(PF6), using ferrocene as supported by other π-bonding interactions (HSOMO−15 and
the internal standard (E1/2 = 0.00 V; scan rate 100 mV s−1). HSOMO−53; see the ESI†). The two highest singly occupied
molecular orbitals (HSOMO and HSOMO−1), with an energy
difference of only 1.3 kcal mol−1, are related to each other (q =
basis set, and calculations were performed in the singlet state 1.0) and principally localized (96%) on the phenoxide ligand
and in the unrestricted triplet state. Energies were calculated (Fig. 5). The next four lower energy α-orbitals matched with
on the optimized structure with the triple-zeta valence polar- their β-orbitals with q > 0.90. Orbital sets HSOMO−5/
ized basis set (def-TZVP). Similar methods have been used to HSOMO−4 and HSOMO−9/HSOMO−8 are mainly composed
study Co systems.37,57 These calculations indicate that the of the [NNN]2− fragment (>94%), while HSOMO−6/HSOMO−7
triplet state is more energetically favored than the singlet state and HSOMO−11/HSOMO−10 are composed of both [NNN]2−
by approximately 7 kcal mol−1 (3a) and 16 kcal mol−1 (3b, 4). and −OPh fragments in ratios of approximately 30 : 70 and
In contrast, both spin states for the cationic complex 2 have 75 : 25, respectively. These data support the ability of both
approximately the same energy (within 0.1 kcal mol−1). These [NNN]2− and −OPh to easily get oxidized, as observed in the
calculations thus support the observed diamagnetism of cyclic voltammetry.
complex 2 and paramagnetism of complexes 3 and 4.
The composition of the frontier orbitals in the molecular
orbital diagram of complexes 2, 3a and 4 was analyzed using
the AOMix software28,29 where the complexes were split into
three fragments: Co3+, [NNN]2−, and the second donor ( py in
2, −OPh in 3a, or −OH in 4). AOMix28,29 was also used to match
the α- and β-orbitals of the paramagnetic complexes and quan-
tify the quality of the match with the resemblance factor q on a
scale from 0 ( poor) to 1 (excellent; see the ESI†).
The lowest unoccupied molecular orbital (LUMO) of
complex 2 consists of a π*-antibonding interaction between
the dxz orbital on cobalt and pz orbitals on the amide nitrogen
atoms. The corresponding π(dxz,pz)-bonding interaction corres-
ponds to the HOMO−10, indicating a Co–N bond order of 1.5,
consistent with the short Co–N2,3 bond length observed in the Fig. 5 HSOMO and HSOMO−1 of complex 3a (0.045 e− Å−3).

This journal is © The Royal Society of Chemistry 2017 Dalton Trans., 2017, 46, 16228–16235 | 16231
View Article Online

Paper Dalton Transactions

As for complex 4, the lowest unoccupied orbitals LSUMO+1 alkyl complexes could not be isolated, presumably due to their
and LSUMO (q = 0.67) are composed of 98% and 58%, respect- high reactivity and homolysis of the new bond formed. Our
ively, of the pyridine π-system from the [NNN]2− ligand. As present work is focused on utilizing these complexes in
observed in 3a, the HSOMO and LSUMO+2 (q = 0.81) of polymerization of alkenes and lactides, and in hydrogenation
complex 4 are also π-antibonding, arising from interactions of alkenes. The results will be reported in due course.
between dxz,pz-atomic orbitals. As noted for 2 and 3a,
additional π-bonding molecular orbitals (HSOMO−11 and
HSOMO−38) result in short Co–N bond lengths (see the ESI†). Experimental
Published on 02 November 2017. Downloaded by Universitat de Barcelona on 8/27/2020 5:18:33 PM.

Orbital sets HSOMO−3/HSOMO−2, HSOMO−4/HSOMO−5,


General considerations
HSOMO−8/HSOMO−7, and HSOMO−10/HSOMO−9 are com-
posed of more than 95% of the [NNN]2− fragment with most of Unless otherwise noted, all manipulations were carried out
the electron density again localized on the xylyl rings. In con- using oven-dried glassware and performed under argon using
trast, the electron density in the orbital set HSOMO−6/ standard Schlenk line techniques or in a nitrogen-filled
HSOMO−1 exhibits significant amido N pz-character. The DFT mBraun glovebox. Solvents were dried using an mBraun
calculations thus proved helpful in interpreting the observed solvent purification system (MB-SPS) fitted with alumina
electrochemistry for complexes 2, 3a and 4. columns and stored over molecular sieves under argon.
All complexes show a large Co contribution (>50% for 2 and Deuterated NMR solvents were purchased from Sigma-Aldrich,
3a, 35% for 4) to the lowest unoccupied molecular orbitals degassed using three freeze–pump–thaw cycles, vacuum dis-
indicating that cathodic peaks at −0.93 (2), −1.47 (3a) and tilled from Na/benzophenone (C6D6) and CaH2 (CD2Cl2), and
−1.82 V (4) correspond to the reduction of Co(III) to Co(II). In stored under argon or nitrogen. NMR spectra were recorded on
contrast, the highest occupied molecular orbitals are generally a Bruker DRX 600, Bruker AV 400 or Bruker AV 300 spectro-
localized on either the [NNN]−2 or −OPh/−OH ligands with meter at room temperature. CoCl2py4 20 was synthesized
small (less than 15%) Co contribution. Orbitals with >15% Co according to the literature procedures. Sodium isopropoxide
contribution are at least 18 kcal mol−1 lower in energy than and sodium phenoxide were prepared from the corresponding
the highest occupied molecular orbitals (HOMO in 2 or alcohol and sodium metal. Bu4NPF6, CoCl2 and AgBF4, and
HSOMO in 3a and 4), indicating that the redox couples at 0.70 LiCH2SiMe3 (1.0 M in pentane) were purchased from Sigma-
(2) and 0.00 V (4) and anodic peaks at −0.02 and 0.89 V (3a) Aldrich and used as received.
are all principally ligand-based. Specifically, these ligand-
based activities are localized on the xylyl ring for complex 2, Synthetic procedures
and on either the xylyl ring or the amido groups for complex 4. [NNN]Co( py) (1). Complex 1 was prepared following an
The HSOMO and HSOMO−1 orbitals of complex 3a are loca- experimental procedure reported by Summerscales and
lized on the aromatic phenoxide ring, with the xylyl groups Gordon for a related complex.12 To a 50 mL toluene suspen-
heavily contributing to the HSOMO−2 to the HSOMO−5. This sion of H2[NNN] (2.23 g, 5.00 mmol) and CoCl2py4 (2.01 g,
suggests that the anodic peaks at −0.02 and 0.89 V correspond 5.00 mmol) cooled to −40 °C was added LiCH2SiMe3 (1.0 M in
to the oxidation of the phenoxide phenyl ring and of the xylyl pentane; 10.5 mL, 10.5 mmol), resulting in a color change
group of the [NNN]2− ligand, respectively. from cyan to red. After stirring overnight at room temperature,
the mixture was filtered using a cannula filter, and the volatiles
were evaporated under vacuum. The crude red solids were
Conclusions washed with cold pentane and dried under vacuum. Yield:
1.84 g, 3.45 mmol, 69%.
A series of low valent Co(III) square planar complexes sup- {[NNN]Co( py)}BF4 (2). A 20 mL toluene solution of complex
ported by a dianionic pincer bis(arylamido)pyridine ligand was 1 (1.84 g, 3.43 mmol) was added to a 20 mL toluene suspen-
synthesized and characterized, including an unprecedented sion of AgBF4 (0.667 g, 3.43 mmol) at room temperature and
and first ever structurally-characterized square planar Co(III) the mixture was stirred for 2 h. The supernatant was removed,
hydroxide complex. The magnetic properties of these com- and the residual solids were washed with toluene. The product
plexes were found to be highly dependent on the charge and was extracted with dichloromethane, and the volatiles from
nature of the coordination sphere imparted by the second the solution removed in vacuo to yield a dark blue solid
ligand, chosen from either pyridine, phenoxide or hydroxide. (1.90 g, 3.09 mmol, 90%). Crystals for X-ray crystallography
While the cationic pyridine complex 2 is diamagnetic, com- were grown by liquid diffusion of pentane into a dichloro-
plexes 3 and 4 exhibit paramagnetism with two unpaired elec- methane solution of 2. 1H NMR (400 MHz; CD2Cl2): δ 9.60 (d,
3
trons. The double bond character between the metal and the J = 5.2 Hz, 2H, o-CHpy), 8.54 (t, 3J = 8.0 Hz, 1H, p-CH) 7.79 (d,
3
anilido nitrogen was evident from X-ray diffraction analysis J = 7.9 Hz, 2H, m-CH), 7.72 (t, 3J = 7.5 Hz, 1H, p-CHpy), 7.15
and further supported by DFT calculations for complex 2. (dd, 2H, m-CHpy), 6.77 (t, 3J = 7.6 Hz, 2H, p-CHxylyl ), 6.45 (d,
3
Complex 2 showed well-behaved redox chemistry. In contrast, J = 7.4 Hz, 4H, m-CHxylyl ), 1.90 (s, 12H, CH3(xylyl)), 1.29 (s, 12H,
both complexes 3a and 4 showed irreversible redox processes C(CH3)2). 13C{1H} NMR (100 MHz, CD2Cl2): δ 172.4 (o-C), 155.4
on either cobalt or ligands. The corresponding hydride and (o-Cpy), 152.2 (o-Cxylyl ), 140.8 ( p-C), 139.1 ( p-Cpy), 128.3 (m-

16232 | Dalton Trans., 2017, 46, 16228–16235 This journal is © The Royal Society of Chemistry 2017
View Article Online

Dalton Transactions Paper

Cxylyl ), 127.1 (ipso-Cxylyl ), 125.5 ( p-Cxylyl, and m-Cpy), 119.7 coated with AgCl was used as the reference electrode. All
(m-C), 91.8 (C(CH3)2), 23.0 (C(CH3)2), 21.5 (CH3xylyl ). 19F NMR potentials were calibrated vs. the Fc+/Fc redox couple. Half-
(376 MHz; CD2Cl2): δ −152.5 (s, BF4). Anal. calcd for wave potential (E1/2) was estimated as the average of anodic
C32H38N4CoBF4: C, 61.55; H, 6.13; N, 8.97. Found: C, 61.29; H, (Epa) and cathodic peak (Epc) potentials.
5.87; N, 8.58.
[NNN]Co(OPh) (3a). A 5 mL dichloromethane solution of 2 Computational details
(106 mg, 0.170 mmol) was added to a 5 mL dichloromethane Geometry optimizations were carried out at the B3LYP21–24
suspension of sodium phenoxide (22.6 mg, 0.185 mmol) hybrid level of theory with the LANL2DZ basis set using
Published on 02 November 2017. Downloaded by Universitat de Barcelona on 8/27/2020 5:18:33 PM.

resulting in a slight colour change to a lighter blue. The Gaussian 09,25 with GaussView 0326 for molecular visualiza-
mixture was stirred at room temperature for 1 h, filtered and tion. These calculations were performed on the Shared
dried in vacuo. The product was isolated by extraction with Hierarchical Academic Research Computing Network
toluene. The volatiles were removed in vacuo and the blue solid (SHARCNET: http://www.sharcnet.ca). The starting geometries
was further washed with pentane (5 × 25 mL). Yield 64 mg, were based on X-ray structures but further optimized in the
0.12 mmol, 68%. Crystals for X-ray crystallography were grown gas phase. Energy calculations were performed at the same
from pentane liquid diffusion into dichloromethane solution functional using the triple-zeta valence polarized basis set
stored at −40 °C. 1H NMR (400 MHz; C6D6): δ 26.49 (d, 3J = (def-TZVP).27 The composition of molecular orbitals as combi-
7.1 Hz, 4H, m-CHxylyl ), 13.31 (dd, 2H, m-CHPh), 10.73 (t, 3J = nations of atomic orbitals was determined based on data
7.1 Hz, 2H, p-CHxylyl ), 3.59 (t, 3J = 6.8 Hz, 1H, p-CHPh), 2.94 (d, generated by Gaussian 09 using AOMix software.28,29
3
J = 7.2 Hz, 2H, m-CH), 2.60 (t, 3J = 7.0 Hz, 1 H, p-CH), 2.21 (s,
12H), −3.91 (s, 12H), −9.74 (d, 3J = 7.0 Hz, 2H, o-CHPh). X-ray crystallography
13
C{1H} NMR (100 MHz; C6D6): δ 246.2, 200.5, 172.5, 137.1
Crystallographic data were collected at the University of
126.8, 108.8, −22.1. Magnetic moment (THF, 0.05 M): 2.4 B.M.
Toronto on a Bruker Kappa APEX-DUO diffractometer using
Anal. calcd for C33H38CoN3O: C, 71.85; H, 6.94; N, 7.62. Found:
monochromated Mo-Kα radiation (Bruker Triumph; λ =
C, 71.59; H, 7.17; N, 7.38.
0.71073 Å) at 150 K. Data were measured using a combination
[NNN]Co(OiPr) (3b). Complex 3b was prepared using the pro-
of ϕ scans and ω scans, and were processed using APEX2 and
cedures described for 3a on a 0.167 mmol scale. The formation
SAINT.58 Absorption corrections were carried out using
of 3b is consistent with 1H NMR spectroscopic data but could
SADABS.58 Structures were solved with WinGX59 or OLEX260 as
not be cleanly isolated due to rapid decomposition. 1H NMR
the graphical user interface using either SHELXL or SHELXT
(400 MHz; C6D6): δ 47.29 (br, 4H, m-CHxylyl ), 35.06 (br, 1H,
with SHELXS-97 for full-matrix least-squares refinement that
OCH(CH3)2), 15.72 (br, 2H, p-CHxylyl ), 6.44 (br, 6H, OCH
was based on F2.61–63 All H atoms were included in calculated
(CH3)2), 0.42 (br, 2H, m-CH), 0.04 (12H), −3.90 (br, 1H, p-CH),
positions, except for the hydroxy hydrogen atom in 4, and
−10.90 (12H).
allowed to refine in riding-motion approximation with Uiso
[NNN]Co(OH) (4). To a 5 mL dichloromethane solution of
tied to the carrier atom.
complex 2 (79 mg, 0.13 mmol) was added excess NaOH pellets
(or CsOH·H2O), with no attempt to exclude moisture or air
from the reaction mixture. The reaction mixture was stirred
vigorously for 1 h at room temperature. The mixture was
Conflicts of interest
filtered, and the solvent from the filtrate evaporated. The There are no conflicts to declare.
product was extracted using toluene. The solvent was evapor-
ated and the product was washed with warm hexanes to give 4
as a purple solid (47 mg, 0.10 mmol, 77%). Crystals for X-ray
crystallography were grown from a dichloromethane/pentane
Acknowledgements
mixture stored at 0 °C. 1H NMR (400 MHz; C6D6): δ 48.58 (d, G. G. L. gratefully acknowledges financial support from the
3
J = 7.3 Hz, 4H, m-CHxylyl ), 14.90 (t, 3J = 7.2 Hz, 2H, p-CHxylyl ), Natural Sciences and Engineering Research Council for a
2.71 (s, 12H), 0.21 (d, 3J = 7.0, 2H, m-CH), −9.98 (s, 12H), Discovery Grant. F. H. wishes to thank Enbridge and the
−11.04 (t, 3J = 6.9 Hz, 1H, p-CH). 13C NMR (150 MHz; C6D6): Government of Ontario for scholarships. The authors wish to
δ 322.8, 297.2, 244.1, 128.9, 91.5, −63.5, −110.0. Magnetic thank Professor A. B. P. Lever and Dr Elaine Dodsworth for
moment (THF, 0.05 M): 2.7 B.M. Anal. calcd for C27H34CoN3O: helpful discussion of DFT calculations.
C, 68.05; H, 7.40; N, 8.82. Found: C, 68.31; H, 7.28; N, 8.55.

Cyclic voltammetry
References
Cyclic voltammograms were produced using a BASi-Epsilon
potentiostat on dichloromethane solutions of the analytes 1 F. Guérin, D. H. McConville and J. J. Vittal, Organometallics,
(0.01 M) containing 0.1 M Bu4NPF6 as the electrolyte. A plati- 1997, 16, 1491–1496.
num disc (1 mm diameter) and a platinum wire were used as 2 F. Guérin, D. H. McConville and J. J. Vittal, Organometallics,
working and auxiliary electrodes, respectively. A silver wire 1996, 7333, 5586–5590.

This journal is © The Royal Society of Chemistry 2017 Dalton Trans., 2017, 46, 16228–16235 | 16233
View Article Online

Paper Dalton Transactions

3 F. Guérin, D. H. McConville and J. J. Vittal, Organometallics, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar,


1995, 14, 3154–3156. J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene,
4 F. Estler, G. Eickerling, E. Herdtweck and R. Anwander, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo,
Organometallics, 2003, 22, 1212–1222. R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin,
5 C. A. Cruz, D. J. H. Emslie, L. E. Harrington, J. F. Britten R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin,
and C. M. Robertson, Organometallics, 2007, 26, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador,
692–701. J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas,
6 C. A. Cruz, D. J. H. Emslie, H. A. Jenkins and J. F. Britten, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox,
Published on 02 November 2017. Downloaded by Universitat de Barcelona on 8/27/2020 5:18:33 PM.

Dalton Trans., 2010, 39, 6626–6628. Gaussian 09, Revision D.01, Wallingford CT, 2013.
7 S. Jie and P. L. Diaconescu, Organometallics, 2010, 29, 26 R. D. Ik, T. Keith, J. Millam, K. Eppinnett, I. W. Hovell and
1222–1230. R. Gilliland, GaussView 3, Carnegie PA, 2003.
8 C. P. Lilly, P. D. Boyle and E. A. Ison, Dalton Trans., 2011, 27 F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys., 2005,
40, 11815–11821. 7, 3297.
9 C. P. Lilly, P. D. Boyle and E. A. Ison, Organometallics, 2012, 28 S. I. Gorelsky and A. B. P. Lever, J. Organomet. Chem., 2001,
31, 4295–4301. 635, 187–196.
10 H. Sugiyama, S. Gambarotta, G. P. A. Yap, D. R. Wilson and 29 S. I. Gorelsky, AOMix: Program for Molecular Orbital
S. K.-H. Thiele, Organometallics, 2004, 23, 5054–5061. Analysis, Version 6.88, University of Ottawa, Ottawa,
11 M. Zimmermann, F. Estler, E. Herdtweck, K. W. Törnroos Canada, 2013.
and R. Anwander, Organometallics, 2007, 26, 6029–6041. 30 J. Y. Jeon, J. J. Lee, J. K. Varghese, S. J. Na, S. Sujith,
12 O. T. Summerscales, J. A. Stull, B. L. Scott and J. C. Gordon, M. J. Go, J. Lee, M. Ok and B. Y. Lee, Dalton Trans., 2013,
Inorg. Chem., 2015, 54, 6885–6890. 42, 9245–9254.
13 B. Y. Tay, C. Wang, S. C. Chia, L. P. Stubbs, P. K. Wong and 31 S. Maria, H. Kaneyoshi, K. Matyjaszewski and R. Poli,
M. Van Meurs, Organometallics, 2011, 30, 6028–6033. Chem. – Eur. J., 2007, 13, 2480–2492.
14 V. C. Gibson, M. J. Humphries, K. P. Tellmann, D. F. Wass, 32 M. D. Fryzuk, D. B. Leznoff, R. C. Thompson and
A. J. P. White and D. J. Williams, Chem. Commun., 2001, S. J. Rettig, J. Am. Chem. Soc., 1998, 120, 10126–10135.
120, 2252–2253. 33 J. Setsune, Y. Ishimaru, T. Moriyama and T. Kitao, J. Chem.
15 T. M. Kooistra, Q. Knijnenburg, J. M. M. Smits, Soc., Chem. Commun., 1991, 555.
A. D. Horton, P. H. M. Budzelaar and A. W. Gal, Angew. 34 M. S. Ram, C. G. Riordan, G. P. A. Yap, L. Liable-Sands,
Chem., Int. Ed., 2001, 113, 4719–4722. A. L. Rheingold, A. Marchaj and J. R. Norton, J. Am. Chem.
16 M. J. Humphries, K. P. Tellmann, V. C. Gibson, Soc., 1997, 119, 1648–1655.
A. J. P. White and D. J. Williams, Organometallics, 2005, 24, 35 A. Sokolowski, E. Bothe, E. Bill, T. Weyhermüller,
2039–2050. K. Wieghardt, H. Stratemeier, D. Reinen and P. F. Knowles,
17 O. Daugulis, M. Brookhart and P. S. White, Chem. Commun., 1996, 28, 1671–1672.
Organometallics, 2003, 22, 4699–4704. 36 B. Adam, E. Bill, E. Bothe, B. Goerdt, G. Haselhorst,
18 A. Debuigne, R. Poli, C. Jérôme, R. Jérôme and K. Hildenbrand, A. Sokolowski, S. Steenken,
C. Detrembleur, Prog. Polym. Sci., 2009, 34, 211–239. T. Weyhermüller and K. Wieghardt, Chem. – Eur. J., 1997, 3,
19 S. A. M. Smith, P. O. Lagaditis, A. Lüpke, A. J. Lough and 308–319.
R. H. Morris, Chem. – Eur. J., 2017, 23, 7212–7216. 37 A. S. Roy, N. Muresan, H. M. Tuononen, S. P. Rath and
20 D. Zhu, F. F. B. J. Janssen and P. H. M. Budzelaar, P. Ghosh, Dalton Trans., 2008, 3438.
Organometallics, 2010, 29, 1897–1908. 38 A. Kochem, H. Kanso, B. Baptiste, H. Arora, C. Philouze,
21 P. J. Stephens, F. J. Devlin, C. F. Chabalowski and O. Jarjayes, H. Vezin, D. Luneau, M. Orio and F. Thomas,
M. J. Frisch, J. Phys. Chem., 1994, 98, 11623–11627. Inorg. Chem., 2012, 51, 10557–10571.
22 S. H. Vosko, L. Wilk and M. Nusair, Can. J. Phys., 1980, 58, 39 A. Sokolowski, B. Adam, T. Weyhermüller, A. Kikuchi,
1200–1211. K. Hildenbrand, R. Schnepf, P. Hildebrandt, E. Bill and
23 C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. K. Wieghardt, Inorg. Chem., 1997, 36, 3702–3710.
Matter, 1988, 37, 785–789. 40 L. Benisvy, E. Bill, A. J. Blake, D. Collison, E. S. Davies,
24 A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652. C. D. Garner, C. I. Guindy, E. J. L. McInnes, G. McArdle,
25 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, J. McMaster, C. Wilson and J. Wolowska, Dalton Trans.,
M. A. Robb, J. R. Cheeseman, V. Scalmani, G. Barone, 2004, 83, 3647.
B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, 41 P. O. Lagaditis, B. Schluschaß, S. Demeshko, C. Würtele
X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, and S. Schneider, Inorg. Chem., 2016, 55, 4529–4536.
J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, 42 M. A. García-Monforte, I. Ara, A. Martín, B. Menjón,
J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, M. Tomás, P. J. Alonso, A. B. Arauzo, J. I. Martínez and
H. Nakai, T. Vreven, J. A. Montgomery Jr., J. E. Peralta, C. Rillo, Inorg. Chem., 2014, 53, 12384–12395.
F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, 43 E. D. McKenzie and J. M. Worthington, Inorg. Chim. Acta,
K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, 1976, 16, 9–15.

16234 | Dalton Trans., 2017, 46, 16228–16235 This journal is © The Royal Society of Chemistry 2017
View Article Online

Dalton Transactions Paper

44 L. Yang, D. R. Powell and R. P. Houser, Dalton Trans., 2007, 53 X. Hu and K. Meyer, J. Am. Chem. Soc., 2004, 126, 16322–
96, 955–964. 16323.
45 F. N. Penkert, T. Weyhermüller, E. Bill, P. Hildebrandt, 54 C. N. Verani, S. Gallert, E. Bill, T. Weyhermüller,
S. Lecomte and K. Wieghardt, J. Am. Chem. Soc., 2000, 122, K. Wieghardt and P. Chaudhuri, Chem. Commun., 1999, 38,
9663–9673. 1747–1748.
46 R. E. Cowley, R. P. Bontchev, J. Sorrell, O. Sarracino, 55 G. M. Zats, H. Arora, R. Lavi, D. Yufit and L. Benisvy,
Y. Feng, H. Wang and J. M. Smith, J. Am. Chem. Soc., 2007, Dalton Trans., 2011, 40, 10889–10896.
129, 2424–2425. 56 C. Bergquist, T. Fillebeen, M. M. Morlok and G. Parkin,
Published on 02 November 2017. Downloaded by Universitat de Barcelona on 8/27/2020 5:18:33 PM.

47 S. Fortier, J. J. Le Roy, C. H. Chen, V. Vieru, M. Murugesu, J. Am. Chem. Soc., 2003, 125, 6189–6199.
L. F. Chibotaru, D. J. Mindiola and K. G. Caulton, J. Am. 57 L. Chiang, L. E. N. Allan, J. Alcantara, M. C. P. Wang, T. Storr
Chem. Soc., 2013, 135, 14670–14678. and M. P. Shaver, Dalton Trans., 2014, 43, 4295–4304.
48 C. Y. Lin, J. C. Fettinger, F. Grandjean, G. J. Long and 58 Bruker AXS, APEX2, SAINT, SADABS, Madison, Wisconsin,
P. P. Power, Inorg. Chem., 2014, 53, 9400–9406. USA, 2007.
49 M. J. Ingleson, M. Pink, H. Fan and K. G. Caulton, Inorg. 59 L. J. Farrugia, J. Appl. Crystallogr., 2012, 45, 849–854.
Chem., 2007, 46, 10321–10334. 60 O. V. Dolomanov, L. J. Bourhis, R. J. Gildea, J. A. K. Howard
50 J. Du, L. Wang, M. Xie and L. Deng, Angew. Chem., Int. Ed., and H. Puschmann, J. Appl. Crystallogr., 2009, 42, 339–341.
2015, 54, 12640–12644. 61 G. M. Sheldrick, Acta Crystallogr., Sect. A: Found.
51 L. Zhang, Y. Liu and L. Deng, J. Am. Chem. Soc., 2014, 136, Crystallogr., 1990, 46, 467–473.
1–26. 62 G. M. Sheldrick, Acta Crystallogr., Sect. C: Struct. Chem.,
52 X.-N. Yao, J.-Z. Du, Y.-Q. Zhang, X.-B. Leng, M.-W. Yang, 2015, 71, 3–8.
S.-D. Jiang, Z.-X. Wang, Z.-W. Ouyang, L. Deng, B.-W. Wang 63 G. M. Sheldrick, Acta Crystallogr., Sect. A: Found.
and S. Gao, J. Am. Chem. Soc., 2017, 139, 373–380. Crystallogr., 2008, 64, 112–122.

This journal is © The Royal Society of Chemistry 2017 Dalton Trans., 2017, 46, 16228–16235 | 16235

You might also like