Abstract
Many coordination complexes and organometallic compounds with the 4d6 and 5d6 valence electron configurations have outstanding photophysical and photochemical properties, which stem from metal-to-ligand charge transfer (MLCT) excited states. This substance class makes extensive use of the most precious and least abundant metal elements, and consequently there has been a long-standing interest in first-row transition metal compounds with photoactive MLCT states. Semiprecious copper(I) with its completely filled 3d subshell is a relatively straightforward and well explored case, but in 3d6 complexes the partially filled d-orbitals lead to energetically low-lying metal-centered (MC) states that can cause undesirably fast MLCT excited state deactivation. Herein, we discuss recent advances made with isoelectronic Cr0, MnI, FeII, and CoIII compounds, for which long-lived MLCT states have become accessible over the past five years. Furthermore, we discuss possible future developments in the search for new first-row transition metal complexes with partially filled 3d subshells and photoactive MLCT states for next-generation applications in photophysics and photochemistry.
Introduction
Transition metal complexes with a metal-centered HOMO and a ligand-based LUMO can have MLCT excited states with lifetimes on the order of nano- to microseconds, long enough for luminescence or diffusion-controlled photochemical processes to occur. According to a photophysical rule of thumb established by exploring many precious metal-based complexes,1 the larger the energy gap between the lowest MLCT excited state and the electronic ground state, the longer the excited-state lifetime and the higher the luminescence quantum yield become. To obtain large energy gaps, chelating polypyridine ligands are often combined with second- or third-row transition metals having six electrons in their outermost d-subshell. These heavier d-metals embedded into an octahedral coordination environment of ligands with combined σ-donor and π-acceptor properties then adopt a low-spin d6 configuration, in which the HOMO corresponds to a triply degenerate set of d-orbitals (named t2g in Figure 1a) and a LUMO with antibonding character on the ligand’s π-system (called π* in Figure 1a). Typical representatives are RuII,2 ReI,3 OsII,4 and IrIII,5 all of which belong to the rarest and most expensive stable elements; hence there is considerable interest in alternatives from the first row of transition metals, which would be more abundant and cheaper.
Figure 1.
(a) Molecular orbital (MO) scheme for low-spin d6 metal complexes as typically encountered for RuII polypyridine, cyclometalated IrIII, and related precious metal-based compounds, simplified to octahedral symmetry. (b) Typical MO scheme for FeII polypyridines, simplified to Oh point group. (c) Single configurational coordinate diagram with the key electronic states in low-spin 3d6 complexes; Q is a nuclear coordinate, and E is relative energy.
FeII is the most obvious choice and by far the best investigated case among 3d6 compounds,6 in which photoactive MLCT excited states can in principle be installed, but this is very difficult.7 The main challenge is that the lowest MLCT excited states of FeII complexes very often deactivate on the femto- to picosecond time scales,8−10 making them unsuitable for many photophysical and photochemical applications. The underlying reason for this behavior is the decrease in ligand field strength when going from the second to the first row of the d-metals, and this in turn is a consequence of the more contracted nature of the 3d-orbitals compared to the 4d- or 5d-orbitals.11 In Figure 1b, this manifests in a situation in which the LUMO is no longer a ligand-based π* orbital but instead becomes metal-centered, comprised of a doubly degenerate set of d-orbitals with eg symmetry in octahedral ligand fields. In other words, the HOMO–LUMO transition is now metal-centered (MC) and the MLCT state is no longer the lowest electronically excited state, causing rapid MLCT excited state deactivation.12 The relaxation from an initially excited MLCT into a lower lying MC state is called internal conversion (IC), and its rate and efficiency depend on the relative energies of the involved states and on the activation barrier for the IC process. In the prototypical [Fe(bpy)3]2+ complex (bpy = 2,2′-bipyridine) (Figure 2a), IC is associated with an activation energy of only roughly 300 cm–1,13 and the MLCT state deactivates within 50–80 fs at room temperature.14,15 Such ultrafast relaxation is often illustrated in potential well diagrams (Figure 1c),7 in which the horizontal axis reflects the altering bonding situations in different electronic states caused by the redistribution of electron density between different orbitals. The lowest MC states of 3d6 metal complexes are strongly distorted with respect to the MLCT states and the electronic ground state,16 owing to the promotion of electrons from formally non-σ-bonding t2g-orbitals to antibonding eg-orbitals, thus leading to an elongation of the average metal–ligand bond distance. This property can make the activation barrier for IC from the lowest MLCT into MC states undesirably small. To maximize that activation barrier, it is helpful to establish strong ligand fields and rigid coordination environments, which shift the MC states to high energies (ideally above the lowest MLCT state) and furthermore allow only minimal displacements of the MC potential wells. Strong ligand fields can be obtained with strong σ-donors (which act on the metal eg-orbitals and raise their energy) and strong π-acceptors (which influence the metal t2g-orbitals and lower their energy), whereas rigidity is typically achievable with bi- or tridentate chelate ligands that offer limited degrees of conformational freedom when coordinated to the metal.17 Over the past few years, significant progress has been made along these lines, not just with 3d6 metal complexes, but also with other types of photoactive first-row transition metal compounds.18 As noted above, FeII has received most attention among 3d6 compounds,10 even though early studies already noted that Cr0 compounds are promising as 3d6 MLCT luminophores,19,20 but then it seems that isoelectronic alternatives to FeII were largely ignored for three decades. Over the past 5 years, new Cr0, MnI and CoIII complexes with photoactive MLCT excited states were discovered, and important progress on FeII complexes has been made. In the following, we focus on these key developments and provide a comparative discussion of the photophysical and photochemical characteristics of these four different, yet closely related 3d6 species.
Figure 2.
Polypyridine complexes of FeII: (a) [Fe(bpy)3]2+.14,15 (b) [Fe(dcpp)2]2+ complex (dcpp = 2,6-bis(2-carboxypyridyl)pyridine) with its particularly strongly π-accepting terdentate ligands.22 (c) [Fe(dcpp)(ddpd)]2+ complex (ddpd = N,N′-dimethyl-N,N′-dipyridine-2-yl-pyridine-2,6-diamine).23 (d) [Fe(dqp)2]2+ complex (dqp = 2,6-di(quinoline-8-yl)pyridine).24 (e) [Fe(pdmmi)2]2+ complex (pdmmi = 3,3′-pyridine-2,6-diyl(methylene)bis(1-methylimidazolylidene).24 (f) FeII is coordinated by the three bpy subunits of a cage-bpy ligand to afford the [Fe(cage-bpy)]2+ complex; two CuI ions can furthermore be coordinated by the four nitrogen atoms of both imine caps of the cage to yield the [FeCu2(cage-bpy)]2+ complex.9 (g) Trinuclear RuII-FeII-RuII compound.26 (h) High-spin [Fe(dftpy)2]2+ (dftpy = 6,6″-difluoro-2,2′:6′,2″-terpyridine), [Fe(dctpy)2]2+ (dctpy = 6,6″-dichloro-2,2′:6′,2″-terpyridine) and [Fe(dbtpy)2]2+ (dbtpy = 6,6″-dibromo-2,2′:6′,2″-terpyridine) complexes with 5/7MLCT excited states.27,28
Iron(II) Compounds
With classical polypyridines, the achievable ligand field strengths for octahedral FeII complexes are typically such that MC states depopulate MLCT excited states in ultrafast manner, and this is usually monitored by following excited state absorption (ESA) bands that are characteristic for the one-electron reduced polypyridine ligands.21 A particularly strongly π-accepting variant of a tridentate polypyridine came close to a point (Figure 2b) at which the order of the two lowest 3MC and 5MC excited states was reversed when compared to the majority of previously investigated compounds, yet MLCT deactivation remained ultrafast.22 When combining that strongly π-accepting chelate with a strongly electron-donating tridentate ligand, a heteroleptic push–pull type FeII complex (Figure 2c) with a similarly strong ligand field was obtained, but its MLCT lifetime could not be measured.23 The use of tridentate polypyridine ligands providing close to perfect octahedral coordination geometries gave access to FeII complexes (Figure 2d/e) with MLCT lifetimes approaching the ps time regime (Table 1).24 A particularly ambitious study reported on a macrocyclic molecular cage (Figure 2f), in which one FeII cation was coordinated by the three central bpy units.9,25 The MLCT decay of the resulting [Fe(cage-bpy)]2+ complex exhibits vibronic coherence, from which information concerning the structural distortions responsible for the internal conversion of the MLCT into the MC state can be gained. The coordination of CuI ions by four nitrogen atoms at the two caps of the cage rigidifies the overall molecular structure in [FeCu2(cage-bpy)]2+ and decelerates the MLCT decay from 110 fs to 2.6 ps (Table 1, entry 7). The authors noted that this represented the longest reported FeII 3MLCT lifetime in a pure polypyridine coordination environment at the time of publication.9 For a trinuclear RuII-FeII-RuII compound integrating a central [Fe(tpy)2]2+-like unit (Figure 2g), a decay with a time constant of 23 ps (Table 1, entry 8) was tentatively attributed to the relaxation of the lowest 3MLCT excited state into MC states.26
Table 1. MLCT Lifetimes (τMLCT) of FeII Complexes with Polypyridine Ligands in Solution at Room Temperature.
| entry | compound | molecular structure | τMLCT |
|---|---|---|---|
| 1 | [Fe(bpy)3]2+ | Figure 2a | 50–80 fs14,15 |
| 2 | [Fe(dcpp)2]2+ | Figure 2b | N/A22 |
| 3 | [Fe(dcpp)(ddpd)]2+ | Figure 2c | N/A23 |
| 4 | [Fe(dqp)2]2+ | Figure 2d | 450 fs24 |
| 5 | [Fe(pdmmi)2]2+ | Figure 2e | 0.8–1.5 ps24 |
| 6 | [Fe(cage-bpy)]2+ | Figure 2f | 110 fs9 |
| 7 | [FeCu2(cage-bpy)]2+ | Figure 2f | 2.6 ps9 |
| 8 | RuII–FeII-RuII | Figure 2g | 23 ps26 |
| 9 | [Fe(dftpy)2]2+a | Figure 2h | 14.0 ps28 |
| 10 | [Fe(dctpy)2]2+a | Figure 2h | 16.0 ps27,28 |
| 11 | [Fe(dbtpy)2]2+a | Figure 2h | 17.4 ps28 |
| 12 | [Fe(tpy)(CN)3]− | Figure 3a | ca. 10 ps29 |
| 13 | [Fe(bpy)(CN4)]2– | Figure 3b | 18 ps30,31 |
The indicated lifetime is for a 5/7MLCT excited state. The respective compounds have a high-spin d6 valence electron configuration in the ground state (5T2).
In a radically different approach, the MLCT lifetimes of [Fe(tpy)2]2+ derivatives were tuned by ligand halogenation (Figure 2h) and steric strain.27,28 With increasing halogen size, steric interactions between the two coordinated tpy ligands increasingly restrict the conformational degrees of freedom, which likely contributes to the MLCT lifetime elongation from 14.0 to 16.0 and 17.4 ps (Table 1, entries 9–11) when going from F to Cl and Br substituents. However, the repulsive interactions between the halogen atoms from one tpy ligand and the other coordinated tpy furthermore cause a weak ligand field; hence these compounds have a high-spin 5T2 ground state (instead of 1A1 as all other complexes in Figure 2). The photoactive MLCT excited states in this situation are either quintets or septets, different from the triplets commonly observed for low-spin FeII complexes. Thus, suitably designed coordination spheres around FeII can provide access to photochemistry via quintet MLCT excited states.
In combination with strong-field cyanido ligands, a tpy ligand (tpy = 2,2′:6′,2″-terpyridine) on FeII (Figure 3a) recently afforded a 3MLCT lifetime of ca. 10 ps (Table 1, entry 12).29 Earlier studies on the related [Fe(bpy(CN)4]2– complex (Figure 3b) revealed an MLCT lifetime of 18 ps (Table 1, entry 13), suggesting that cyanido ligands are helpful to decelerate relaxation to MC states,30,31 but they also tend to make the photophysics of the resulting complexes very susceptible to interactions with the solvent.32 Borylation of the peripheral nitrogen atoms of the cyanido ligands presumably further strengthens the ligand field,33 yet the detectable excited-state lifetime of 28 ps in [Fe(bpy)(BCF)4]2– (BCF = tris(pentafluorophenyl)) (Figure 3c) was attributed to an MC rather than an MLCT state.34 Further MLCT excited-state tuning in iron(II) cyanido complexes is possible through variation of the α-diimine ligands.35
Figure 3.
Heteroleptic FeII complexes with coordination environments containing a combination of polypyridine and cyanido ligands: (a) [Fe(tpy)(CN)3]−.29 (b) [Fe(bpy)(CN)4]2–.30,31 (c) [Fe(bpy)(BCF)4]2–.34
N-Heterocyclic carbene (NHC) ligands emerged as better alternatives to polypyridines as far as long 3MLCT lifetimes are concerned, largely owing due their stronger σ-donor properties.6,7,36,37 Following an initial study published in 2013,38 a series of conceptually related FeII compounds (Figure 4a-f) with chelating NHC ligands featuring MLCT lifetimes on the order of a few picoseconds were discovered (Table 2),39−45 which represented record lifetimes at that point.7 The only outlier in this series is [Fe(pbbi)2]2+ (Figure 4c) with its lifetime of merely 0.3 ps,38 which is likely the result of a weakened ligand field caused by the steric strain imposed by the tert-butyl substituents. The attachment of peripheral anthracenyl- or pyrenyl-substituents has no important effect on the 3MLCT lifetime (Figure 4g/h),46 because the triplet energies of the respective molecular entities are too high relative to the lowest 3MLCT state; hence triplet reservoir effects cannot be expected.47 In the very close to perfectly octahedral coordination environment of [Fe(dpmi)2]2+ (Figure 4i), a 3MLCT lifetime of 9.2 ps was obtained (Table 2, entry 9).48 The 3MLCT excited-state decay behavior has furthermore been investigated in what could be considered an FeII NHC–ZnII porphyrin dyad (Figure 4j), and a lifetime of 160 ps was reported (Table 2, entry 10).49 This dyad was furthermore reported to be emissive from an FeII NHC-based 3MLCT state.
Figure 4.

FeII complexes with N-heterocyclic carbene ligands: (a) [Fe(bpmi)2]2+ (pbmi = pyridine-2,6-diyl)bis(1-methyl-imidazol-2-ylidene).38 (b) [Fe(pbmbi)2]2+ (pbmbi = (pyridine-2,6-diyl)bis(1-methyl-benzimidazol-2-ylidene).41 (c) [Fe(pbbi)2]2+ (pbbi = pyridine-2,6-diyl)bis(1-tert-butyl)-imidazol-2-ylidene).38 (d) [Fe(cpbmi)2]2+ (cpbmi = (carboxypyridine-2,6-diyl)bis(1-methyl-imidazol-2-ylidene).40,42 (e) [Fe(cpbmbi)2]2+ (cpbmbi = (carboxypyridine-2,6-diyl)bis(1-methyl-benzimidazol-2-ylidene).7,41 (f) [Fe(pzbmi)2]2+ (pzbmi = (pyrazine-2,6-diyl)bis(1-methyl-imidazol-2-ylidene).44 (g) [Fe(bim-ant)2]2+ (bim-ant = (4-anthracene-pyridine-2,6-diyl)bis(1-methylimidazol-2-ylidene)).46 (h) [Fe(bim-pyr)2]2+ (bim-pyr = (4-pyrene-pyridine-2,6-diyl)bis(1-methylimidazol-2-ylidene)).46 (i) [Fe(dpmi)2]2+ (dpmi = di(pyridine-2-yl)(3-methylimidazol-2-yl)methane).48 (j) Dinuclear [FeNHC-ZnP]2+ compound.49
Table 2. MLCT Lifetimes (τMLCT) of FeII Complexes with NHC or Mesoionic Carbene Ligands in Solution at Room Temperature.
| entry | compound | molecular structure | τMLCT |
|---|---|---|---|
| 1 | [Fe(pbmi)2]2+ | Figure 4a | 9 ps38 |
| 2 | [Fe(pbmbi)2]2+ | Figure 4b | 16 ps41 |
| 3 | [Fe(pbbi)2]2+ | Figure 4c | 0.3 ps38 |
| 4 | [Fe(cpbmi)2]2+ | Figure 4d | 16.5–18 ps40,42 |
| 5 | [Fe(cpbmbi)2]2+ | Figure 4e | 26 ps7,41 |
| 6 | [Fe(pzbmi)2]2+ | Figure 4f | 21–25 ps44 |
| 7 | [Fe(bim-ant)2]2+ | Figure 4g | 13.4 ps46 |
| 8 | [Fe(bim-pyr)2]2+ | Figure 4h | 12.8 ps46 |
| 9 | [Fe(dpmi)2]2+ | Figure 4i | 9.2 ps48 |
| 10 | [FeNHC-ZnP]2+ | Figure 4j | 160 ps49 |
| 11 | [Fe(btz)3]2+ | Figure 5a | 528 ps50 |
| 12 | [Fe(btz)2(bpy)]2+ | Figure 5b | 7.6/13 ps30,57 |
| 13 | [Fe(btp)2]2+ | Figure 5c | 8.7 ps58 |
Mesoionic carbenes allowed an MLCT lifetime elongation up to 0.5 ns (Figure 5a),50 owing to a further increase in ligand field strength caused by their strong π-acceptor properties, while remaining similarly strong σ-donors as NHCs (Table 2, entry 11).51 The electron donating properties of these mesoionic carbenes are in fact sufficiently pronounced to make FeIII the most stable oxidation state in a homoleptic tris(bidentate) complex,52 and this led to the (presumably accidental) discovery of FeIII complexes with luminescent ligand-to-metal charge transfer (LMCT) excited states with lifetimes up to the nanosecond regime.53,54 The accessibility of both the FeII and FeIII oxidation states in the compound of Figure 5a furthermore gave access to an unusual operation mode for photoredox catalysis, in which the two different oxidation states were excited consecutively.55 Moreover, an uncommon symmetry-breaking charge separation reaction was observable after excitation of [Fe(btz)3]3+, leading to FeII and FeIV photoproducts, thus corresponding essentially to a photoinduced disproportionation reaction.56 Turning back to FeII 3MLCT excited states and their lifetimes, we note that a heteroleptic FeII complex with the same btz ligand (Figure 5b)30,57 and a tridendate ligand combining mesoionic carbene and pyridine ligand motifs (Figure 5c)58 give less spectacular results in the sense that the MLCT lifetime is much shorter (Table 2, entries 12–13).
Figure 5.
FeII complexes with mesoionic carbene ligands: (a) [Fe(btz)3]2+ (btz = 3,3′-dimethyl-1,1′-bis(p-tolyl)-4,4′-bis(1,2,3-triazol-5-ylidene)).50 (b) [Fe(btz)2(bpy)]2+.30,57 (c) [Fe(btp)2]2+ (btp = 4,4′-(pyridine-2,6-diyl)bis(1-ethyl-3-methyl-1,2,3-triazol-5-ylidene)).58
In a conceptually much different approach, tridentate ligands incorporating π-donating amido groups enabled strong mixing between the metal t2g orbitals and the filled ligand orbitals (Figure 6a), such that the resulting HOMO obtained substantial ligand character,59 an effect that had been predicted computationally.60,61 In combination with π-accepting phenanthridine and quinoline moieties in the same (tridentate) chelate (Figure 6b), this “HOMO inversion” effect led to absorption over large parts of the UV–vis spectrum.62 The lowest electronically excited states in this compound class (leading to absorption bands between 650 to 900 nm) involve the relocation of electron density from the π* (p+d) Namido-Fe antibonding orbitals (HOMO in Figure 6a) to phenanthridine-based π* orbitals, and were termed PALCT for “πantibonding-to-ligand charge transfer”.59 Transitions with cleaner MLCT character seem to occur at higher energy in those compounds. The 3PALCT excited states have lifetimes on the order of 2–3 ns in solution at room temperature (Table 3, entries 1–3). X-ray techniques confirmed the high degree of metal–ligand bond covalence in this compound class,63 and furthermore led to the insight that the ensuing low Racah B parameter overcompensates for the comparatively weak ligand field splitting (10 Dq) imparted by the π-donating amido units. Consequently, the 10 Dq/B ratio is larger for [Fe(pqa)2(Cl)] than for [Fe(tpy)2]2+. The combined effects of a stabilized PALCT state, minimal stabilization of the MC states, and a substantial inner-sphere reorganization energy cause a greater barrier for deactivation of the lowest charge-transfer excited state by MC states.63
Figure 6.
(a) Simplified molecular orbital diagram illustrating the situation with phenanthridine-based π* orbitals, metal-based t2g orbitals, and amido-centered N(2p) orbitals, leading to a mixed and destabilized HOMO.59 Molecular structures of: (b) [Fe(pqa)2(R)] (pqa = (phenanthridin-4-yl)(quinoline-8-yl)amido, R = tBu, CF3, Cl).59,62,63 (c) [Fe(pbpy)(tpy)]2+ (pbpy = 6-phenyl-2,2′-bipyridine).72 (d) [Fe(C^Npy^Npy)(6-NHC-bpy)]+ (HC^Npy^Npy = 6-(phenyl)-2,2′-bipyridine; 6-NHC-bpy = 6-(1-methyl-imidazol-2-ylidene)-2,2′-bipyridine.73 (e) [Fe(pphen)2] (pphen = 9-phenylphenanthroline).74
Table 3. Charge-Transfer Excited-State Lifetimes (τCT) of FeII Complexes with Amido or Cyclometalating Ligands in Solution at Room Temperature.
Cyclometalated FeII complexes have long received attention from a computational chemistry perspective,64−70 but air- and water-stable cyclometalated FeII compounds amenable to experimental photochemistry have been elusive.71 Recently, a robust heteroleptic FeII complex with one N^N^C coordinating ligand and one tpy chelate (Figure 6c), was found to have an MLCT lifetime of 0.8 ps (Table 3, entry 4),72 a factor of 5 longer than in the structurally related [Fe(tpy)2]2+. This finding is in line with earlier theoretical studies that considered N^N^Caryl ligands promising for elongating the MLCT lifetimes of FeII complexes,61,67,68 though a lifetime of 0.8 ps for [Fe(pbpy)(tpy)]2+ is short in comparison to 528 ps for [Fe(btz)3]2+ (Table 2, entry 11). The combination of a cyclometalating ligand with a bpy ligand bearing an additional NHC coordination site in the [Fe(C^Npy^Npy)(6-NHC-bpy)]+ complex (Figure 6d) led to a 3MLCT lifetime of 21.4 ps (Table 3, entry 5).73
9-Phenylphenanthroline was later used as a tridentate chelate ligand for a homoleptic FeII complex (Figure 6e) with a rigid N^N^C coordination environment that led to an MLCT excited state featuring a 1 ns lifetime in solution at room temperature (Table 3, entry 6).74 Photoluminescence in the NIR spectral range between 1170 and 1230 nm was reported for this compound, which seems very unusual given that in isoelectronic RuII and OsII compounds, the energy gap law typically makes MLCT emission in the NIR improbable,1 unless specifically tailored ligands (with particularly π-extended frameworks) are present.75 To date, this is the only mononuclear FeII complex for which an emissive MLCT excited state has been claimed, whereas selected earlier studies merely reported on ligand-based fluorescence from FeII compounds.76−78 As long as the respective ligand-centered (singlet) excited states are sufficiently decoupled from the metal-centered excited states, the occurrence of such fluorescence seems very plausible.
Chromium(0) Carbonyls and Isocyanides
To stabilize the low oxidation state of Cr0 and to obtain a low-spin 3d6 valence electron configuration in octahedral complexes, strongly π-accepting ligands such as carbonyls and isocyanides are well-suited. Studies performed in the 1970s already noted the similarity between the electronic structures of zerovalent group 6 metal isocyanide complexes (Figure 7a) and RuII polypyridines.19,79 For hexakis(arylisocyanide) complexes with the heavier homologues Mo0 and W0, emission in solution at room temperature was readily observable.19 Related studies reported on luminescence from [M(α-diimine)(CO)4] complexes (Figure 7b; M = Cr0, Mo0, W0), with some of the Cr0 compounds apparently featuring dual emission from two MLCT excited states in benzene at room temperature.20,80 Aside from this uncommon behavior in conflict with Kasha’s rule,81 later studies demonstrated that dissociation of CO ligands can be an important excited-state deactivation pathway in [Cr(α-diimine)(CO)4] complexes.82
Figure 7.

(a) Generic molecular structure of hexakis(arylisocyanide) complexes with low-valent metals (Cr0, MnI, Mo0, W0).19,79 (b) Generic molecular structure of [M(α-diimine)(CO)4] compounds (M = Cr0, Mo0, W0), here with α-diimine = 1,10-phenanthroline.20,80 (c) Molecular structures of Cr0 complexes with bidentate arylisocyanide ligands (R = H: LH, R = pyrenyl: LPy), emitting from 3MLCT excited states.83,85 (d) Competing MLCT decay pathways in [Cr(LPy)3] depending on solvent polarity.85 (e) Molecular structures of [Mn(Lbi)3]+ (Lbi = 2,5-bis(3,5-di-tert-butyl-2-isocyanophenyl)thiophene).90 (f) [Mn(Ltri)2]+ (Ltri = 5,5′-(2-isocyano-5-methyl-1,3-phenylene)bis(2-(3,5-di-tert-butyl-2-isocyanophenyl)thiophene)).90 (g) [Mn(CNdippPhOMe2)6]+ (CNdippPhOMe2 = 4-(3,5-dimethoxyphenyl)-2,6-diisopropylphenylisocyanide)),94 and (h) [Co(bimca)2]+ (bimca = 1,8-bis(imidazoline-2-yliden-1-yl)carbazolide).108
A chelating arylisocyanide ligand provided access to clear-cut 3MLCT emission in a homoleptic Cr0 complex with a tris(bidentate) coordination environment (Figure 7c, R = H).83 In dry and deaerated THF at room temperature, the 3MLCT lifetime was 2.2 ns (Table 4, entry 1), but the photoluminescence quantum yield (ϕ) was only 10–5, barely above the detection limit. A Mo0 complex with the exact same coordination environment showed 3MLCT photoluminescence with a quantum yield of 5.8 × 10–2 (5800 times higher) and an average lifetime of 880 ns (400 times longer) under identical conditions,84 hence it seems plausible that the major nonradiative deactivation pathway in the Cr0 complex involves the same types of internal conversion processes into lower-lying 3MC (and possibly 5MC) states as in typical octahedral FeII compounds (Figure 1c).83
Table 4. MLCT Lifetimes (τMLCT) and Emission Properties of Cr0, Mo0, MnI, and CoIII Complexes in Solution at Room Temperaturea.
| entry | compound | molecular structure | τMLCT | emission properties |
|---|---|---|---|---|
| 1 | [Cr(LH)3]b | Figure 7c (R = H) | 2.2 ns | ϕ = 10–5, λmax = 630 nm83 |
| 2 | [Cr(LPy)3]c | Figure 7c (R = pyrenyl) | 6.1 nsd | ϕ = 9 × 10–4, λmax = 682 nm85 |
| 3 | [Mn(Lbi)3]+e | Figure 7e | 0.74 nsf | ϕ = 5 × 10–4, λmax = 478 nm90 |
| 4 | [Mn(Ltri)2]+e | Figure 7f | 1.73 nsf | ϕ = 3 × 10–4, λmax = 525 nm90 |
| 5 | [Mn(CNdippPhOMe2)6]+ | Figure 7g | N/A | nonemissive94 |
| 6 | [Co(bimca)2]+g | Figure 7h | 1.2 nsh | nonemissive108 |
ϕ is the photoluminescence quantum yield, λmax is the emission band maximum.
In deaerated THF at 20 °C.83
In cyclooctane at 20 °C.
Weighted average lifetime of a biexponential decay, attributed to different conformers.
In deaerated CH3CN at 20 °C.
Weighted average lifetime of a triexponential decay, attributed to different conformers.
In deaerated CH3CN at 22 °C.108
Likely not a pure MLCT state, but instead containing substantial intraligand character.
Nonetheless, the distortion of the emissive 3MLCT state plays an non-negligible role according to a follow-up study, in which pyrenyl substituents were attached at the periphery of the diisocyanide chelate (Figure 7c, R = pyrenyl).85 Though these substituents are oriented more or less orthogonally to the m-terphenyl backbone in the electronic ground state, transient absorption spectroscopy provides evidence that the emissive MLCT state has admixed pyrene character, compatible with the view that the MLCT-excited electron is more delocalized in [Cr(LPy)3] than in [Cr(LH)3]. Thus it seems plausible that the luminescent MLCT excited state of [Cr(LPy)3] is somewhat less distorted than in [Cr(LH)3], analogously to what is known from RuII and OsII polypyridines with enlarged π-conjugation networks.75,86−88 This picture of weaker excited-state distortion seems useful to help understand the 90 times higher luminescence quantum yield of [Cr(LPy)3] in comparison to [Cr(LH)3] (Table 4, entry 2), and it is supported by the observation of a roughly 20% narrower MLCT luminescence bandwidth for [Cr(LPy)3] compared to [Cr(LH)3]. An uncommon bell-shaped dependence of the MLCT lifetime on solvent polarity is obtained for [Cr(LPy)3], because the highest polarity solvents lower the MLCT energy and thereby facilitate nonradiative relaxation directly to the ground state, whereas too strongly apolar solvents raise the MLCT energy to the extent that internal conversion to MC states becomes increasingly favorable (Figure 7d).85,89 The longest lifetime was measured in cyclooctane (6.1 ns), representing the record 3MLCT lifetime reported for a 3d6 complex in solution at room temperature so far.
Manganese(I) Arylisocyanides
[Mn(Lbi)3]+ and [Mn(Ltri)2]+ (Figure 7e/f) are analogues of [Ru(bpy)3]2+ and [Ru(tpy)2]2+, and furthermore represent the first examples of MnI compounds emitting from an MLCT excited state in solution at room temperature.89,90 Owing to the higher oxidation state of MnI with respect to Cr0 (implying a more stabilized HOMO), the MLCT absorption and emission bands are considerably blue-shifted when using similar types of ligands. Thiophene instead of benzene units were used in the ligand backbones to connect individual arylisocyanide subunits, in an attempt to optimize the bite angles of the chelate ligands and the LUMO energies. Studies of Mo0 complexes indicated that the change from six-membered benzene rings to five-membered thiophene linkers on isocyanide ligand backbones can decrease the energy of the lowest MLCT excited state by roughly 3000 cm–1,91,92 and this seemed desirable to ensure that an MLCT state indeed becomes the energetically lowest excited state, even for the less readily oxidizable MnI. Despite this design principle, the lowest intraligand π–π* state is energetically close, and a triplet state of this type seems to play a non-negligible role in the photophysics and photochemistry of [Mn(Lbi)3]+ and [Mn(Ltri)2]+.90 The MLCT luminescence quantum yields in deaerated acetonitrile at 20 °C are below 0.1% and the excited-state lifetimes are shorter than 2 ns (Table 4, entries 3–4), yet both photoinduced electron transfer and triplet–triplet energy transfer to various substrates proceeded readily. Lbi and Ltri require multistep syntheses, and therefore more readily accessible monodentate arylisocyanide ligands would be attractive alternatives for luminescent MnI complexes. However, with a monodentate ligand that gave a brightly emissive W0 complex in earlier studies,93 the resulting hexakis(arylisocyanide)manganese(I) complex (Figure 7g) was nonluminescent and instead featured light-induced ligand dissociation,94 similar to what is known from related carbonyl compounds.95
Cobalt(III) Complex in a Coordination Environment of Amido and Carbene Ligand Motifs
The high oxidation state of CoIII relative to the other 3d6 species considered herein leads to comparatively strong ligand fields and furthermore increases the energy of the lowest MLCT state.96 When combined with strong-field π-acceptor ligands such as cyanides or strongly σ-donating NHC ligands, the energetically lowest-lying MC state is sufficiently high above the ground state to become emissive, despite its substantial distortion.97−100 The same electronic state that causes ultrafast nonradiative MLCT deactivation in FeII polypyridines then features microsecond lifetimes in solution at room temperature.100 In combination with electron-rich imine ligands, an LMCT state can become the energetically lowest (and luminescent) excited state,101 representing a reversal of the preferred charge transfer direction relative to Cr0, MnI, and FeII. This is unsurprising, because higher oxidation states are of course more easily reduced, and this design principle has found much recent attention for the development of LMCT luminophores,102 for example, based on ZrIV,103,104 MnIV,105 or FeIII.52,53,106 Installing an energetically lowest-lying MLCT state in a CoIII complex therefore requires particularly careful ligand design.
The bimca ligand (Figure 7h) known from previous studies of four-coordinate square-planar complexes in other contexts107 seemed promising to increase the electron density at the cobalt center, because bimca’s central amido subunit is strongly π-donating, which could help establish an energetically low-lying MLCT state. At the same time, the tightly flanking NHC units of bimca provide an overall rigid strong-field coordination environment that looked suitable to shift MC states to high energies. Compared to related cobalt complexes, the potential of the CoIV/III redox couple of [Co(bimca)2]+ is shifted cathodically by roughly 0.5 V,108 confirming the anticipated strong influence of the amido subunit regarding the ease of metal oxidation. Transient UV–vis absorption spectroscopy combined with cyclic voltammetry and spectro-electrochemical studies support a picture, in which an intraligand charge transfer (ILCT) and an MLCT excited state are energetically close, and in which a photoactive excited state at ∼2.6 eV featuring a lifetime of 1.2 ns in acetonitrile at 22 °C (Table 4, entry 6) has substantial MLCT character, thus resembling in its composition the luminescent excited states of many cyclometalated IrIII compounds.109 In [Co(bimca)2]+ the respective excited state is nonemissive, but readily undergoes photoinduced electron transfer to methyl viologen, albeit only with a low cage-escape quantum yield of 2% in acetonitrile. Similar cage escape yields, in the range 1–7% were reported for electron transfers between an FeIII photosensitizer and organic electron donors (dimethyltoluidine, dimethylaniline, and tritolylamine) in polar solvents (acetonitrile and dimethylformamide).53,110 Sub-picosecond transient absorption spectroscopy signaled the presence of a higher-lying excited state with a lifetime of roughly 100 ps and a spectral signature compatible with the above-mentioned ILCT state, suggesting that internal conversion from the ILCT to the MLCT state is unusually slow, perhaps because the respective two states are largely decoupled from one another. The roles and the energies of MC excited states in [Co(bimca)2]+ have remained somewhat unclear, and in principle the above-mentioned photoinduced electron transfer reactivity could also emerge from an MC state, similar to what is known for FeII polypyridines.111 Pump–probe X-ray experiments and detailed (time-dependent) density functional calculations would seem desirable to further elucidate the roles and energies of ILCT, MLCT, and MC states in [Co(bimca)2]+.
Lessons Learned and Future Challenges
Over the past ten years, carbene ligands came into focus for FeII complexes, because they allowed an elongation of the 3MLCT lifetimes from the subpicosecond range14,112,113 to tens of picoseconds for N-heterocyclic carbenes7,10,36,114 and up to the subnanosecond time scale for a mesoionic carbene.50 This 3MLCT lifetime elongation is largely attributable to slower relaxation into MC states, principally 3T1 and 5T2 in octahedral symmetry (Figure 1c), because these MC states are shifted to higher energies when using carbene instead of polypyridine ligands.115 The strongly σ-donating character of NHCs raises the 3T1 and 5T2 state energies due to the destabilization of the metal eg (dz2 and dx2-y2) orbitals, and the additional π-acceptor character of mesoionic carbenes stabilizes the t2g (dxy, dxz, dyz) orbitals. Thus, the very classical design principle of maximizing the ligand field strength (Figure 8a) is pursed with the carbene ligand approach, and the same holds true for heteroleptic FeII complexes incorporating cyanido or cyanoborylated ligands in addition to polypyridines, resulting in MLCT lifetimes in the tens of picoseconds (Table 5, entry 1).29−32,34 Recent studies of CoIII complexes suggest that the factors governing the energies of MC and MLCT states can indeed be controlled independently,116 which seems plausible given that MC states usually depend more strongly on ligand field parameters whereas MLCT states are more influenced by the metal and ligand redox properties.
Figure 8.
Illustration of effects having an important influence on MLCT lifetimes in 3d6 complexes. (a) σ-donating and π-accepting ligands to maximize the ligand field strength. (b) π-donor ligands can increase the metal–ligand bond covalence and lower the Racah B parameter. (c) Optimized bite angles are important to maximize the overlaps between metal and ligand orbitals. (d) Rigidity of the ligands and the overall complexes can help minimize unwanted excited-state distortions and decelerate nonradiative relaxation. (e) Delocalization of MLCT-excited electron density can help reduce excited state distortion. (f) Deuteration can slow down nonradiative deactivation. (g) Faster radiative relaxation (kr) can make photoluminescence more competitive with nonradiative processes.
Table 5. Possible Strategies to Enhance Photophysical and Photochemical Properties of d6 (and Other Types of) Metal Complexesa.
| entry | strategy | approaches used | examplesb |
|---|---|---|---|
| 1 | increase ligand field strength | strong σ donor ligands, π acceptor ligands | Cr0,83,85 MnI,90 FeII6,7,10,12,24,29,30,34,36,38,39,40−46,50 CoIII,100 Mo084,91,133 |
| 2 | increase metal–ligand bond covalence | π donor ligands | FeII,59,63 CoIII,108 CrIII117,118,134,135 |
| 3 | optimize metal–ligand orbital overlap | chelates with optimized bite angles | MnI,90 FeII,22,23,48 CoIII,108 CrIII124,125,128,136 |
| 4 | minimize excited state distortion | structural rigidity, enlarged π conjugation networks | Cr0,83,85 FeII,9 CoIII,108 NiII,137 CuI,130,131,138 ZrIV104 |
| 5 | decrease vibrational energies | ligand deuteration | VIII,139,140 CrIII124,141 |
| 6 | increase radiative relaxation rates | enhanced transition probability | FeII,142 CuI138 |
Not restricted to MLCT transitions.
Particular focus here is on 3d6 compounds and on other types of coordination complexes based on Earth-abundant metals; not a comprehensive list.
The energies of many MC states are however not merely a function of the ligand field parameter 10 Dq, but rather depend on the ratio of 10 Dq and B, where B is the Racah parameter characterizing the metal–ligand bond covalence. This aspect is now receiving increased attention in the design of photoactive first-row transition metal complexes,117 even though the nature of bonding in this compound class is usually considered more ionic in comparison to the second or third row.11,96 Amido π-donor ligands (Figure 6b/Table 5, entry 2) help reduce the repulsion between d-electrons owing to the nephelauxetic effect,63,108 and the Racah B parameter can be lowered markedly (Figure 8b).118 With FeII complexes, this has been useful to approach the HOMO inversion scenario predicted computationally,60 leading to so-called PALCT excited states,59,63 which can be viewed as a form of charge transfer with a particularly strong contribution of electron density from the covalent Fe–Namido interaction. Whereas the π-donor ability of amido ligands is evident, the investigation of a spectrochemical series of CoIII complexes revealed that polypyridines can act as net π-donors, thus contrasting the behavior of these ligands toward second- and third-row transition metal complexes, in which they are net π-acceptors.96 This illustrates that well-established principles for precious metal compounds are not necessarily applicable to the first-row of transition metals, here in this case mainly because the respective metal orbital energies differ strongly.96 Computational studies forecasted this effect,119 and furthermore predicted that cyclometalating ligands could lead to long-lived MLCT states.61,64−68 Synthetic advances now provide access to such cyclometalated FeII complexes,6,72,120 and in one case led to the claim of an emissive MLCT state in solution at room temperature (Figure 6d).74 The energy gap between this MLCT state and the electronic ground state is only 1.1 eV, whereas RuII and OsII polypyridines with comparable energy gaps are usually nonemissive.87,121 Thus, the results reported for this FeII compound seem surprising, particularly for an excited state that is substantially distorted, as the apparent Stokes shift of ∼5000 cm–1 and the emission bandwidth of ∼1300 cm–1 imply.74
The optimization of ligand field strength (10 Dq) and covalence (B) discussed above is intimately tied to the bite angles of chelating ligands (Figure 8c/Table 5, entry 3), which crucially affect the overlap of metal and ligand orbitals.22,96 This aspect is illustrated by the remarkable 3MLCT lifetime difference between [Ru(bpy)3]2+ and [Ru(tpy)2]2+,122,123 and is furthermore seen in first-row transition metal complexes.124−126 Various types of bite-angle optimized chelate ligands, some of them known from earlier studies of precious metal complexes,107,127 have now been used successfully to enhance the photophysical properties of 3d compounds.22,48,108,118,125,126,128
The issue of bite angles is connected to structural rigidity (Figure 8d/Table 5, entry 4). While the ligand rigidity in itself can be helpful, as illustrated for example by the difference between 1,10-phenanthroline (phen) and bpy ligands in CuI complexes,129 an equally important question is to what extent cooperative steric effects between individual ligands can increase the overall rigidity of the entire coordination unit. The significance of this aspect is illustrated by CuI complexes featuring sec-butyl or cyclohexyl substituents on their phen ligands,130,131 and by ZrIV complexes with mesityl substituents on the ligand periphery.103,104 As far as photoactive 3d6 complexes are concerned, the strategy of cooperative rigidity seems underexplored yet, though NMR experiments on a Cr0 complex with tert-butyl substituents in α-position to the ligating isocyanide units (Figure 7c) provide direct evidence for steric interactions between substituents from different chelate ligands.85 Bulky tert-butyl substituents seemed indispensable for obtaining luminescent Cr0 and MnI complexes,89 because in earlier studies with Mo0 (4d6) complexes the replacement of methyl by tert-butyl groups improved the MLCT lifetimes and luminescence quantum yield by an order of magnitude.84 Overall rigidity with mutually interlocked ligands can be useful to minimize excited state distortion and to restrict the vibrational degrees of freedom, both of which can help decelerate nonradiative relaxation processes. Not all vibrational modes deactivate a given excited state equally well, and a recent study of a tailor-made FeII compound to monitor coherence phenomena provided direct insight into what kinds of complex distortions are particularly relevant.9,25 Such investigations can inform future ligand designs,132 and this specific case study furthermore illustrated how rigidity slows MLCT excited state decay.
Another approach to minimize excited state distortion involves the use of extended π-conjugation networks on the ligands, to delocalize the MLCT-excited electron density (Figure 8e/Table 5, entry 4). This strategy is well-known from RuII polypyridines,75,87 and recently has been applied successfully to a Cr0 compound (Figure 7c), resulting in the longest-lived 3MLCT excited state of a 3d6 compound in solution at room temperature known to date.85 Alternatively, polyaromatic hydrocarbon (PAH) groups (for example anthracenyl or pyrenyl substituents) at the ligand periphery could lead to the so-called triplet reservoir effect,47 in which the long-lived T1 state of the PAH is in equilibrium with the photoactive 3MLCT state. However, for this to occur, energy matching between the respective two states is indispensable, otherwise the PAH groups have no significant effect in terms of MLCT lifetime elongation (Figure 4g/h).46
Higher energy vibrations mediate nonradiative relaxation better than lower energy modes, and hence ligand deuteration can have a beneficial influence (Figure 8f). This strategy has long been applied to lanthanide complexes as well as 4d6 and 5d6 polypyridines.143,144 Recent work demonstrated that it can also be very relevant for first-row transition metal compounds,128,139−141 though 3d6 complexes played no important role until now. Site-selective deuteration experiments showed that the C–H/C–D oscillators in immediate proximity to the metal core have the biggest influence, and furthermore the energetic match between vibrational overtones and the photoactive excited state is a key factor (Table 5, entry 5).114,120 The isocyanide ligands needed to obtain emissive Cr0 and MnI complexes feature high-energy CN stretching vibrations (∼1950 cm–1) in immediate proximity to the metal core,83,85,89,90,92,94,145 which could represent an inherent disadvantage of this class of compounds.
In addition to slowing down nonradiative MLCT excited state relaxation, accelerating radiative emission can help enhance the photoluminescence quantum yields (Figure 8g/Table 5, entry 6). While the energy gap law correctly predicts the nonradiative rate constants for a family of related complexes, the current understanding of how to control radiative MLCT excited state decay seems less well developed. Radiative relaxation from 3MLCT states in d6 metal complexes is associated with comparatively low rate constants due to the spin-forbidden nature of the transition to the electronic (singlet) ground state. Mixing of the relevant 3MLCT state with singlet excited states could in principle be expected to lead to higher radiative decay rates, but this aspect seems underexplored for complexes of Cr0, MnI, FeII, and CoIII. Stark spectroscopy has provided evidence for spin disallowed absorption transitions in [Fe(bpy)3]2+,142 which seems encouraging for the idea that greater radiative excited-state decay rates could be obtained by singlet–triplet mixing in low-spin d6 compounds.
No matter what exact design principles are applied, the photoactive MLCT excited states of d6 complexes will likely remain considerably more strongly distorted than the spin-flip excited states in d3 metal compounds, as becomes evident when comparing the Huang–Rhys factors of 3MLCT states in [Ru(bpy)3]2+ and its congeners with those of the lowest spin-flip excited states of CrIII.146,147 Fighting nonradiative relaxation is therefore inherently more difficult when dealing with 3d6 MLCT excited states, and this should be kept in mind when comparing photophysical figures of merit between d6 MLCT luminophores and d3 spin-flip (MC) emitters.128,148
Photocatalysis
Whereas 4d6 and 5d6 complexes and their MLCT excited states are extensively used in organic photoredox and energy transfer catalysis,2,149,150 this is not yet the case for 3d6 analogues. There is an increasing body of literature reporting on photocatalysis with FeII complexes (mostly with polypyridine ligands),151−154 though this usually implies reactivity from the lowest MC excited state.111 A recent study claimed that the MLCT excited state of a cyclometalated FeII complex (Figure 6e) undergoes direct photoreaction, though this was a stoichiometric, not catalytic process.74 A clear-cut case of catalytic MLCT photoreaction with 3d6 metal complexes has been observed with MnI isocyanides (Figure 7e/f), but only for stilbene photoisomerization, as a proof of principle for triplet energy transfer catalysis.89,90 Clearly, the photoreactivity of 3MLCT excited states of 3d6 complexes is underexplored yet, mainly due to the fact that most of the compounds known to date have too short MLCT lifetimes. Work on Mo0 and W0 isocyanide complexes suggests that Cr0 isocyanides could be very potent photoreductants.84,93,133,145,155−157 Photochemical reduction reactions seem therefore most attractive with Cr0 and MnI complexes, going through an oxidative 3MLCT excited-state quenching pathway, followed by subsequent reduction of the resulting 3d5 species to regenerate the initial 3d6 complex. Reductive excited-state quenching does not currently seem to be a viable option for these electron-rich isocyanide complexes, simply because they are very difficult to reduce. Conversely, CoIII complexes are potentially attractive photo-oxidants, and, in the form of LMCT chromophores, have already been used in this capacity.101 The observation of photoinduced electron transfer from [Co(bimca)2]+ to methyl viologen indicates however that photoreductions are also realistic, despite the high (formal) metal oxidation state.108 FeII complexes would seem well-suited to tackle both photo-oxidations and photoreduction reactions, analogously to RuII compounds. In a recent study, the 2LMCT excited state of [Fe(btz)3]3+ was quenched reductively, and the formed [Fe(btz)3]2+ photoproduct was then excited again to its 3MLCT excited state, which initiated an atom transfer radical addition (ATRA) reaction.55 Such excitation strategies resembling the Z-scheme of natural photosynthesis have recently become of interest also with other first-row transition metal complexes.158,159
Parallels and Differences between 3d6 and 3d10/3d8 Complexes with MLCT Excited States
Energetically low-lying MLCT excited states are typically obtained for late, electron-rich transition metals, and aside from d6, the d8 and d10 electron configurations are particularly relevant. Semiprecious CuI with its completely filled d-subshell has received vast attention.129,160−163 In the absence of low-lying MC states, the installment of long-lived MLCT states becomes relatively straightforward, particularly when compared to the 3d6 configuration. The focus has long been on four-coordinate, tetrahedral CuI complexes, but recently there is a trend toward two-coordinate linear CuI complexes with carbene ligands, often featuring a push–pull design between different ligands, such that ligand-to-ligand charge transfer (LLCT) states become the lowest excited states.18,138,164 Since those LLCT states cause less molecular distortion than the Jahn–Teller susceptible MLCT excited states of d10 complexes,165 the emission properties of this new compound class are particularly favorable and will likely enable future applications in lighting and sensing.166
From a more fundamental curiosity-driven perspective, isoelectronic analogues based on Ni0 and ZnII seem attractive. Photoactive tetrahedral Ni0 complexes are underexplored in comparison to CuI,167 likely due to stability issues that seem however not insurmountable.168 ZnII complexes are better explored,169 though they typically feature ligand-based fluorescent excited states rather than photoactive MLCT states.170 Recently, there has been increasing focus on harnessing triplet excited states of ZnII complexes and on introducing charge transfer character into the photoactive excited states,171−173 including the development of emitters displaying thermally activated delayed fluorescence (TADF).174 Lacking ligand field stabilization energy, ZnII coordination compounds are often substitution-labile and tend to form pentacoordinate or even polynuclear compounds,175 and multiple species can exist in dynamic equilibrium with each other.176 Possible future focal points with ZnII could be the design of compounds with photoactive LLCT or ILCT excited states, whereas Ni0 compounds seem attractive as classical MLCT luminophores and photosensitizers.
Even bigger challenges await with photoactive MLCT states in 3d8 complexes. Many square-planar PtII and AuIII compounds luminesce from charge-transfer excited states,177−179 and with the recent surge of interest in combined photoredox and nickel cross-coupling catalysis, the possible involvement of MLCT states in square-planar NiII complexes has received increasing attention.180 Complexes of the type in Figure 9a are thermally stable variants of catalytically relevant reaction intermediates in cross-coupling reactions,181−183 and time-resolved spectroscopic studies revealed that their MLCT excited states relax to an MC state within a few picoseconds, thereby undergoing a structural distortion from square-planar to tetrahedral geometry.184 Recent work demonstrated that when this structural distortion is made more difficult by suitable molecular design (Figure 9b), then longer MLCT lifetimes are achievable.137 Many different research avenues are currently pursued with photoactive NiII complexes,183,185−187 and much remains to be discovered yet concerning the basic photophysics and photochemistry of square-planar 3d8 complexes. Luminescent and photochemically exploitable MLCT excited states could be one of the future targets, and this challenge seems even greater than in the case of 3d6 complexes.137
Figure 9.
Next big fundamental challenge concerning long-lived MLCT excited states in first-row transition metal complexes with a partially filled d-subshell:180 (a) Square-planar NiII complexes undergoing facile distortion to tetrahedral geometries; R1 = OMe, tBu, H, Ph, CO2Et with R2 = H and X = Cl; R1 = tBu with R2 = OMe, H, CF3 and X = Br.184 (b) Longer MLCT excited state lifetimes are obtained by making the structural distortion from square-planar to tetrahedral more difficult.137 (c) Square-planar NiII complex with an MLCT excited state lifetime of 48 ps in acetonitrile at room temperature.137
Conclusions
In modern research on first-row transition metals and their coordination complexes, the relatively high abundance as well as the lower costs of these elements in comparison to platinum group metals are often emphasized.188,189 In the big picture, societal, environmental, and geopolitical aspects are furthermore very relevant. The mining of some first-row transition metals can be problematic.190,191 The overall complexity of the situation when trying to identify advantages of using first-row transition metals instead of precious metals therefore necessitates a very nuanced approach.
In the specific field of photoactive transition metal complexes, the cost argument is currently particularly tricky, because many of the ligands required to install the desirable photophysical and photochemical properties rely on multistep syntheses. It has therefore been questioned whether these expensive ligands outweigh the cost savings made with the metal.192 Most of the contemporary research in this field seems however driven by curiosity and the desire to understand fundamentally new aspects.17,95,193 Even after decades of research dominated by work on the platinum group metals, many key aspects of excited-state relaxation, which are particularly relevant to the first row of the transition series, have remained poorly understood. The combined advances in synthesis, spectroscopy, and computational chemistry now permit much progress in understanding the fundamental photophysics and photochemistry of 3d compounds.
Keeping an eye on possible future applications in solar energy conversion, lighting, sensing, or photocatalysis, it seems desirable that the full diversity of electronically excited states occurring in metal complexes is accessible, including LMCT,102,106 LLCT,166 ILCT,173 and MC states124 in addition to MLCT excited states.194 For complexes with a partially filled 3d subshell, rapid nonradiative MLCT excited state deactivation is particularly prevalent, and to date luminescent MLCT excited states with nanosecond lifetimes have remained very rare among 3d6 compounds.18,83,85,89,90 Consequently, MLCT-based photocatalysis has remained extremely scarce in this compound class.90 Most attention still goes to FeII,6 but CoIII seems to attract growing interest,96,100,108 particularly in combination with carbene ligands. The best performers in terms of luminescence and MLCT lifetimes are currently based on Cr0 and MnI, for which chelating arylisocyanides are important.89 NiIV does not seem to have been considered yet from a photophysical perspective.195,196 Square-planar NiII (3d8) compounds represent a great challenge that has now been taken on by several research groups.
Acknowledgments
Funding from the Swiss National Science Foundation through grant number 200020_207329 is acknowledged.
The authors declare no competing financial interest.
References
- Caspar J. V.; Meyer T. J. Application of the Energy-Gap Law to Nonradiative Excited-State Decay. J. Phys. Chem. 1983, 87, 952–957. 10.1021/j100229a010. [DOI] [Google Scholar]
- Arias-Rotondo D. M. The fruit fly of photophysics. Nat. Chem. 2022, 14, 716–716. 10.1038/s41557-022-00955-8. [DOI] [PubMed] [Google Scholar]
- Knopf K. M.; Murphy B. L.; MacMillan S. N.; Baskin J. M.; Barr M. P.; Boros E.; Wilson J. J. In Vitro Anticancer Activity and in Vivo Biodistribution of Rhenium(I) Tricarbonyl Aqua Complexes. J. Am. Chem. Soc. 2017, 139, 14302–14314. 10.1021/jacs.7b08640. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vittardi S. B.; Thapa Magar R.; Breen D. J.; Rack J. J. A Future Perspective on Phototriggered Isomerizations of Transition Metal Sulfoxides and Related Complexes. J. Am. Chem. Soc. 2021, 143, 526–537. 10.1021/jacs.0c08820. [DOI] [PubMed] [Google Scholar]
- DiLuzio S.; Connell T. U.; Mdluli V.; Kowalewski J. F.; Bernhard S. Understanding Ir(III) Photocatalyst Structure–Activity Relationships: A Highly Parallelized Study of Light-Driven Metal Reduction Processes. J. Am. Chem. Soc. 2022, 144, 1431–1444. 10.1021/jacs.1c12059. [DOI] [PubMed] [Google Scholar]
- Dierks P.; Vukadinovic Y.; Bauer M. Photoactive iron complexes: more sustainable, but still a challenge. Inorg. Chem. Front. 2022, 9, 206–220. 10.1039/D1QI01112J. [DOI] [Google Scholar]
- Liu Y. Z.; Persson P.; Sundström V.; Wärnmark K. Fe N-heterocyclic carbene complexes as promising photosensitizers. Acc. Chem. Res. 2016, 49, 1477–1485. 10.1021/acs.accounts.6b00186. [DOI] [PubMed] [Google Scholar]
- Oppermann M.; Zinna F.; Lacour J.; Chergui M. Chiral control of spin-crossover dynamics in Fe(II) complexes. Nat. Chem. 2022, 14, 739–745. 10.1038/s41557-022-00933-0. [DOI] [PubMed] [Google Scholar]
- Paulus B. C.; Adelman S. L.; Jamula L. L.; McCusker J. K. Leveraging excited-state coherence for synthetic control of ultrafast dynamics. Nature 2020, 582, 214–218. 10.1038/s41586-020-2353-2. [DOI] [PubMed] [Google Scholar]
- Cebrián C.; Pastore M.; Monari A.; Assfeld X.; Gros P. C.; Haacke S. Ultrafast Spectroscopy of Fe(II) Complexes Designed for Solar-Energy Conversion: Current Status and Open Questions. ChemPhysChem 2022, 23, e202100659. 10.1002/cphc.202100659. [DOI] [PubMed] [Google Scholar]
- McCusker J. K. Electronic structure in the transition metal block and its implications for light harvesting. Science 2019, 363, 484–488. 10.1126/science.aav9104. [DOI] [PubMed] [Google Scholar]
- Zhang W. K.; Kjaer K. S.; Alonso-Mori R.; Bergmann U.; Chollet M.; Fredin L. A.; Hadt R. G.; Hartsock R. W.; Harlang T.; Kroll T.; Kubicek K.; Lemke H. T.; Liang H. W.; Liu Y. Z.; Nielsen M. M.; Persson P.; Robinson J. S.; Solomon E. I.; Sun Z.; Sokaras D.; van Driel T. B.; Weng T. C.; Zhu D. L.; Wärnmark K.; Sundström V.; Gaffney K. J. Manipulating charge transfer excited state relaxation and spin crossover in iron coordination complexes with ligand substitution. Chem. Sci. 2017, 8, 515–523. 10.1039/C6SC03070J. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Carey M. C.; Adelman S. L.; McCusker J. K. Insights into the excited state dynamics of Fe(II) polypyridyl complexes from variable-temperature ultrafast spectroscopy. Chem. Sci. 2019, 10, 134–144. 10.1039/C8SC04025G. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Auböck G.; Chergui M. Sub-50-fs photoinduced spin crossover in Fe(bpy3)2+. Nat. Chem. 2015, 7, 629–633. 10.1038/nchem.2305. [DOI] [PubMed] [Google Scholar]
- Monat J. E.; McCusker J. K. Femtosecond Excited-State Dynamics of an Iron(II) Polypyridyl Solar Cell Sensitizer Model. J. Am. Chem. Soc. 2000, 122, 4092–4097. 10.1021/ja992436o. [DOI] [Google Scholar]
- Zhang K.; Ash R.; Girolami G. S.; Vura-Weis J. Tracking the metal-centered triplet in photoinduced spin crossover of Fe(phen)32+ with tabletop femtosecond M-edge X-ray absorption near-edge structure spectroscopy. J. Am. Chem. Soc. 2019, 141, 17180–17188. 10.1021/jacs.9b07332. [DOI] [PubMed] [Google Scholar]
- Förster C.; Heinze K. Photophysics and photochemistry with earth-abundant metals – fundamentals and concepts. Chem. Soc. Rev. 2020, 49, 1057–1070. 10.1039/C9CS00573K. [DOI] [PubMed] [Google Scholar]
- Wegeberg C.; Wenger O. S. Luminescent First-Row Transition Metal Complexes. JACS Au 2021, 1, 1860–1876. 10.1021/jacsau.1c00353. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mann K. R.; Gray H. B.; Hammond G. S. Excited-state reactivity patterns of hexakisarylisocyano complexes of chromium(0), molybdenum(0), and tungsten(0). J. Am. Chem. Soc. 1977, 99, 306–307. 10.1021/ja00443a083. [DOI] [Google Scholar]
- Lees A. J. Luminescence properties of organometallic complexes. Chem. Rev. 1987, 87, 711–743. 10.1021/cr00080a003. [DOI] [Google Scholar]
- Creutz C.; Chou M.; Netzel T. L.; Okumura M.; Sutin N. Lifetimes, spectra, and quenching of the excited states of polypyridine complexes of iron(II), ruthenium(II), and osmium(II). J. Am. Chem. Soc. 1980, 102, 1309–1319. 10.1021/ja00524a014. [DOI] [Google Scholar]
- Jamula L. L.; Brown A. M.; Guo D.; McCusker J. K. Synthesis and Characterization of a High-Symmetry Ferrous Polypyridyl Complex: Approaching the 5T2/3T1 Crossing Point for FeII. Inorg. Chem. 2014, 53, 15–17. 10.1021/ic402407k. [DOI] [PubMed] [Google Scholar]
- Mengel A. K. C.; Förster C.; Breivogel A.; Mack K.; Ochsmann J. R.; Laquai F.; Ksenofontov V.; Heinze K. A Heteroleptic Push–Pull Substituted Iron(II) Bis(tridentate) Complex with Low-Energy Charge-Transfer States. Chem.—Eur. J. 2015, 21, 704–714. 10.1002/chem.201404955. [DOI] [PubMed] [Google Scholar]
- Darari M.; Francés-Monerris A.; Marekha B.; Doudouh A.; Wenger E.; Monari A.; Haacke S.; Gros P. C. Towards Iron(II) Complexes with Octahedral Geometry: Synthesis, Structure and Photophysical Properties. Molecules 2020, 25, 5991. 10.3390/molecules25245991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paulus B. C.; McCusker J. K. On the use of vibronic coherence to identify reaction coordinates for ultrafast excited-state dynamics of transition metal-based chromophores. Faraday Discuss. 2022, 237, 274–299. 10.1039/D2FD00106C. [DOI] [PubMed] [Google Scholar]
- Wächtler M.; Kübel J.; Barthelmes K.; Winter A.; Schmiedel A.; Pascher T.; Lambert C.; Schubert U. S.; Dietzek B. Energy transfer and formation of long-lived 3MLCT states in multimetallic complexes with extended highly conjugated bis-terpyridyl ligands. Phys. Chem. Chem. Phys. 2016, 18, 2350–2360. 10.1039/C5CP04447B. [DOI] [PubMed] [Google Scholar]
- Shepard S. G.; Fatur S. M.; Rappe A. K.; Damrauer N. H. Highly Strained Iron(II) Polypyridines: Exploiting the Quintet Manifold To Extend the Lifetime of MLCT Excited States. J. Am. Chem. Soc. 2016, 138, 2949–2952. 10.1021/jacs.5b13524. [DOI] [PubMed] [Google Scholar]
- Fatur S. M.; Shepard S. G.; Higgins R. F.; Shores M. P.; Damrauer N. H. A Synthetically Tunable System To Control MLCT Excited-State Lifetimes and Spin States in Iron(II) Polypyridines. J. Am. Chem. Soc. 2017, 139, 4493–4505. 10.1021/jacs.7b00700. [DOI] [PubMed] [Google Scholar]
- Vittardi S. B.; Magar R. T.; Schrage B. R.; Ziegler C. J.; Jakubikova E.; Rack J. J. Evidence for a lowest energy 3MLCT excited state in [Fe(tpy)(CN)3]−. Chem. Commun. 2021, 57, 4658–4661. 10.1039/D1CC01090E. [DOI] [PubMed] [Google Scholar]
- Liu Y. Z.; Kjaer K. S.; Fredin L. A.; Chabera P.; Harlang T.; Canton S. E.; Lidin S.; Zhang J. X.; Lomoth R.; Bergquist K. E.; Persson P.; Wärnmark K.; Sundström V. A Heteroleptic Ferrous Complex with Mesoionic Bis(1,2,3-triazol-5-ylidene) Ligands: Taming the MLCT Excited State of Iron(II). Chem.—Eur. J. 2015, 21, 3628–3639. 10.1002/chem.201405184. [DOI] [PubMed] [Google Scholar]
- Winkler J. R.; Creutz C.; Sutin N. Solvent tuning of the excited-state properties of (2,2′-bipyridine)tetracyanoferrate(II): direct observation of a metal-to-ligand charge-transfer excited state of iron(II). J. Am. Chem. Soc. 1987, 109, 3470–3471. 10.1021/ja00245a053. [DOI] [Google Scholar]
- Zederkof D. B.; Mo̷ller K. B.; Nielsen M. M.; Haldrup K.; González L.; Mai S. Resolving Femtosecond Solvent Reorganization Dynamics in an Iron Complex by Nonadiabatic Dynamics Simulations. J. Am. Chem. Soc. 2022, 144, 12861–12873. 10.1021/jacs.2c04505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schmid L.; Kerzig C.; Prescimone A.; Wenger O. S. Photostable Ruthenium(II) Isocyanoborato Luminophores and Their Use in Energy Transfer and Photoredox Catalysis. JACS Au 2021, 1, 819–832. 10.1021/jacsau.1c00137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schmid L.; Chábera P.; Rüter I.; Prescimone A.; Meyer F.; Yartsev A.; Persson P.; Wenger O. S. Borylation in the Second Coordination Sphere of FeII Cyanido Complexes and Its Impact on Their Electronic Structures and Excited-State Dynamics. Inorg. Chem. 2022, 61, 15853–15863. 10.1021/acs.inorgchem.2c01667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Curtin G. M.; Jakubikova E. Extended π-Conjugated Ligands Tune Excited-State Energies of Iron(II) Polypyridine Dyes. Inorg. Chem. 2022, 61, 18850–18860. 10.1021/acs.inorgchem.2c02362. [DOI] [PubMed] [Google Scholar]
- Duchanois T.; Liu L.; Pastore M.; Monari A.; Cebrian C.; Trolez Y.; Darari M.; Magra K.; Frances-Monerris A.; Domenichini E.; Beley M.; Assfeld X.; Haacke S.; Gros P. C. NHC-based iron sensitizers for DSSCs. Inorganics 2018, 6, 63. 10.3390/inorganics6020063. [DOI] [Google Scholar]
- Lindh L.; Chabera P.; Rosemann N. W.; Uhlig J.; Wärnmark K.; Yartsev A.; Sundström V.; Persson P. Photophysics and Photochemistry of Iron Carbene Complexes for Solar Energy Conversion and Photocatalysis. Catalysts 2020, 10, 315. 10.3390/catal10030315. [DOI] [Google Scholar]
- Liu Y. Z.; Harlang T.; Canton S. E.; Chabera P.; Suarez-Alcantara K.; Fleckhaus A.; Vithanage D. A.; Goransson E.; Corani A.; Lomoth R.; Sundström V.; Wärnmark K. Towards Longer-Lived Metal-to-Ligand Charge Transfer States of Iron(II) Complexes: An N-Heterocyclic Carbene Approach. Chem. Commun. 2013, 49, 6412–6414. 10.1039/c3cc43833c. [DOI] [PubMed] [Google Scholar]
- Duchanois T.; Etienne T.; Beley M.; Assfeld X.; Perpete E. A.; Monari A.; Gros P. C. Heteroleptic Pyridyl-Carbene Iron Complexes with Tuneable Electronic Properties. Eur. J. Inorg. Chem. 2014, 2014, 3747–3753. 10.1002/ejic.201402356. [DOI] [Google Scholar]
- Duchanois T.; Etienne T.; Cebrian C.; Liu L.; Monari A.; Beley M.; Assfeld X.; Haacke S.; Gros P. C. An Iron-Based Photosensitizer with Extended Excited-State Lifetime: Photophysical and Photovoltaic Properties. Eur. J. Inorg. Chem. 2015, 2015, 2469–2477. 10.1002/ejic.201500142. [DOI] [Google Scholar]
- Liu L.; Duchanois T.; Etienne T.; Monari A.; Beley M.; Assfeld X.; Haacke S.; Gros P. C. A New Record Excited State 3MLCT Lifetime for Metalorganic Iron(II) Complexes. Phys. Chem. Chem. Phys. 2016, 18, 12550–12556. 10.1039/C6CP01418F. [DOI] [PubMed] [Google Scholar]
- Harlang T. C. B.; Liu Y. Z.; Gordivska O.; Fredin L. A.; Ponseca C. S.; Huang P.; Chabera P.; Kjaer K. S.; Mateos H.; Uhlig J.; Lomoth R.; Wallenberg R.; Styring S.; Persson P.; Sundström V.; Wärnmark K. Iron Sensitizer Converts Light to Electrons with 92% Yield. Nat. Chem. 2015, 7, 883–889. 10.1038/nchem.2365. [DOI] [PubMed] [Google Scholar]
- Zimmer P.; Burkhardt L.; Friedrich A.; Steube J.; Neuba A.; Schepper R.; Müller P.; Florke U.; Huber M.; Lochbrunner S.; Bauer M. The connection between NHC ligand count and photophysical properties in Fe(II) photosensitizers: an experimental study. Inorg. Chem. 2018, 57, 360–373. 10.1021/acs.inorgchem.7b02624. [DOI] [PubMed] [Google Scholar]
- Darari M.; Domenichini E.; Francés-Monerris A.; Cebrián C.; Magra K.; Beley M.; Pastore M.; Monari A.; Assfeld X.; Haacke S.; Gros P. C. Iron(II) complexes with diazinyl-NHC ligands: impact of π-deficiency of the azine core on photophysical properties. Dalton Trans. 2019, 48, 10915–10926. 10.1039/C9DT01731C. [DOI] [PubMed] [Google Scholar]
- Vukadinovic Y.; Burkhardt L.; Päpcke A.; Miletic A.; Fritsch L.; Altenburger B.; Schoch R.; Neuba A.; Lochbrunner S.; Bauer M. When Donors Turn into Acceptors: Ground and Excited State Properties of FeII Complexes with Amine-Substituted Tridentate Bis-imidazole-2-ylidene Pyridine Ligands. Inorg. Chem. 2020, 59, 8762–8774. 10.1021/acs.inorgchem.0c00393. [DOI] [PubMed] [Google Scholar]
- Dierks P.; Päpcke A.; Bokareva O. S.; Altenburger B.; Reuter T.; Heinze K.; Kühn O.; Lochbrunner S.; Bauer M. Ground- and Excited-State Properties of Iron(II) Complexes Linked to Organic Chromophores. Inorg. Chem. 2020, 59, 14746–14761. 10.1021/acs.inorgchem.0c02039. [DOI] [PubMed] [Google Scholar]
- Howarth A. J.; Majewski M. B.; Wolf M. O. Photophysical properties and applications of coordination complexes incorporating pyrene. Coord. Chem. Rev. 2015, 282, 139–149. 10.1016/j.ccr.2014.03.024. [DOI] [Google Scholar]
- Reuter T.; Kruse A.; Schoch R.; Lochbrunner S.; Bauer M.; Heinze K. Higher MLCT lifetime of carbene iron(II) complexes by chelate ring expansion. Chem. Commun. 2021, 57, 7541–7544. 10.1039/D1CC02173G. [DOI] [PubMed] [Google Scholar]
- Jiang T.; Bai Y. S.; Zhang P.; Han Q. W.; Mitzi D. B.; Therien M. J. Electronic structure and photophysics of a supermolecular iron complex having a long MLCT-state lifetime and panchromatic absorption. Proc. Natl. Acad. Sci. U. S. A. 2020, 117, 20430–20437. 10.1073/pnas.2009996117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chábera P.; Kjaer K. S.; Prakash O.; Honarfar A.; Liu Y. Z.; Fredin L. A.; Harlang T. C. B.; Lidin S.; Uhlig J.; Sundström V.; Lomoth R.; Persson P.; Wärnmark K. FeII hexa N-heterocyclic carbene complex with a 528 ps metal-to-ligand charge-transfer excited-state lifetime. J. Phys. Chem. Lett. 2018, 9, 459–463. 10.1021/acs.jpclett.7b02962. [DOI] [PubMed] [Google Scholar]
- Sarkar B.; Suntrup L. Illuminating Iron: Mesoionic Carbenes as Privileged Ligands in Photochemistry. Angew. Chem., Int. Ed. 2017, 56, 8938–8940. 10.1002/anie.201704522. [DOI] [PubMed] [Google Scholar]
- Chábera P.; Liu Y.; Prakash O.; Thyrhaug E.; El Nahhas A.; Honarfar A.; Essén S.; Fredin L. A.; Harlang T. C. B.; Kjaer K. S.; Handrup K.; Ericsson F.; Tatsuno Y.; Morgan K.; Schnadt J.; Häggström L.; Ericsson T.; Sobkowiak A.; Lidin S.; Huang P.; Styring S.; Uhlig J.; Bendix J.; Lomoth R.; Sundström V.; Persson P.; Wärnmark K. A low-spin FeIII complex with 100-ps ligand-to-metal charge transfer photoluminescence. Nature 2017, 543, 695–699. 10.1038/nature21430. [DOI] [PubMed] [Google Scholar]
- Kjær K. S.; Kaul N.; Prakash O.; Chábera P.; Rosemann N. W.; Honarfar A.; Gordivska O.; Fredin L. A.; Bergquist K. E.; Häggström L.; Ericsson T.; Lindh L.; Yartsev A.; Styring S.; Huang P.; Uhlig J.; Bendix J.; Strand D.; Sundström V.; Persson P.; Lomoth R.; K W. Luminescence and reactivity of a charge-transfer excited iron complex with nanosecond lifetime. Science 2019, 363, 249–253. 10.1126/science.aau7160. [DOI] [PubMed] [Google Scholar]
- Prakash O.; Lindh L.; Kaul N.; Rosemann N. W.; Losada I. B.; Johnson C.; Chábera P.; Ilic A.; Schwarz J.; Gupta A. K.; Uhlig J.; Ericsson T.; Häggström L.; Huang P.; Bendix J.; Strand D.; Yartsev A.; Lomoth R.; Persson P.; Wärnmark K. Photophysical Integrity of the Iron(III) Scorpionate Framework in Iron(III)–NHC Complexes with Long-Lived 2LMCT Excited States. Inorg. Chem. 2022, 61, 17515–17526. 10.1021/acs.inorgchem.2c02410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ilic A.; Schwarz J.; Johnson C.; de Groot L. H. M.; Kaufhold S.; Lomoth R.; Wärnmark K. Photoredox Catalysis via Consecutive 2LMCT- and 3MLCT-Excitation of an Fe(III/II)-N-Heterocyclic Carbene Complex. Chem. Sci. 2022, 13, 9165–9175. 10.1039/D2SC02122F. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaul N.; Lomoth R. The Carbene Cannibal: Photoinduced Symmetry-Breaking Charge Separation in an Fe(III) N-Heterocyclic Carbene. J. Am. Chem. Soc. 2021, 143, 10816–10821. 10.1021/jacs.1c03770. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tatsuno H.; Kjær K. S.; Kunnus K.; Harlang T. C. B.; Timm C.; Guo M.; Chàbera P.; Fredin L. A.; Hartsock R. W.; Reinhard M. E.; Koroidov S.; Li L.; Cordones A. A.; Gordivska O.; Prakash O.; Liu Y.; Laursen M. G.; Biasin E.; Hansen F. B.; Vester P.; Christensen M.; Haldrup K.; Németh Z.; Sárosiné Szemes D.; Bajnóczi É.; Vankó G.; Van Driel T. B.; Alonso-Mori R.; Glownia J. M.; Nelson S.; Sikorski M.; Lemke H. T.; Sokaras D.; Canton S. E.; Dohn A. O.; Mo̷ller K. B.; Nielsen M. M.; Gaffney K. J.; Wärnmark K.; Sundström V.; Persson P.; Uhlig J. Hot Branching Dynamics in a Light-Harvesting Iron Carbene Complex Revealed by Ultrafast X-ray Emission Spectroscopy. Angew. Chem., Int. Ed. 2020, 59, 364–372. 10.1002/anie.201908065. [DOI] [PubMed] [Google Scholar]
- Dierks P.; Kruse A.; Bokareva O. S.; Al-Marri M. J.; Kalmbach J.; Baltrun M.; Neuba A.; Schoch R.; Hohloch S.; Heinze K.; Seitz M.; Kühn O.; Lochbrunner S.; Bauer M. Distinct photodynamics of κ-N and κ-C pseudoisomeric iron(II) complexes. Chem. Commun. 2021, 57, 6640–6643. 10.1039/D1CC01716K. [DOI] [PubMed] [Google Scholar]
- Braun J. D.; Lozada I. B.; Kolodziej C.; Burda C.; Newman K. M. E.; van Lierop J.; Davis R. L.; Herbert D. E. Iron(II) coordination complexes with panchromatic absorption and nanosecond charge-transfer excited state lifetimes. Nat. Chem. 2019, 11, 1144–1150. 10.1038/s41557-019-0357-z. [DOI] [PubMed] [Google Scholar]
- Mukherjee S.; Torres D. E.; Jakubikova E. HOMO inversion as a strategy for improving the light-absorption properties of Fe(II) chromophores. Chem. Sci. 2017, 8, 8115–8126. 10.1039/C7SC02926H. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ashley D. C.; Jakubikova E. Ironing out the Photochemical and Spin-Crossover Behavior of Fe(II) Coordination Compounds with Computational Chemistry. Coord. Chem. Rev. 2017, 337, 97–111. 10.1016/j.ccr.2017.02.005. [DOI] [Google Scholar]
- Braun J. D.; Lozada I. B.; Herbert D. E. In Pursuit of Panchromatic Absorption in Metal Coordination Complexes: Experimental Delineation of the HOMO Inversion Model Using Pseudo-Octahedral Complexes of Diarylamido Ligands. Inorg. Chem. 2020, 59, 17746–17757. 10.1021/acs.inorgchem.0c02973. [DOI] [PubMed] [Google Scholar]
- Larsen C. B.; Braun J. D.; Lozada I. B.; Kunnus K.; Biasin E.; Kolodziej C.; Burda C.; Cordones A. A.; Gaffney K. J.; Herbert D. E. Reduction of Electron Repulsion in Highly Covalent Fe-Amido Complexes Counteracts the Impact of a Weak Ligand Field on Excited-State Ordering. J. Am. Chem. Soc. 2021, 143, 20645–20656. 10.1021/jacs.1c06429. [DOI] [PubMed] [Google Scholar]
- Mukherjee S.; Bowman D. N.; Jakubikova E. Cyclometalated Fe(II) Complexes as Sensitizers in Dye-Sensitized Solar Cells. Inorg. Chem. 2015, 54, 560–569. 10.1021/ic502438g. [DOI] [PubMed] [Google Scholar]
- Mukherjee S.; Liu C.; Jakubikova E. Comparison of Interfacial Electron Transfer Efficiency in [Fe(ctpy)2]2+–TiO2 and [Fe(cCNC)2]2+–TiO2 Assemblies: Importance of Conformational Sampling. J. Phys. Chem. A 2018, 122, 1821–1830. 10.1021/acs.jpca.7b10932. [DOI] [PubMed] [Google Scholar]
- Dixon I. M.; Boissard G.; Whyte H.; Alary F.; Heully J. L. Computational Estimate of the Photophysical Capabilities of Four Series of Organometallic Iron(II) Complexes. Inorg. Chem. 2016, 55, 5089–5091. 10.1021/acs.inorgchem.6b00223. [DOI] [PubMed] [Google Scholar]
- Dixon I. M.; Alary F.; Boggio-Pasqua M.; Heully J. L. Reversing the relative 3MLCT-3MC order in Fe(II) complexes using cyclometallating ligands: a computational study aiming at luminescent Fe(II) complexes. Dalton Trans. 2015, 44, 13498–13503. 10.1039/C5DT01214G. [DOI] [PubMed] [Google Scholar]
- Dixon I. M.; Khan S.; Alary F.; Boggio-Pasqua M.; Heully J. L. Probing the photophysical capability of mono and bis(cyclometallated) Fe(II) polypyridine complexes using inexpensive ground state DFT. Dalton Trans. 2014, 43, 15898–15905. 10.1039/C4DT01939C. [DOI] [PubMed] [Google Scholar]
- Dixon I. M.; Alary F.; Boggio-Pasqua M.; Heully J. L. The (N4C2)2- Donor Set as Promising Motif for Bis(tridentate) Iron(II) Photoactive Compounds. Inorg. Chem. 2013, 52, 13369–13374. 10.1021/ic402453p. [DOI] [PubMed] [Google Scholar]
- Ashley D. C.; Mukherjee S.; Jakubikova E. Designing air-stable cyclometalated Fe(II) complexes: stabilization via electrostatic effects. Dalton Trans. 2019, 48, 374. 10.1039/C8DT04402C. [DOI] [PubMed] [Google Scholar]
- Estrada-Montaño A. S.; Ryabov A. D.; Gries A.; Gaiddon C.; Le Lagadec R. Iron(III) Pincer Complexes as a Strategy for Anticancer Studies. Eur. J. Inorg. Chem. 2017, 2017, 1673–1678. 10.1002/ejic.201601350. [DOI] [Google Scholar]
- Steube J.; Burkhardt L.; Päpcke A.; Moll J.; Zimmer P.; Schoch R.; Wölper C.; Heinze K.; Lochbrunner S.; Bauer M. Excited-State Kinetics of an Air-Stable Cyclometalated Iron(II) Complex. Chem.—Eur. J. 2019, 25, 11826–11830. 10.1002/chem.201902488. [DOI] [PubMed] [Google Scholar]
- Tang Z.; Chang X.-Y.; Wan Q.; Wang J.; Ma C.; Law K.-C.; Liu Y.; Che C.-M. Bis(tridentate) Iron(II) Complexes with a Cyclometalating Unit: Photophysical Property Enhancement with Combinatorial Strong Ligand Field Effect. Organometallics 2020, 39, 2791–2802. 10.1021/acs.organomet.0c00149. [DOI] [Google Scholar]
- Leis W.; Argüello Cordero M. A.; Lochbrunner S.; Schubert H.; Berkefeld A. A Photoreactive Iron(II) Complex Luminophore. J. Am. Chem. Soc. 2022, 144, 1169–1173. 10.1021/jacs.1c13083. [DOI] [PubMed] [Google Scholar]
- Treadway J. A.; Strouse G. F.; Ruminski R. R.; Meyer T. J. Long-Lived Near-Infrared MLCT Emitters. Inorg. Chem. 2001, 40, 4508–4509. 10.1021/ic010660k. [DOI] [PubMed] [Google Scholar]
- Santoro A.; Kershaw Cook L. J.; Kulmaczewski R.; Barrett S. A.; Cespedes O.; Halcrow M. A. Iron(II) Complexes of Tridentate Indazolylpyridine Ligands: Enhanced Spin-Crossover Hysteresis and Ligand-Based Fluorescence. Inorg. Chem. 2015, 54, 682–693. 10.1021/ic502726q. [DOI] [PubMed] [Google Scholar]
- Chartres J. D.; Busby M.; Riley M. J.; Davis J. J.; Bernhardt P. V. A Turn-on Fluorescent Iron Complex and Its Cellular Uptake. Inorg. Chem. 2011, 50, 9178–9183. 10.1021/ic201495r. [DOI] [PubMed] [Google Scholar]
- Kurz H.; Schötz K.; Papadopoulos I.; Heinemann F. W.; Maid H.; Guldi D. M.; Köhler A.; Hörner G.; Weber B. A Fluorescence-Detected Coordination-Induced Spin State Switch. J. Am. Chem. Soc. 2021, 143, 3466–3480. 10.1021/jacs.0c12568. [DOI] [PubMed] [Google Scholar]
- Mann K. R.; Cimolino M.; Geoffroy G. L.; Hammond G. S.; Orio A. A.; Albertin G.; Gray H. B. Electronic structures and spectra of hexakisphenylisocyanide complexes of Cr0, Mo0, W0, MnI, and MnII. Inorg. Chim. Acta 1976, 16, 97–101. 10.1016/S0020-1693(00)91697-9. [DOI] [Google Scholar]
- Manuta D. M.; Lees A. J. Emission and photochemistry of M(CO)4(diimine) (M = chromium, molybdenum, tungsten) complexes in room-temperature solution. Inorg. Chem. 1986, 25, 1354–1359. 10.1021/ic00229a012. [DOI] [Google Scholar]
- Röhrs M.; Escudero D. Multiple Anti-Kasha Emissions in Transition-Metal Complexes. J. Phys. Chem. Lett. 2019, 10, 5798–5804. 10.1021/acs.jpclett.9b02477. [DOI] [PubMed] [Google Scholar]
- Farrell I. R.; Matousek P.; Towrie M.; Parker A. W.; Grills D. C.; George M. W.; Vlcek A. Direct observation of competitive ultrafast CO dissociation and relaxation of an MLCT excited state: Picosecond time-resolved infrared spectroscopic study of Cr(CO)4(2,2′-bipyridine). Inorg. Chem. 2002, 41, 4318–4323. 10.1021/ic020218h. [DOI] [PubMed] [Google Scholar]
- Büldt L. A.; Guo X.; Vogel R.; Prescimone A.; Wenger O. S. A tris(diisocyanide)chromium(0) complex is a luminescent analog of Fe(2,2′-bipyridine)32+. J. Am. Chem. Soc. 2017, 139, 985–992. 10.1021/jacs.6b11803. [DOI] [PubMed] [Google Scholar]
- Herr P.; Glaser F.; Büldt L. A.; Larsen C. B.; Wenger O. S. Long-lived, strongly emissive, and highly reducing excited states in Mo(0) complexes with chelating isocyanides. J. Am. Chem. Soc. 2019, 141, 14394–14402. 10.1021/jacs.9b07373. [DOI] [PubMed] [Google Scholar]
- Wegeberg C.; Häussinger D.; Wenger O. S. Pyrene-Decoration of a Chromium(0) Tris(diisocyanide) Enhances Excited State Delocalization: A Strategy to Improve the Photoluminescence of 3d6 Metal Complexes. J. Am. Chem. Soc. 2021, 143, 15800–15811. 10.1021/jacs.1c07345. [DOI] [PubMed] [Google Scholar]
- Strouse G. F.; Schoonover J. R.; Duesing R.; Boyde S.; Jones W. E. Jr.; Meyer T. J. Influence of Electronic Delocalization In Metal-to-Ligand Charge Transfer Excited States. Inorg. Chem. 1995, 34, 473–487. 10.1021/ic00106a009. [DOI] [Google Scholar]
- Boyde S.; Strouse G. F.; Jones W. E.; Meyer T. J. The effect on MLCT excited states of electronic delocalization in the acceptor ligand. J. Am. Chem. Soc. 1990, 112, 7395–7396. 10.1021/ja00176a049. [DOI] [Google Scholar]
- Damrauer N. H.; Boussie T. R.; Devenney M.; McCusker J. K. Effects of Intraligand Electron Delocalization, Steric Tuning, and Excited-State Vibronic Coupling on the Photophysics of Aryl-Substituted Bipyridyl Complexes of Ru(II). J. Am. Chem. Soc. 1997, 119, 8253–8268. 10.1021/ja971321m. [DOI] [Google Scholar]
- Wegeberg C.; Wenger O. S. Luminescent chromium(0) and manganese(I) complexes. Dalton Trans. 2022, 51, 1297–1302. 10.1039/D1DT03763C. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Herr P.; Kerzig C.; Larsen C. B.; Häussinger D.; Wenger O. S. Manganese(I) complexes with metal-to-ligand charge transfer luminescence and photoreactivity. Nat. Chem. 2021, 13, 956–962. 10.1038/s41557-021-00744-9. [DOI] [PubMed] [Google Scholar]
- Bilger J. B.; Kerzig C.; Larsen C. B.; Wenger O. S. A Photorobust Mo(0) Complex Mimicking [Os(2,2′-bipyridine)3]2+ and Its Application in Red-to-Blue Upconversion. J. Am. Chem. Soc. 2021, 143, 1651–1663. 10.1021/jacs.0c12805. [DOI] [PubMed] [Google Scholar]
- Herr P.; Schwab A.; Kupfer S.; Wenger O. S. Deep-Red Luminescent Molybdenum(0) Complexes with Bi- and Tridentate Isocyanide Chelate Ligands. ChemPhotoChem. 2022, 6, e202200052. 10.1002/cptc.202200052. [DOI] [Google Scholar]
- Sattler W.; Henling L. M.; Winkler J. R.; Gray H. B. Bespoke photoreductants: tungsten arylisocyanides. J. Am. Chem. Soc. 2015, 137, 1198–1205. 10.1021/ja510973h. [DOI] [PubMed] [Google Scholar]
- Ossinger S.; Prescimone A.; Häussinger D.; Wenger O. S. Manganese(I) Complex with Monodentate Arylisocyanide Ligands Shows Photodissociation Instead of Luminescence. Inorg. Chem. 2022, 61, 10533–10547. 10.1021/acs.inorgchem.2c01438. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ford P. C. From curiosity to applications. A personal perspective on inorganic photochemistry. Chem. Sci. 2016, 7, 2964–2986. 10.1039/C6SC00188B. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yarranton J. T.; McCusker J. K. Ligand-Field Spectroscopy of Co(III) Complexes and the Development of a Spectrochemical Series for Low-Spin d6 Charge-Transfer Chromophores. J. Am. Chem. Soc. 2022, 144, 12488–12500. 10.1021/jacs.2c04945. [DOI] [PubMed] [Google Scholar]
- Miskowski V. M.; Gray H. B.; Wilson R. B.; Solomon E. I. Position of the 3T1g←1A1g transition in hexacyanocobaltate(III). Analysis of absorption and emission results. Inorg. Chem. 1979, 18, 1410–1412. 10.1021/ic50195a057. [DOI] [Google Scholar]
- Viaene L.; D’Olieslager J.; Ceulemans A.; Vanquickenborne L. G. Excited-state spectroscopy of hexacyanocobaltate(III). J. Am. Chem. Soc. 1979, 101, 1405–1409. 10.1021/ja00500a009. [DOI] [Google Scholar]
- Viaene L.; D’Olieslager J. Luminescence from and absorption by the 3T1g level of the hexacyanocobaltate(III) ion. Inorg. Chem. 1987, 26, 960–962. 10.1021/ic00253a039. [DOI] [Google Scholar]
- Kaufhold S.; Rosemann N. W.; Chabera P.; Lindh L.; Losada I. B.; Uhlig J.; Pascher T.; Strand D.; Wärnmark K.; Yartsev A.; Persson P. Microsecond Photoluminescence and Photoreactivity of a Metal-Centered Excited State in a Hexacarbene-Co(III) Complex. J. Am. Chem. Soc. 2021, 143, 1307–1312. 10.1021/jacs.0c12151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pal A. K.; Li C. F.; Hanan G. S.; Zysman-Colman E. Blue-emissive cobalt(III) complexes and their use in the photocatalytic trifluoromethylation of polycyclic aromatic hydrocarbons. Angew. Chem., Int. Ed. 2018, 57, 8027–8031. 10.1002/anie.201802532. [DOI] [PubMed] [Google Scholar]
- Wenger O. S. A bright future for photosensitizers. Nat. Chem. 2020, 12, 323–324. 10.1038/s41557-020-0448-x. [DOI] [PubMed] [Google Scholar]
- Zhang Y.; Petersen J. L.; Milsmann C. A Luminescent Zirconium(IV) Complex as a Molecular Photosensitizer for Visible Light Photoredox Catalysis. J. Am. Chem. Soc. 2016, 138, 13115–13118. 10.1021/jacs.6b05934. [DOI] [PubMed] [Google Scholar]
- Zhang Y.; Lee T. S.; Favale J. M.; Leary D. C.; Petersen J. L.; Scholes G. D.; Castellano F. N.; Milsmann C. Delayed fluorescence from a zirconium(IV) photosensitizer with ligand-to-metal charge-transfer excited states. Nat. Chem. 2020, 12, 345–352. 10.1038/s41557-020-0430-7. [DOI] [PubMed] [Google Scholar]
- Harris J. P.; Reber C.; Colmer H. E.; Jackson T. A.; Forshaw A. P.; Smith J. M.; Kinney R. A.; Telser J. Near-infrared 2Eg → 4A2g and visible LMCT luminescence from a molecular bis-(tris(carbene)borate) manganese(IV) complex. Can. J. Chem. 2017, 95, 547–552. 10.1139/cjc-2016-0607. [DOI] [Google Scholar]
- Juliá F. Ligand-to-Metal Charge Transfer (LMCT) Photochemistry at 3d-Metal Complexes: An Emerging Tool for Sustainable Organic Synthesis. ChemCatChem. 2022, 14, e202200916. 10.1002/cctc.202200916. [DOI] [Google Scholar]
- Moser M.; Wucher B.; Kunz D.; Rominger F. 1,8-Bis(imidazolin-2-yliden-1-yl)carbazolide (bimca): A New CNC Pincer-Type Ligand with Strong Electron-Donating Properties. Facile Oxidative Addition of Methyl Iodide to Rh(bimca)(CO). Organometallics 2007, 26, 1024–1030. 10.1021/om060952z. [DOI] [Google Scholar]
- Sinha N.; Pfund B.; Wegeberg C.; Prescimone A.; Wenger O. S. Cobalt(III) Carbene Complex with an Electronic Excited-State Structure Similar to Cyclometalated Iridium(III) Compounds. J. Am. Chem. Soc. 2022, 144, 9859–9873. 10.1021/jacs.2c02592. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mills I. N.; Porras J. A.; Bernhard S. Judicious Design of Cationic, Cyclometalated IrIII Complexes for Photochemical Energy Conversion and Optoelectronics. Acc. Chem. Res. 2018, 51, 352–364. 10.1021/acs.accounts.7b00375. [DOI] [PubMed] [Google Scholar]
- Aydogan A.; Bangle R. E.; Cadranel A.; Turlington M. D.; Conroy D. T.; Cauët E.; Singleton M. L.; Meyer G. J.; Sampaio R. N.; Elias B.; Troian-Gautier L. Accessing Photoredox Transformations with an Iron(III) Photosensitizer and Green Light. J. Am. Chem. Soc. 2021, 143, 15661–15673. 10.1021/jacs.1c06081. [DOI] [PubMed] [Google Scholar]
- Woodhouse M. D.; McCusker J. K. Mechanistic Origin of Photoredox Catalysis Involving Iron(II) Polypyridyl Chromophores. J. Am. Chem. Soc. 2020, 142, 16229–16233. 10.1021/jacs.0c08389. [DOI] [PubMed] [Google Scholar]
- McCusker J. K.; Walda K. N.; Dunn R. C.; Simon J. D.; Magde D.; Hendrickson D. N. Subpicosecond 1MLCT-5T2 Intersystem Crossing of Low-Spin Polypyridyl Ferrous Complexes. J. Am. Chem. Soc. 1993, 115, 298–307. 10.1021/ja00054a043. [DOI] [Google Scholar]
- Zhang W. K.; Gaffney K. J. Mechanistic Studies of Photoinduced Spin Crossover and Electron Transfer in Inorganic Complexes. Acc. Chem. Res. 2015, 48, 1140–1148. 10.1021/ar500407p. [DOI] [PubMed] [Google Scholar]
- Nair S. S.; Bysewski O. A.; Kupfer S.; Wächtler M.; Winter A.; Schubert U. S.; Dietzek B. Excitation Energy-Dependent Branching Dynamics Determines Photostability of Iron(II)–Mesoionic Carbene Complexes. Inorg. Chem. 2021, 60, 9157–9173. 10.1021/acs.inorgchem.1c01166. [DOI] [PubMed] [Google Scholar]
- Wenger O. S. Is Iron the New Ruthenium?. Chem.—Eur. J. 2019, 25, 6043–6052. 10.1002/chem.201806148. [DOI] [PubMed] [Google Scholar]
- Paulus B. C.; Nielsen K. C.; Tichnell C. R.; Carey M. C.; McCusker J. K. A Modular Approach to Light Capture and Synthetic Tuning of the Excited-State Properties of Fe(II)-Based Chromophores. J. Am. Chem. Soc. 2021, 143, 8086–8098. 10.1021/jacs.1c02451. [DOI] [PubMed] [Google Scholar]
- Cheng Y.; Yang Q.; He J.; Zou W.; Liao K.; Chang X.; Zou C.; Lu W. The energy gap law for NIR-phosphorescent Cr(III) complexes. Dalton Trans. 2023, 10.1039/D2DT02872G. [DOI] [PubMed] [Google Scholar]
- Sinha N.; Jiménez J. R.; Pfund B.; Prescimone A.; Piguet C.; Wenger O. S. A near-infrared-II emissive chromium(III) complex. Angew. Chem., Int. Ed. 2021, 60, 23722–23728. 10.1002/anie.202106398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ashley D. C.; Jakubikova E. Tuning the Redox Potentials and Ligand Field Strength of Fe(II) Polypyridines: The Dual π-Donor and π-Acceptor Character of Bipyridine. Inorg. Chem. 2018, 57, 9907–9917. 10.1021/acs.inorgchem.8b01002. [DOI] [PubMed] [Google Scholar]
- Law K.-C.; Tang Z.; Wu L.; Wan Q.; To W.-P.; Chang X.; Low K.-H.; Liu Y.; Che C.-M. Cyclometalated Iron and Ruthenium Complexes Supported by a Tetradentate Ligand Scaffold with Mixed O, N, and C Donor Atoms: Synthesis, Structures, and Excited-State Properties. Organometallics 2022, 41, 418–429. 10.1021/acs.organomet.1c00677. [DOI] [Google Scholar]
- Caspar J. V.; Kober E. M.; Sullivan B. P.; Meyer T. J. Application of the energy gap law to the decay of charge-transfer excited states. J. Am. Chem. Soc. 1982, 104, 630–632. 10.1021/ja00366a051. [DOI] [Google Scholar]
- Islam A.; Ikeda N.; Yoshimura A.; Ohno T. Nonradiative Transition of Phosphorescent Charge-Transfer States of Ruthenium(II)-to-2,2‘-Biquinoline and Ruthenium(II)-to-2,2‘:6‘,2”-Terpyridine in the Solid State. Inorg. Chem. 1998, 37, 3093–3098. 10.1021/ic9702429. [DOI] [Google Scholar]
- Medlycott E. A.; Hanan G. S. Designing Tridentate Ligands for Ruthenium(II) Complexes with Prolonged Room Temperature Luminescence Lifetimes. Chem. Soc. Rev. 2005, 34, 133–142. 10.1039/b316486c. [DOI] [PubMed] [Google Scholar]
- Kitzmann W. R.; Moll J.; Heinze K. Spin-flip luminescence. Photochem. Photobiol. Sci. 2022, 21, 1309–1331. 10.1007/s43630-022-00186-3. [DOI] [PubMed] [Google Scholar]
- Otto S.; Grabolle M.; Förster C.; Kreitner C.; Resch-Genger U.; Heinze K. [Cr(ddpd)2]3+: A Molecular, Water-Soluble, Highly NIR-Emissive Ruby Analogue. Angew. Chem., Int. Ed. 2015, 54, 11572–11576. 10.1002/anie.201504894. [DOI] [PubMed] [Google Scholar]
- Jimenez J.-R.; Doistau B.; Cruz C. M.; Besnard C.; Cuerva J. M.; Campana A. G.; Piguet C. Chiral Molecular Ruby [Cr(dqp)2]3+ with Long-Lived Circularly Polarized Luminescence. J. Am. Chem. Soc. 2019, 141, 13244–13252. 10.1021/jacs.9b06524. [DOI] [PubMed] [Google Scholar]
- Abrahamsson M.; Jäger M.; Osterman T.; Eriksson L.; Persson P.; Becker H. C.; Johansson O.; Hammarström L. A 3.0 μs Room Temperature Excited State Lifetime of a Bistridentate RuII-polypyridine Complex for Rod-Like Molecular Arrays. J. Am. Chem. Soc. 2006, 128, 12616–12617. 10.1021/ja064262y. [DOI] [PubMed] [Google Scholar]
- Reichenauer F.; Wang C.; Förster C.; Boden P.; Ugur N.; Báez-Cruz R.; Kalmbach J.; Carrella L. M.; Rentschler E.; Ramanan C.; Niedner-Schatteburg G.; Gerhards M.; Seitz M.; Resch-Genger U.; Heinze K. Strongly Red-Emissive Molecular Ruby [Cr(bpmp)2]3+ Surpasses [Ru(bpy)3]2+. J. Am. Chem. Soc. 2021, 143, 11843–11855. 10.1021/jacs.1c05971. [DOI] [PubMed] [Google Scholar]
- Beaudelot J.; Oger S.; Peruško S.; Phan T.-A.; Teunens T.; Moucheron C.; Evano G. Photoactive Copper Complexes: Properties and Applications. Chem. Rev. 2022, 122, 16365–16609. 10.1021/acs.chemrev.2c00033. [DOI] [PubMed] [Google Scholar]
- McCusker C. E.; Castellano F. N. Design of a Long-Lifetime, Earth-Abundant, Aqueous Compatible Cu(I) Photosensitizer Using Cooperative Steric Effects. Inorg. Chem. 2013, 52, 8114–8120. 10.1021/ic401213p. [DOI] [PubMed] [Google Scholar]
- Rosko M. C.; Wells K. A.; Hauke C. E.; Castellano F. N. Next Generation Cuprous Phenanthroline MLCT Photosensitizer Featuring Cyclohexyl Substituents. Inorg. Chem. 2021, 60, 8394–8403. 10.1021/acs.inorgchem.1c01242. [DOI] [PubMed] [Google Scholar]
- Schrauben J. N.; Dillman K. L.; Beck W. F.; McCusker J. K. Vibrational coherence in the excited state dynamics of Cr(acac)3: probing the reaction coordinate for ultrafast intersystem crossing. Chem. Sci. 2010, 1, 405–410. 10.1039/c0sc00262c. [DOI] [Google Scholar]
- Büldt L. A.; Guo X.; Prescimone A.; Wenger O. S. A Molybdenum(0) Isocyanide Analogue of Ru(2,2′-Bipyridine)32+: A Strong Reductant for Photoredox Catalysis. Angew. Chem., Int. Ed. 2016, 55, 11247–11250. 10.1002/anie.201605571. [DOI] [PubMed] [Google Scholar]
- Sawicka N.; Craze C. J.; Horton P. N.; Coles S. J.; Richards E.; Pope S. J. A. Long-lived, near-IR emission from Cr(III) under ambient conditions. Chem. Commun. 2022, 58, 5733–5736. 10.1039/D2CC01434C. [DOI] [PubMed] [Google Scholar]
- Stein L.; Boden P.; Naumann R.; Förster C.; Niedner-Schatteburg G.; Heinze K. The overlooked NIR luminescence of Cr(ppy)3. Chem. Commun. 2022, 58, 3701–3704. 10.1039/D2CC00680D. [DOI] [PubMed] [Google Scholar]
- Jiménez J.-R.; Poncet M.; Míguez-Lago S.; Grass S.; Lacour J.; Besnard C.; Cuerva J. M.; Campaña A. G.; Piguet C. Bright Long-Lived Circularly Polarized Luminescence in Chiral Chromium(III) Complexes. Angew. Chem., Int. Ed. 2021, 60, 10095–10102. 10.1002/anie.202101158. [DOI] [PubMed] [Google Scholar]
- Ogawa T.; Sinha N.; Pfund B.; Prescimone A.; Wenger O. S. Molecular Design Principles to Elongate the Metal-to-Ligand Charge Transfer Excited-State Lifetimes of Square-Planar Nickel(II) Complexes. J. Am. Chem. Soc. 2022, 144, 21948–21960. 10.1021/jacs.2c08838. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hamze R.; Peltier J. L.; Sylvinson D.; Jung M.; Cardenas J.; Haiges R.; Soleilhavoup M.; Jazzar R.; Djurovich P. I.; Bertrand G.; Thompson M. E. Eliminating nonradiative decay in CuI emitters: > 99% quantum efficiency and microsecond lifetime. Science 2019, 363, 601–606. 10.1126/science.aav2865. [DOI] [PubMed] [Google Scholar]
- Dorn M.; Kalmbach J.; Boden P.; Kruse A.; Dab C.; Reber C.; Niedner-Schatteburg G.; Lochbrunner S.; Gerhards M.; Seitz M.; Heinze K. Ultrafast and long-time excited state kinetics of an NIR-emissive vanadium(III) complex I: synthesis, spectroscopy and static quantum chemistry. Chem. Sci. 2021, 12, 10780–10790. 10.1039/D1SC02137K. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dorn M.; Kalmbach J.; Boden P.; Päpcke A.; Gómez S.; Förster C.; Kuczelinis F.; Carrella L. M.; Büldt L.; Bings N.; Rentschler E.; Lochbrunner S.; González L.; Gerhards M.; Seitz M.; Heinze K. A vanadium(III) complex with blue and NIR-II spin-flip luminescence in solution. J. Am. Chem. Soc. 2020, 142, 7947–7955. 10.1021/jacs.0c02122. [DOI] [PubMed] [Google Scholar]
- Wang C.; Otto S.; Dorn M.; Kreidt E.; Lebon J.; Srsan L.; Di Martino-Fumo P.; Gerhards M.; Resch-Genger U.; Seitz M.; Heinze K. Deuterated Molecular Ruby with Record Luminescence Quantum Yield. Angew. Chem., Int. Ed. 2018, 57, 1112–1116. 10.1002/anie.201711350. [DOI] [PubMed] [Google Scholar]
- Maurer A. B.; Meyer G. J. Stark Spectroscopic Evidence that a Spin Change Accompanies Light Absorption in Transition Metal Polypyridyl Complexes. J. Am. Chem. Soc. 2020, 142, 6847–6851. 10.1021/jacs.9b13602. [DOI] [PubMed] [Google Scholar]
- Bünzli J.-C. G. On the design of highly luminescent lanthanide complexes. Coord. Chem. Rev. 2015, 293–294, 19–47. 10.1016/j.ccr.2014.10.013. [DOI] [Google Scholar]
- Browne W. R.; O’Connor C. M.; Killeen J. S.; Guckian A. L.; Burke M.; James P.; Burke M.; Vos J. G. Routes to Regioselective Deuteriation of Heteroaromatic Compounds. Inorg. Chem. 2002, 41, 4245–4251. 10.1021/ic020226y. [DOI] [PubMed] [Google Scholar]
- Büldt L. A.; Wenger O. S. Chromium(0), Molybdenum(0), and Tungsten(0) Isocyanide Complexes as Luminophores and Photosensitizers with Long-Lived Excited States. Angew. Chem., Int. Ed. 2017, 56, 5676–5682. 10.1002/anie.201701210. [DOI] [PubMed] [Google Scholar]
- Dongare P.; Myron B. D. B.; Wang L.; Thompson D. W.; Meyer T. J. [Ru(bpy)3]2+* revisited. Is it localized or delocalized? How does it decay?. Coord. Chem. Rev. 2017, 345, 86–107. 10.1016/j.ccr.2017.03.009. [DOI] [Google Scholar]
- Wenger O. S.; Güdel H. U. Optical spectroscopy of CrCl63– doped Cs2NaScCl6: Broadband near-infrared luminescence and Jahn-Teller effect. J. Chem. Phys. 2001, 114, 5832–5841. 10.1063/1.1352730. [DOI] [Google Scholar]
- Kitzmann W. R.; Heinze K. Charge-Transfer and Spin-Flip States: Thriving as Complements. Angew. Chem., Int. Ed. 2023, 10.1002/anie.202213207. [DOI] [PubMed] [Google Scholar]
- Arias-Rotondo D. M.; McCusker J. K. The photophysics of photoredox catalysis: a roadmap for catalyst design. Chem. Soc. Rev. 2016, 45, 5803–5820. 10.1039/C6CS00526H. [DOI] [PubMed] [Google Scholar]
- Strieth-Kalthoff F.; Glorius F. Triplet Energy Transfer Photocatalysis: Unlocking the Next Level. Chem. 2020, 6, 1888–1903. 10.1016/j.chempr.2020.07.010. [DOI] [Google Scholar]
- Gualandi A.; Marchini M.; Mengozzi L.; Natali M.; Lucarini M.; Ceroni P.; Cozzi P. G. Organocatalytic Enantioselective Alkylation of Aldehydes with [Fe(bpy)3]Br2 Catalyst and Visible Light. ACS Catal. 2015, 5, 5927–5931. 10.1021/acscatal.5b01573. [DOI] [Google Scholar]
- Parisien-Collette S.; Hernandez-Perez A. C.; Collins S. K. Photochemical Synthesis of Carbazoles Using an [Fe(phen)3](NTf2)2/O2 Catalyst System: Catalysis toward Sustainability. Org. Lett. 2016, 18, 4994–4997. 10.1021/acs.orglett.6b02456. [DOI] [PubMed] [Google Scholar]
- Larsen C. B.; Wenger O. S. Photoredox Catalysis with Metal Complexes Made from Earth-Abundant Elements. Chem.—Eur. J. 2018, 24, 2039–2058. 10.1002/chem.201703602. [DOI] [PubMed] [Google Scholar]
- Zhou W.-J.; Wu X.-D.; Miao M.; Wang Z.-H.; Chen L.; Shan S.-Y.; Cao G.-M.; Yu D.-G. Light Runs Across Iron Catalysts in Organic Transformations. Chem.—Eur. J. 2020, 26, 15052–15064. 10.1002/chem.202000508. [DOI] [PubMed] [Google Scholar]
- Sattler W.; Ener M. E.; Blakemore J. D.; Rachford A. A.; LaBeaume P. J.; Thackeray J. W.; Cameron J. F.; Winkler J. R.; Gray H. B. Generation of Powerful Tungsten Reductants by Visible Light Excitation. J. Am. Chem. Soc. 2013, 135, 10614–10617. 10.1021/ja4047119. [DOI] [PubMed] [Google Scholar]
- Barth A. T.; Morales M.; Winkler J. R.; Gray H. B. Photoredox Catalysis Mediated by Tungsten(0) Arylisocyanides in 1,2-Difluorobenzene. Inorg. Chem. 2022, 61, 7251–7255. 10.1021/acs.inorgchem.1c03767. [DOI] [PubMed] [Google Scholar]
- Fajardo J. Jr.; Barth A. T.; Morales M.; Takase M. K.; Winkler J. R.; Gray H. B. Photoredox Catalysis Mediated by Tungsten(0) Arylisocyanides. J. Am. Chem. Soc. 2021, 143, 19389–19398. 10.1021/jacs.1c07617. [DOI] [PubMed] [Google Scholar]
- Glaser F.; Wenger O. S. Red Light-Based Dual Photoredox Strategy Resembling the Z-Scheme of Natural Photosynthesis. JACS Au 2022, 2, 1488–1503. 10.1021/jacsau.2c00265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Glaser F.; Wenger O. S. Sensitizer-controlled photochemical reactivity via upconversion of red light. Chem. Sci. 2023, 14, 149–161. 10.1039/D2SC05229F. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lazorski M. S.; Castellano F. N. Advances in the Light Conversion Properties of Cu(I)-Based Photosensitizers. Polyhedron 2014, 82, 57–70. 10.1016/j.poly.2014.04.060. [DOI] [Google Scholar]
- Housecroft C. E.; Constable E. C. The emergence of copper(I)-based dye sensitized solar cells. Chem. Soc. Rev. 2015, 44, 8386–8398. 10.1039/C5CS00215J. [DOI] [PubMed] [Google Scholar]
- Hossain A.; Bhattacharyya A.; Reiser O. Copper’s rapid ascent in visible-light photoredox catalysis. Science 2019, 364, 450–461. 10.1126/science.aav9713. [DOI] [PubMed] [Google Scholar]
- Czerwieniec R.; Leitl M. J.; Homeier H. H. H.; Yersin H. Cu(I) Complexes - Thermally Activated Delayed Fluorescence. Photophysical Approach and Material Design. Coord. Chem. Rev. 2016, 325, 2–28. 10.1016/j.ccr.2016.06.016. [DOI] [Google Scholar]
- Gernert M.; Balles-Wolf L.; Kerner F.; Müller U.; Schmiedel A.; Holzapfel M.; Marian C. M.; Pflaum J.; Lambert C.; Steffen A. Cyclic (amino)(aryl)carbenes Enter the Field of Chromophore Ligands: Expanded π System Leads to Unusually Deep Red Emitting CuI Compounds. J. Am. Chem. Soc. 2020, 142, 8897–8909. 10.1021/jacs.0c02234. [DOI] [PubMed] [Google Scholar]
- Li T.-y.; Muthiah Ravinson D. S.; Haiges R.; Djurovich P. I.; Thompson M. E. Enhancement of the Luminescent Efficiency in Carbene-Au(I)-Aryl Complexes by the Restriction of Renner–Teller Distortion and Bond Rotation. J. Am. Chem. Soc. 2020, 142, 6158–6172. 10.1021/jacs.9b13755. [DOI] [PubMed] [Google Scholar]
- Shi S.; Jung M. C.; Coburn C.; Tadle A.; Sylvinson M. R D.; Djurovich P. I.; Forrest S. R.; Thompson M. E. Highly Efficient Photo- and Electroluminescence from Two-Coordinate Cu(I) Complexes Featuring Nonconventional N-Heterocyclic Carbenes. J. Am. Chem. Soc. 2019, 141, 3576–3588. 10.1021/jacs.8b12397. [DOI] [PubMed] [Google Scholar]
- Malzkuhn S.; Wenger O. S. Luminescent Ni(0) complexes. Coord. Chem. Rev. 2018, 359, 52–56. 10.1016/j.ccr.2018.01.003. [DOI] [Google Scholar]
- Büldt L. A.; Larsen C. B.; Wenger O. S. Luminescent Ni0 Diisocyanide Chelates as Analogues of CuI Diimine Complexes. Chem.—Eur. J. 2017, 23, 8577–8580. 10.1002/chem.201700103. [DOI] [PubMed] [Google Scholar]
- Bizzarri C.; Spuling E.; Knoll D. M.; Volz D.; Bräse S. Sustainable metal complexes for organic light-emitting diodes (OLEDs). Coord. Chem. Rev. 2018, 373, 49–82. 10.1016/j.ccr.2017.09.011. [DOI] [Google Scholar]
- Tungulin D.; Leier J.; Carter A. B.; Powell A. K.; Albuquerque R. Q.; Unterreiner A. N.; Bizzarri C. Chasing BODIPY: Enhancement of Luminescence in Homoleptic Bis(dipyrrinato) ZnII Complexes Utilizing Symmetric and Unsymmetrical Dipyrrins. Chem.—Eur. J. 2019, 25, 3816–3827. 10.1002/chem.201806330. [DOI] [PubMed] [Google Scholar]
- Lozada I. B.; Braun J. D.; Williams J. A. G.; Herbert D. E. Yellow-Emitting, Pseudo-Octahedral Zinc Complexes of Benzannulated N^N^O Pincer-Type Ligands. Inorg. Chem. 2022, 61, 17568–17578. 10.1021/acs.inorgchem.2c02585. [DOI] [PubMed] [Google Scholar]
- Mrózek O.; Gernert M.; Belyaev A.; Mitra M.; Janiak L.; Marian C. M.; Steffen A. Ultra-Long Lived Luminescent Triplet Excited States in Cyclic (Alkyl)(amino)carbene Complexes of Zn(II) Halides. Chem.—Eur. J. 2022, 28, e202201114. 10.1002/chem.202201114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kübler J. A.; Pfund B.; Wenger O. S. Zinc(II) Complexes with Triplet Charge-Transfer Excited States Enabling Energy-Transfer Catalysis, Photoinduced Electron Transfer, and Upconversion. JACS Au 2022, 2, 2367–2380. 10.1021/jacsau.2c00442. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sakai Y.; Sagara Y.; Nomura H.; Nakamura N.; Suzuki Y.; Miyazaki H.; Adachi C. Zinc complexes exhibiting highly efficient thermally activated delayed fluorescence and their application to organic light-emitting diodes. Chem. Commun. 2015, 51, 3181–3184. 10.1039/C4CC09403D. [DOI] [PubMed] [Google Scholar]
- Cheng G.; So G. K. M.; To W. P.; Chen Y.; Kwok C. C.; Ma C. S.; Guan X. G.; Chang X.; Kwok W. M.; Che C. M. Luminescent zinc(II) and copper(I) complexes for high-performance solution-processed monochromic and white organic light-emitting devices. Chem. Sci. 2015, 6, 4623–4635. 10.1039/C4SC03161J. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kurz H.; Hörner G.; Weser O.; Li Manni G.; Weber B. Quenched Lewis Acidity: Studies on the Medium Dependent Fluorescence of Zinc(II) Complexes. Chem.—Eur. J. 2021, 27, 15159–15171. 10.1002/chem.202102086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kalinowski J.; Fattori V.; Cocchi M.; Williams J. A. G. Light-emitting devices based on organometallic platinum complexes as emitters. Coord. Chem. Rev. 2011, 255, 2401–2425. 10.1016/j.ccr.2011.01.049. [DOI] [Google Scholar]
- Malmberg R.; Venkatesan K. Conceptual advances in the preparation and excited-state properties of neutral luminescent (C^N) and (C^C*) monocyclometalated gold(III) complexes. Coord. Chem. Rev. 2021, 449, 214182. 10.1016/j.ccr.2021.214182. [DOI] [Google Scholar]
- Tang M.-C.; Chan M.-Y.; Yam V. W.-W. Molecular Design of Luminescent Gold(III) Emitters as Thermally Evaporable and Solution-Processable Organic Light-Emitting Device (OLED) Materials. Chem. Rev. 2021, 121, 7249–7279. 10.1021/acs.chemrev.0c00936. [DOI] [PubMed] [Google Scholar]
- Wenger O. S. Photoactive Nickel Complexes in Cross-Coupling Catalysis. Chem.—Eur. J. 2021, 27, 2270–2278. 10.1002/chem.202003974. [DOI] [PubMed] [Google Scholar]
- Welin E. R.; Le C.; Arias-Rotondo D. M.; McCusker J. K.; MacMillan D. W. C. Photosensitized, Energy Transfer-Mediated Organometallic Catalysis through Electronically Excited Nickel(II). Science 2017, 355, 380–384. 10.1126/science.aal2490. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tian L.; Till N. A.; Kudisch B.; MacMillan D. W. C.; Scholes G. D. Transient Absorption Spectroscopy Offers Mechanistic Insights for an Iridium/Nickel-Catalyzed C–O Coupling. J. Am. Chem. Soc. 2020, 142, 4555–4559. 10.1021/jacs.9b12835. [DOI] [PubMed] [Google Scholar]
- Cagan D. A.; Bím D.; Silva B.; Kazmierczak N. P.; McNicholas B. J.; Hadt R. G. Elucidating the Mechanism of Excited-State Bond Homolysis in Nickel–Bipyridine Photoredox Catalysts. J. Am. Chem. Soc. 2022, 144, 6516–6531. 10.1021/jacs.2c01356. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ting S. I.; Garakyaraghi S.; Taliaferro C. M.; Shields B. J.; Scholes G. D.; Castellano F. N.; Doyle A. G. 3d-d Excited States of Ni(II) Complexes Relevant to Photoredox Catalysis: Spectroscopic Identification and Mechanistic Implications. J. Am. Chem. Soc. 2020, 142, 5800–5810. 10.1021/jacs.0c00781. [DOI] [PubMed] [Google Scholar]
- Wong Y.-S.; Tang M.-C.; Ng M.; Yam V. W.-W. Toward the Design of Phosphorescent Emitters of Cyclometalated Earth-Abundant Nickel(II) and Their Supramolecular Study. J. Am. Chem. Soc. 2020, 142, 7638–7646. 10.1021/jacs.0c02172. [DOI] [PubMed] [Google Scholar]
- Lauenstein R.; Mader S. L.; Derondeau H.; Esezobor O. Z.; Block M.; Römer A. J.; Jandl C.; Riedle E.; Kaila V.; Hauer J.; Thyrhaug E.; Hess C. R. The central role of the metal ion for photoactivity: Zn- vs. Ni-Mabiq. Chem. Sci. 2021, 12, 7521–7532. 10.1039/D0SC06096H. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eskelinen T.; Buss S.; Petrovskii S. K.; Grachova E. V.; Krause M.; Kletsch L.; Klein A.; Strassert C. A.; Koshevoy I. O.; Hirva P. Photophysics and Excited State Dynamics of Cyclometalated [M(Phbpy)(CN)] (M = Ni, Pd, Pt) Complexes: A Theoretical and Experimental Study. Inorg. Chem. 2021, 60, 8777–8789. 10.1021/acs.inorgchem.1c00680. [DOI] [PubMed] [Google Scholar]
- Gandeepan P.; Müller T.; Zell D.; Cera G.; Warratz S.; Ackermann L. 3d Transition Metals for C–H Activation. Chem. Rev. 2019, 119, 2192–2452. 10.1021/acs.chemrev.8b00507. [DOI] [PubMed] [Google Scholar]
- Hockin B. M.; Li C. F.; Robertson N.; Zysman-Colman E. Photoredox Catalysts Based on Earth-Abundant Metal Complexes. Catal. Sci. Technol. 2019, 9, 889–915. 10.1039/C8CY02336K. [DOI] [Google Scholar]
- Banza Lubaba Nkulu C.; Casas L.; Haufroid V.; De Putter T.; Saenen N. D.; Kayembe-Kitenge T.; Musa Obadia P.; Kyanika Wa Mukoma D.; Lunda Ilunga J.-M.; Nawrot T. S.; Luboya Numbi O.; Smolders E.; Nemery B. Sustainability of artisanal mining of cobalt in DR Congo. Nat. Sustain. 2018, 1, 495–504. 10.1038/s41893-018-0139-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Calvão F.; McDonald C. E. A.; Bolay M. Cobalt mining and the corporate outsourcing of responsibility in the Democratic Republic of Congo. Extr. Ind. Soc. 2021, 8, 100884. 10.1016/j.exis.2021.02.004. [DOI] [Google Scholar]
- Shon J. H.; Teets T. S. Molecular Photosensitizers in Energy Research and Catalysis: Design Principles and Recent Developments. ACS Energy Lett. 2019, 4, 558–566. 10.1021/acsenergylett.8b02388. [DOI] [Google Scholar]
- Hauser A.; Reber C. Spectroscopy and Chemical Bonding in Transition Metal Complexes. Struct. Bonding (Berlin) 2016, 172, 291–312. 10.1007/430_2015_195. [DOI] [Google Scholar]
- Wood C. J.; Summers G. H.; Clark C. A.; Kaeffer N.; Braeutigam M.; Carbone L. R.; D’Amario L.; Fan K.; Farré Y.; Narbey S.; Oswald F.; Stevens L. A.; Parmenter C. D. J.; Fay M. W.; La Torre A.; Snape C. E.; Dietzek B.; Dini D.; Hammarström L.; Pellegrin Y.; Odobel F.; Sun L.; Artero V.; Gibson E. A. A comprehensive comparison of dye-sensitized NiO photocathodes for solar energy conversion. Phys. Chem. Chem. Phys. 2016, 18, 10727–10738. 10.1039/C5CP05326A. [DOI] [PubMed] [Google Scholar]
- D’Accriscio F.; Borja P.; Saffon-Merceron N.; Fustier-Boutignon M.; Mézailles N.; Nebra N. C–H Bond Trifluoromethylation of Arenes Enabled by a Robust, High-Valent Nickel(IV) Complex. Angew. Chem., Int. Ed. 2017, 56, 12898–12902. 10.1002/anie.201706237. [DOI] [PubMed] [Google Scholar]
- Milbauer M. W.; Kampf J. W.; Sanford M. S. Nickel(IV) Intermediates in Aminoquinoline-Directed C(sp2)–C(sp3) Coupling. J. Am. Chem. Soc. 2022, 144, 21030–21034. 10.1021/jacs.2c10778. [DOI] [PubMed] [Google Scholar]








