Abstract
Recently reported manganese(I) complexes with chelating arylisocyanide ligands exhibit luminescent metal-to-ligand charge-transfer (MLCT) excited states, similar to ruthenium(II) polypyridine complexes with the same d6 valence electron configuration used for many different applications in photophysics and photochemistry. However, chelating arylisocyanide ligands require substantial synthetic effort, and therefore it seemed attractive to explore the possibility of using more readily accessible monodentate arylisocyanides instead. Here, we synthesized the new Mn(I) complex [Mn(CNdippPhOMe2)6]PF6 with the known ligand CNdippPhOMe2 = 4-(3,5-dimethoxyphenyl)-2,6-diisopropylphenylisocyanide. This complex was investigated by NMR spectroscopy, single-crystal structure analysis, high-resolution electrospray ionization mass spectrometry (HR-ESI-MS) measurements, IR spectroscopy supported by density functional theory (DFT) calculations, cyclic voltammetry, and time-resolved as well as steady-state UV–vis absorption spectroscopy. The key finding is that the new Mn(I) complex is nonluminescent and instead undergoes arylisocyanide ligand loss during continuous visible laser irradiation into ligand-centered and charge-transfer absorption bands, presumably owed to the population of dissociative d–d excited states. Thus, it seems that chelating bi- or tridentate binding motifs are essential for obtaining emissive MLCT excited states in manganese(I) arylisocyanides. Our work contributes to understanding the basic properties of photoactive first-row transition metal complexes and could help advance the search for alternatives to precious metal-based luminophores, photocatalysts, and sensors.
Short abstract
We report the synthesis, characterization, and X-ray crystal structure of an octahedral manganese(I) complex with six monodentate arylisocyanide ligands that undergoes photoinduced ligand loss.
1. Introduction
Metal-to-ligand charge-transfer (MLCT) excited states play a key role in many coordination complexes and organometallic compounds because they enable a range of different applications in photophysics and photochemistry. Precious and rare elements such as ruthenium(II),1−500 osmium(II),5−9 rhenium(I),10−18 or iridium(III)15,19−25 in polypyridine or cyclometalating coordination environments often feature a long-lived MLCT excited state,26,27 whereas among first-row d6 transition metal elements, this is yet a very rare occurrence. Iron(II) is by far most investigated in this regard,28−44 yet only a handful of iron(II) complexes with MLCT lifetimes in the nanosecond time regime are known.28,29,41 Building on early reports of hexakis(arylisocyanide) manganese(I) complexes with a focus on UV–vis absorption and electrochemical properties,45−52 we recently discovered that manganese(I) complexes with chelating bi- and tridentate ligands have luminescent and photoredox active MLCT states.53 Until now, the synthesis of these chelates has remained laborious,54 and therefore it seemed attractive to explore the possibility of using more directly accessible monodentate arylisocyanides for manganese(I) complexes with luminescent MLCT states. Recently, there has been increased interest in first-row transition metal elements, partly because they are cheaper and more abundant than the traditionally used metals from the platinum group,55 and because they seem to offer ample opportunities for groundbreaking discoveries.56−59 Advances in ligand design, photophysical techniques, and theoretical understanding60 have made an unexpectedly broad range of transition metals in different oxidation states useable for photophysical or photochemical applications. In particular, luminescent first-row transition metal complexes with V,61−64 Cr,65−73 Mn,53,74−76 Fe,28,34,42,77−79 Co,71,80−800 Ni,83−86 and Cu87−96 with different types of electronically excited states featuring promising photoreactivity and photoluminescence behavior have been discovered recently.56,57,59,97,98
As early as 1977, Mann, Gray, and Hammond reported on tungsten(0) arylisocyanide complexes that showed MLCT luminescence in solution at room temperature.99 More recently, the GRAY group revisited this topic and discovered a whole range of [W(CNAryl)6],100−105 complexes with outstanding photoluminescence properties and very strong photoreducing behavior with applications in catalysis. In addition to diisopropyl substituents at the ortho-position to the ligating isocyanide functional group, the attachment of further phenyl rings at the para-position proved valuable to enhance the chemical stability and the photophysical properties of this compound class, for example in [W(CNdippPhOMe2)6] (CNdippPhOMe2 = 4-(3,5-dimethoxyphenyl)-2,6-diisopropylphenylisocyanide).101 This specific tungsten(0) complex featured a high photoluminescence quantum yield (0.42) paired with a long MLCT lifetime (1.65 μs in toluene at room temperature), hence our interest in this particular ligand for manganese(I).101
In the recent past, other oxidation states of manganese received attention in the context of photoluminescence; for instance, many manganese(II) compounds are known to emit in the solid state.74 Furthermore, a bis-(tris(carbene)borate) manganese(IV) complex featured dual luminescence from LMCT (ligand-to-metal charge-transfer) and metal-centered (MC) excited states.76 At present, the photophysics of manganese(I) isocyanide complexes seems underexplored yet, particularly in comparison to isoelectronic iron(II) polypyridines.28,29,31−33,37−44,79,106−116 By contrast, tricarbonyl complexes of manganese(I) represent a fairly well-investigated class of compounds, which can undergo phototriggered loss of CO ligands. Consequently, they belong to the so-called photoCORM family of compounds, i. e., photoactive carbon monoxide-releasing molecules, that are interesting for the light-controlled delivery of CO for therapeutic purposes.75,117−130
Here, we report on the new complex [Mn(CNdippPhOMe2)6]PF6 (CNdippPhOMe2 = 4-(3,5-dimethoxyphenyl)-2,6-diisopropylphenylisocyanide), including detailed structural, vibrational, electrochemical, and photophysical studies.
2. Experimental Section
2.1. Physical Measurements
NMR spectra were recorded in deuterated solvents on a Bruker Avance III NMR spectrometer operating at a 1H frequency of 600 or 400 MHz, a 13C frequency of 151 or 100 MHz, 31P and 19F frequencies of 565 and 243 MHz, respectively, at 298 K. All chemical shifts are reported in δ values in ppm relative to tetramethylsilane (TMS), referred to protons of the residual nondeuterated solvent used or its carbon atoms, respectively (1H: δ(CD2Cl2) = 5.32 ppm, δ(CDCl3) = 7.26 ppm), the solvent signal (13C: δ(CD2Cl2) = 53.84 ppm, δ(CDCl3) = 77.16 ppm), or an external standard (31P: 85% aqueous H3PO4 in sealed capillary; 19F: CFCl3 in CDCl3).131−133 Signals were assigned with the help of DEPT-135 and two-dimensional correlation spectra (1H,1H-COSY, 1H,13C-HSQC, and 1H,13C-HMBC). Signal multiplicities are abbreviated as s (singlet), d (doublet), t (triplet), sept (septett), m (multiplet), and br. (broad signal).
The 55Mn (I = 5/2) NMR experiments were performed on a 14.1 T instrument at a frequency of 148.56 MHz using a broad-band direct observe probe (BBFO). The transmitter offset was chosen as −1400 ppm and a spectral width of 100 ppm (14.8 kHz) was applied. The 90° pulse was 13 μs at a power level of 100 W. We used a 90° excitation mode, an acquisition time of 2.0 s, and a recycling delay of 100 ms. Sixty four scans were acquired and multiplied with an exponential line broadening function (lb = 5 Hz) before Fourier transformation. The 55Mn chemical shifts were referenced to a saturated solution of KMnO4 in D2O as δ = 0.00 ppm.134 The recommended reference concentration of 0.82 mol kg–1 (13% w/w) is more than twice higher than the solubility in water at room temperature and seems incorrect.
Elemental analyses were performed using a vario MICRO cube CHN element analyzer from Elementar. Samples were burned in sealed tin containers using a stream of oxygen.
High-resolution ESI mass spectra were recorded on a Bruker maXis 4G ESI-Q-TOF spectrometer under direct injection conditions with CH3CN as a solvent.
IR spectra were recorded on a Shimadzu IRTracer-100 with QATR 10 spectrometer. Signal intensities are marked as vs (very strong), s (strong), m (medium), w (weak), br (broad), and vw (very weak).
X-ray powder diffraction (XRPD) measurements were performed with a STOE STADI P diffractometer with a microfocused Cu Kα radiation source (λ = 1.542 Å) equipped with a DECTRIS MYTHEN 1K detector.
Cyclic voltammetry was performed using a Versastat4-200 potentiostat from Princeton Applied Research in a glovebox. A glassy carbon disk electrode served as a working electrode, and two silver wires served as counter and pseudo-reference electrodes. Ferrocene was added as an internal reference (E1/2 vs saturated calomel electrode (SCE) in dichloromethane (DCM) = 0.475 V).135 The solvent was dry and deaerated CH2Cl2 with 0.1 M TBAPF6 (tetra-n-butylammonium hexafluorophosphate) as an electrolyte. Potential sweep rates were, unless otherwise stated, 100 mV s–1.
All photophysical measurements were performed, unless otherwise stated, under an Ar atmosphere in either Schlenk or screw cap cuvette with dry and argon-saturated solvents (CH3CN and CH2Cl2).
UV/vis absorption spectroscopy was performed using a Cary 5000 instrument from Varian.
For photostability experiments, a Flame UV/vis spectrometer system from Ocean optics was used with a continuous-wave (cw) laser from Roithner Lasertechnik emitting at 405 nm (optical output up to 375 mA and 526 mW, circular beam dimension with a diameter of 2.5 mm) and a cw-laser emitting at 447 nm (optical output up to 1050 mA and 1.1 W, rectangular beam dimension with a diameter of 2 × 5 mm2). Spectroscopic experiments with that cw-laser as a light source were carried out at 293 K using self-built cuvette holders allowing temperature control and stirring of the solution with a small magnetic stir bar. Photochemical processes involving free radicals and triplet-excited states are usually sensitive to dissolved oxygen. Therefore, oxygen was removed by saturating the solution with argon (for 5 min) before the measurements.
2.2. Single-Crystal X-ray Structure Analysis
[Mn(CNdippPhOMe2)6]PF6 was recrystallized from DCM and cyclohexane at room temperature. Single yellow block-shaped crystals were obtained after a few days. A suitable crystal with dimensions 0.20 × 0.16 × 0.12 mm3 was selected, and the crystal was mounted on a mylar loop in perfluoroether oil on a Cu-Stoe diffractometer. The crystal was kept at a steady T = 150 K during data collection. The structure was solved with the olex2.solve 1.5136 solution program using iterative methods and using Olex2 1.5137 as the graphical interface. The model was refined with ShelXL 2018/3138 using full-matrix least-squares minimization on F2.
Crystallographic drawings were made with the software packages Mercury139,140 and CYLview.141 Selected crystal data and details of the structure determinations are presented in Tables S1 and S2.
CCDC 2165895 ([Mn(CNdippPhOMe2)6]PF6 at 150 K) contains the supplementary crystallographic data for this article. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre via http://www.ccdc.cam.ac.uk/data_request/cif.(142−144)
2.3. Density Functional Theory (DFT) Calculations
All calculations were carried out for the gas phase using the ORCA 4.1.2 package.145−147 Geometry optimization and the calculation of infrared modes were performed with the composite approach PBEh-3c,148 the basis def2-mSVP(C,H,N,O)148 and def2-mTZVP(Mn)149 with the atom pairwise dispersion correction with the Becke–Johnson damping scheme (D3BJ),150,151 the geometrical counterpoise correction gCP,152 the RIJCOSX approximation, the auxiliary basis def2/J,153 fine numerical integration grids (grid5, gridX7, and NoFinalGrid in ORCA 4 nomenclature), and KDIIS+SOSCF. The crystallographic structure of the complex was used as a starting point. The libint2 library was used for the computation of 2-el integrals.154 The calculated frequencies were scaled by a factor of 0.95.148 No imaginary frequencies were obtained. Calculations were performed without symmetry. The Cartesian coordinates from the geometry optimization are in the Supporting Information (SI). The output results were handled with the ChemCraft software package.155
2.4. General Synthesis Procedures and Product Characterizations
The synthesis and characterization of the ligand CNdippPhOMe2 is described in the SI (see Schemes S1 and S2). The complexation was carried out in dry solvents and under a nitrogen atmosphere using standard Schlenk techniques. [Mn(CO)5Br] and KPF6 were purchased and used without further purification. THF was purified by a solvent purification system (SPS) by Innovative Technology and further degassing by at least three cycles of the freeze–pump–thaw method using nitrogen.
Synthesis of [Mn(CNdippPhOMe2)6]PF6
[Mn(CO)5Br] (20.4 mg, 72.7 μmol) and CNdippPhOMe2 (141 mg, 436 μmol) were dissolved in dry and degassed THF (5 mL) and refluxed for 24 h. After cooling to room temperature, a saturated aqueous solution of KPF6 (5–10 mL) was added. Overnight, a bright yellow to orange solid formed. The precipitate was collected by suction filtration and washed with deionized water. The solid was dissolved in DCM, dried over Na2SO4, filtered off, and washed with DCM. The solvent was removed in vacuum, resulting in an orange solid. This solid was recrystallized in DCM and cyclohexane at room temperature. Yellow crystals were obtained after a few days. The crystals were filtered off and dried under air, leading to a yellow powder. Yield: 82 mg (38.3 μmol, 53% relative to [Mn(CO)5Br]).
Elemental analysis calculated for C126H150F6MnN6O12P: C 70.70, H 3.93, N 7.06%, found: C 70.41, H 3.62, N 6.68%.
HR-ESI-MS(+)(CH3CN): m/z (%) = [M – PF6]+ calcd 1995.0720, found 1995.0679 (100), [M – PF6 – L + CH3CN]+ calcd 1712.9106, found 1712.9063 (36), [M – PF6 – L + CO]+ calcd 1699.8790, found 1699.8749 (20.9).
1H NMR (600 MHz, CD2Cl2): δ/ppm = 7.42 (s, 2H, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 6.74 (d, 3JH–H = 2.3 Hz, 2H, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 6.51 (t, 3JH–H = 2.2 Hz, 1H, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 3.86 (s, 6H, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 3.43 (sept, 3JH–H = 6.9 Hz, 2H, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 1.23 (d, 3JH–H = 7.0 Hz, 12H, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn).
13C{1H} NMR (151 MHz, CD2Cl2): δ/ppm = 176.0–182.0 (m, Cq, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 161.7 (Cq, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 145.0 (Cq, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 142.6 (Cq, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 142.1 (Cq, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 125.3 (Cq, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 122.9 (CH, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 106.1 (CH, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 99.5 (CH, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 55.9 (CH3, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 31.0 (CH, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn), 22.6 (CH3, {(Me2CH)2C6H2(C6H3OMe2)(NC)}6Mn).
55Mn NMR (148.5 MHz, CD2Cl2): δ/ppm = −1389.52 (s, Mn).
19F NMR (565 MHz, CD2Cl2): δ/ppm = −73.68 (d, J = 710.2 Hz, PF6–).
31P NMR (243 MHz, CD2Cl2): δ/ppm = −144.54 (sept, J = 710.2 Hz, PF6–).
Ir (ATR, 298 K): υ/cm–1 = 3075 (w, υ[=C–H]), 2964 (m, υ[−C–H]), 2932 (m, υ[−C–H]), 2875 (w, υ[−C–H]), 2846 (w, υ[−C–H]), 2070 (s, υ[−N≡C]), 1592 (s), 1573 (w), 1457 (m), 1434 (m), 1403 (m), 1386 (w), 1355 (m), 1324 (w), 1310 (w), 1266 (m), 1203 (s), 1153 (s), 1133 (w), 1111 (w), 1064 (m), 1038 (m), 992 (w), 962 (w), 943 (w), 928 (w), 904 (w), 891 (w), 877 (w), 836 (vs, ν[PF6]), 788 (s), 757 (m), 735 (m), 697 (m), 660 (w), 586 (vs), 556 (m), 525 (w), 480 (w), 464 (w).
The experimental XRPD pattern as well as the calculated pattern obtained from the single-crystal structure are shown in Figure S7. Not surprisingly, the XRPD pattern of the single-crystal structure containing 0.8 water and 2 cyclohexane molecules per unit cell does not match the XRPD pattern of the yellow powder of [Mn(CNdippPhOMe2)6]PF6 because of the loss of solvent by drying the crystals under air.
3. Results and Discussion
3.1. Synthesis, Characterization, Crystal Structure, and DFT Calculations
The ligand 4-(3,5-dimethoxyphenyl)-2,6-diisopropylphenylisocyanide (CNdippPhOMe2) was synthesized following a previously published procedure,101 which was adapted with modifications inspired by other studies (Schemes S1 and S2).156,157 The complex was prepared by reacting 6 equiv of the monodentate isocyanide ligand CNdippPhOMe2 with [Mn(CO)5Br] in THF under reflux and inert gas for 24 h (Scheme 1). Afterward, the complex was precipitated with an aqueous hexafluorophosphate solution and yellow crystals were obtained from DCM/cyclohexane. As a solid, the compound [Mn(CNdippPhOMe2)6]PF6 can be handled under an ambient atmosphere without noticeable degradation, and 1H, 13C, 19F, and 31P NMR spectra in deuterated DCM confirm that the compound is pure (Figures S1–S5). The ESI-HR measurements in CH3CN exhibit the expected mass peak for [M – PF6],+ as well as the mass peaks for [M – PF6 – L + CH3CN]+ and [M – PF6 – L + CO]+ (Figure S8 and Table S3). Over time, the solution stored under ambient conditions and ambient light becomes intensely yellow colored, and measurements performed after a few days exhibit the mass peaks for [M – PF6 – L + CH3CN]+, [M – PF6 − 2L + 2CH3CN]+, [M – PF6 − 3L + 2O]+, and [M – PF6 − 4L + 2O]+, along with some additional minor mass peaks (Figure S9 and Table S3). Evidently, ligand exchange occurs in coordinating CH3CN solvent over time, and furthermore, the presence of oxygen seems to be detrimental.
Scheme 1. Synthesis of [Mn(CNdippPhOMe2)6]PF6.
In the 13C NMR spectrum of [Mn(CNdippPhOMe2)6]PF6 in CD2Cl2 (Figures S2 and S3), the coordinating C-atoms of the aromatic isocyanide ligands appear between 170 and 185 ppm in the form of a multiplet, likely reflecting an overlay of 55Mn (I = 5/2) and 14N (I = 1) coupling patterns. This finding fits perfectly with the previously reported three compounds [Mn(CNPh)6]BF4, [Mn(CNPhpara–Me)6]BF4, and [Mn(CNPhp–OMe)6]BF4 in CD3CN, which exhibit identical signals in the chemical shift range from 170 to 185 ppm.158
The 55Mn NMR spectrum in CD2Cl2 (Figure S6) exhibits a remarkably sharp resonance at -1390 ppm with a linewidth (fwhh) of only 69 Hz, in line with an almost perfect Oh symmetry of the primary coordination sphere around Mn. The observable chemical shift is in the expected range for Mn(I) between -1000 ppm and -1500 ppm and thus confirms the +I oxidation state. This signal is not changing over a few hours in CD2Cl2 at 298 K in the dark, hence ligand exchange is evidently slower in this solvent than in CH3CN. The chemical shift observed for [Mn(CNdippPhOMe2)6]+ (−1390 ppm) lies between the values found for our two previously investigated manganese(I) complexes with chelating arylisocyanides,53 which featured 55Mn resonances at -1225 and -1419 ppm, suggesting that the new complex has an electron density at Mn(I) somewhere in between the two previously studied related compounds. More striking is the very large difference in linewidth associated with the 55Mn resonances, which is only 69 Hz for [Mn(CNdippPhOMe2)6]+ but amounts to 2100 Hz and 5900 Hz in the previously studied chelate complexes, and this reflects the higher symmetry of the new Mn(I) complex with monodentate arylisocyanide ligands. Furthermore, this finding of a relatively narrow 55Mn resonance bandwidth is similar to the abovementioned other Mn(I) complexes with monodentate arylisocyanide ligands of the type [Mn(CNR)6]BF4. For example, [Mn(CNPh)6]BF4 featured its 55Mn resonance at -1382 ppm with a linewidth of 42 Hz in CD3CN at 18 °C.158
An additional resonance at -1387 ppm is observable in the 55Mn NMR spectrum of [Mn(CNdippPhOMe2)6]PF6 (Figure S6), but its integral is more than 10 times lower than that of the main resonance at -1389 ppm. We tentatively attribute the minor resonance at −1387 ppm to [Mn(CO)(CNdippPhOMe2)5]PF6, the mass peak of which is furthermore observable in the high-resolution electrospray ionization mass spectrometry (HR-ESI-MS) measurements.
The compound [Mn(CNdippPhOMe2)6]PF6 crystallizes in the P1̅ space group with Z = 1 molecule in the unit cell. The obtained crystal structure contains 0.8 water and 2 cyclohexane molecules per formula unit. The asymmetric unit consists of the manganese cation, three CNdippPhOMe2 ligands, and the PF6– anion. The Mn(I) cation is coordinated by six neutral CNdippPhOMe2 ligands (Figure 1). The Mn–C bond distances are between 1.878(8) and 1.881(7) Å at 150 K (Table S2). The slightly distorted octahedral geometry is evident from the small differences in Mn–C bond lengths and the small deviations in the C–Mn–C bond angles from 90° (Table S2). For a more quantitative analysis of the deviation from perfect octahedral symmetry, we employed the Σ and Θ parameters, where Σ is a general measure for the deviation from ideal octahedral geometry and Θ quantifies the distortion from an octahedral toward a trigonal prismatic structure (eq S1 and Table S2).159−161 A perfectly octahedral complex would give Σ = Θ = 0.159−161 From the crystal structure of [Mn(CNdippPhOMe2)6]PF6, we obtain Σ = 22.55° and Θ = 58.37°, relatively close to Oh symmetry. For reference, the [Mn(CNPh)6]I3 compound has Σ = 25.84° and Θ = 65.51° (Table S2).162 Furthermore, the structure of [Mn(CNdippPhOMe2)6]PF6 is also somewhat less distorted than the analogous W(0) complex [W(CNdippPhOMe2)6] (Σ = 26.52° and Θ = 66.33°, Table S2).101
Figure 1.
Crystal structure of [Mn(CNdippPhOMe2)6]PF6. Hydrogen atoms and solvents (H2O and cyclohexane) are omitted for clarity. Color Code: C (gray), O (red), N (blue), Mn (purple), F (green), and P (orange).
The DFT geometry optimization of [Mn(CNdippPhOMe2)6]PF6 (PBEh-3c/def2-mSVP(C,H,N,O)/def2-mTZVP(Mn)) yields similar Mn–C bond lengths as the X-ray crystal structure (Table S2). Furthermore, the DFT-optimized geometry features a similar deviation from ideal octahedral geometry as the X-ray crystal structure (Σ = 20.83° and Θ = 56.41°, Table S2).
[Mn(CNdippPhOMe2)6]PF6 can be isolated as a yellow crystalline powder by filtration and subsequent drying under air. Not surprisingly, the experimental XRPD pattern of that yellow crystalline powder does not match the calculated X-ray pattern based on the single-crystal structure (Figure S7), due to the loss of crystal solvents during filtration and air drying. Combustion analysis of the crystalline yellow powder confirms that it is solvent-free.
3.2. Vibrational Spectroscopy and DFT Analysis
Infrared spectra of the complex and the free ligand were measured in the range of 400–3500 cm–1 (Figure 2). The C≡N stretching frequency (υ(C ≡ N)) of isocyanides is usually strongly influenced by the metal coordination,46,52,53,156,163−178 and this is also the case here. In the IR spectrum of the free ligand, υ(C≡N) is observed at 2116 cm–1, which is nearly the same for the analogous W-complex (2115 cm–1), whereas in the complex [Mn(CNdippPhOMe2)6]PF6, it is shifted to 2070 cm–1, due to π-backbonding from the metal center to the π* antibonding orbital of the isocyanide moiety. Furthermore, the respective IR band is significantly broader in the complex than in the free ligand, presumably due to the somewhat distorted octahedral coordination geometry resulting in slightly different metal–ligand bond lengths and angles (Table S2), and consequently to a continuum of slightly differing vibrational frequencies (see below). Another explanation for the broadening could be the fact that the symmetry of the vibrational band in the complex is T1u, whereas in the free ligand the symmetry is simply A. Two prominent IR bands at 586 and 789 cm–1 appearing in the spectrum of the complex but not in the free ligand are tentatively attributed to Mn(I)-C vibrational modes, in analogy to a recently explored isoelectronic Cr(0) arylisocyanide complex, where the Cr(0)–C vibration appeared at 590 cm–1.66
Figure 2.
Solid-state ATR-IR spectra of the compound [Mn(CNdippPhOMe2)6]PF6 (top) and the free ligand (4) (bottom). The red marked frequencies of the complex are assigned in Table 1.
For an ideal octahedral complex of the type [Mn–(C≡–N–C(Ar))6], six IR-active T1u modes are expected (υ(C≡N), υ(Mn–C), υ(N-C(Ar)), δ(Mn–C≡N), δ(C–Mn–C), and δ(C≡N-C(Ar))).177 The infrared spectrum was analyzed only in terms of [Mn(−C≡N)] core vibrations (Table 1). The core vibrations of [Mn(CNdippPhOMe2)6]+ were obtained from DFT calculations (PBEh-3c/def2-mSVP(C,H,N,O)/def2-mTZVP(Mn)) (Figures S11–S14) and compared with the experimentally obtained frequencies (Table 1 and Figures S12–S14). In addition, we compared the assigned vibrational modes with previous reports.46,52,53,171,177
Table 1. Selected Infrared Vibrational Frequencies for [Mn(CNdippPhOMe2)6]PF6 (Red Marked Frequencies in Figure 2) in Comparison with DFT Calculated Frequencies (PBEh-3c/def2-mSVP(C,H,N,O)/def2-mTZVP(Mn))a Scaled by a Factor of 0.95148,b.
complex | mode | IRexpc (cm–1) | IRDFT (cm–1) |
---|---|---|---|
[Mn(CNdippPhOMe2)6]PF6 | υ(C≡N) T1u | 2070 | 2176 (2973), 2181 (2863), 2184 (1231)d and 2184 (2134)d |
υ(C≡N) Eg | 2184 (1231),d 2184 (2134)d and 2188 (218) | ||
υ(C≡N) A1g | 2261 (1.07) | ||
PF6– | 834 | ||
υ(Mn–C) T1u | 789 | 789.72 (118), 789.77 (121) and 790.38 (118) | |
υ(Mn–C) Eg | 792.46 (0.29) and 792.70 (1.66) | ||
υ(Mn–C) A1g | 793.12 (0.71) | ||
δ(C–Mn–C) and δ(Mn–C≡N) T1u | 586 | 599.57 (99), 601.19 (169), 608.21 (41), 610.13 (168), 613.87 (91) and 618.43 (131) | |
[Mn(Ltri)2]PF653 e | υ(C≡N) T1u | 2081 | |
Mn–C | 584 | ||
[Mn(Lbi)3]PF653 f | υ(C≡N) T1u | 2064 | |
Mn–C | 568 | ||
[Mn(CNPh)6]PF646 | υ(C≡N) T1u | 2088 (vs) | |
PF6– | 840 (m) | ||
[Mn(CNPh)6]I/Br/Cl46 | υ(C≡N) T1u | 2088 (vs) | |
[Mn(CNPh)6]Cl52 | υ(C≡N) T1u | 2080g | |
[Mn(CNPh)6]BF4171 | υ(C≡N) T1u | 2084 | |
[Mn(CNPh)6]I177 | υ(C≡N) T1uh | 2085 (vs) and 1993 (sh) | |
υ(N(isocyanide)-C(pheny)) T1uh | 1210 | ||
δ(Mn–C≡N) T1uh | 600 (vs) | ||
υ(Mn–C) T1uh | 297 and 319 (multiplet)i | ||
δ(C–Mn–C) T1uh | 113 (w) |
The intensities of the vibrational frequencies obtained from DFT calculations (PBEh-3c/def2-mSVP(C,H,N,O)/def2-mTZVP(Mn)) are indicated in parentheses. (Abbreviations: bend., bending; str., stretching; ip, in-plane).
IR data from previously reported comparable compounds are included for comparison.
of the solid state.
One T1u and Eg modes are mixed together at 2184 cm–1 leading to two IR observable modes.
Ltri = 5,5′-(2-isocyano-5-methyl-1,3-phenylene)bis(2-(3,5-di-tert-butyl-2-isocyanophenyl)thiophene).
Lbi = 2,5-bis(3,5-di-tert-butyl-2-isocyanophenyl)thiophene.
The peak shows a shoulder at lower wavenumbers.52
Approximated local symmetry Oh.177
Due to a breakdown of the T1u degeneracy.177
For the complex [Mn(CNdippPhOMe2)6]+ (approximated local symmetry Oh), two symmetrical Raman-active (A1g and doubly degenerate Eg) and one asymmetrical IR-active (triply degenerate T1u) metal–ligand stretching modes (υ(Mn–C)) are expected.177,179,180 DFT vibrational analysis reveals a splitting of the IR-active υ(Mn–C) T1u mode into three stretching modes with observable Mn motion in the x, y, and z directions for [Mn(CNdippPhOMe2)6]+ (Table 1 and Figures S12–S14) at 789.72, 789.77, and 790.38 cm–1, probably because of the small deviation from an ideal octahedral symmetry (Table S2). These modes are forming one band at 790 cm–1 in a simulated IR spectrum (Figure S11).
A complex with Oh symmetry should furthermore exhibit one Raman-active (triply degenerate T2g), one forbidden (triply degenerate T2u), and one IR-active (triply degenerate T1u) metal–ligand bending mode (δ(Mn–C≡N) as well as for δ(C–Mn–C)).177,179,180 The calculations reveal between 599 and 618 cm–1 six-core vibration bending modes in x, y, and z directions with dominant Mn motions (Table 1). Specifically, these modes are at 599.57, 601.19, 608.21, 610.13, 613.87, and 618.43 cm–1. It seems that these modes originate from the combination of the δ(Mn–C≡N) and δ(C–Mn–C) bending modes, related to the fact that the complex shows deviation from ideal octahedral geometry (Table S2). All computed modes form together one band at 610 cm–1 in a simulated IR spectrum (Figure S11) and correspond to the expected T1u bending mode.
Finally, the υ(C≡N) T1u mode is also split into a few IR-active motions in x, y, and z directions. The IR-active modes can be identified in DFT calculations at 2176 and 2181 as well as at 2184 and 2184 cm–1 due to a mixing of one T1u and one Eg motion (Table 1 and see SI, Figures S12–S14). These modes are forming one intense broad band at 2181 cm–1 in a simulated IR spectrum (Figure S11) and correspond also to an expected T1u stretching mode.
On the basis of these DFT results supported by literature reports, the most relevant vibrational modes observed experimentally for [Mn(CNdippPhOMe2)6]PF6 can be assigned (Table 1). The calculated T1u modes are split into three components due to small deviations from the Oh symmetry but these splittings cannot be observed experimentally. The observed band at 2070 cm–1 is attributed to the υ(C≡N) T1u mode, calculated to appear at ca. 2181 cm–1. This assignment is in agreement with several similar manganese(I) isocyanide complexes, which exhibit this mode between 2064 and 2088 cm–1.46,52,53,171,177 The respective IR band exhibits a weak shoulder at lower wavenumbers, which is also reported in the literature, and which results from some deviation from linearity of the CNR axis at the N-atom.52,177 The X-ray crystal structure of [Mn(CNdippPhOMe2)6]PF6 shows that such a distortion is indeed present in the solid state. Alternatively, the respective IR shoulder could stem from a CO vibration of [Mn(CO)(CNdippPhOMe2)5]PF6, for which the 55Mn NMR spectrum in Figure S6 provides some evidence (see above). The prominent broad band at 834 cm–1 is attributable to the PF6– anion, supported also by the fact that for [Mn(CNPh)6]PF6 this frequency is reported at 840 cm–1.46 The υ(Mn–C), δ(Mn–C≡N), and δ(C–Mn–C) T1u modes resulting from the octahedral coordination skeleton of [Mn(CNdippPhOMe2)6]PF6 are observed at 789 and 586 cm–1 and can be assigned to the calculated IR-active T1u modes at ca. 790 (υ(Mn–C) T1u) and ca. 610 (δ(Mn–C≡N) and δ(C–Mn–C) T1u) cm–1. In the two previously investigated manganese(I) complexes with bidentate and tridentate isocyanides ligands, Mn–C vibrations at 568 and 584 cm–1 were found.53 For [Mn(CNPh)6]I, splitting into the two T1u bending modes δ(Mn–C≡N) (600 cm–1) and δ(C–Mn–C) (113 cm–1) was reported.177 As mentioned above, in our calculations, these two modes appear together around 610 cm–1 (see gifs in the SI).
3.3. Electrochemistry
The cyclic voltammogram of [Mn(CNdippPhOMe2)6]PF6 in dry and deaerated CH2Cl2 containing 0.1 M TBAPF6 shows an oxidation wave at 1.21 V vs SCE (E1/2), which can be attributed to the one-electron oxidation of Mn(I) to Mn(II) (Figure 3A). A value of E1/2 = 1.22 V was determined by differential pulse voltammetry (Figure 3B). In the potential range between 0 and −2.3 V vs SCE (Figure S15), no redox waves are detectable, and thus the reduction potentials of the coordinated ligands cannot be determined under these conditions. [Mn(CNdippPhOMe2)6]PF6 undergoes a second one-electron oxidation at E1/2 = 1.74 V vs SCE, likely attributable to the redox couple Mn(II/III) (Figure S15). The shape of the wave in Figure 3A and the peak–current ratio indicate a quasi-reversible chemical and electrochemical process. The difference between the anodic (Epa) and cathodic (Epc) peak potentials (peak-to-peak separation, ΔEp) is 92 mV (at a scan rate of 100 mV s–1) instead of the expectable 57 mV for a one-electron process at 25 °C, and the ratio of the anodic and cathodic peak currents is not equal to one (|Ipa/Ipc| ≠ 1).181
Figure 3.
(A) Cyclic voltammogram for the Mn(I)/Mn(II) redox couple of [Mn(CNdippPhOMe2)6]PF6 in dry and deaerated CH2Cl2 containing 0.1 M TBAPF6 at 20 °C, recorded with a scan rate of 100 mV s–1. (B) Differential pulse voltammetry for the same redox couple under identical conditions.
To investigate the electrochemical reversibility of the single electron transfer process at 1.21 V in more detail, the cyclic voltammogram was measured at different scan rates (Figure S16A), and the respective data were used to make a Randles–Ševcík plot (Figure S16B).181 The observed slight deviation from linearity in the plots of Ipa and Ipc vs υ1/2 (Figure S16B) suggests electrochemical quasi-reversibility, further confirmed by the observation that the peak-to-peak separation gets larger with increasing potential scan rate (see Trumpet plot in Figure S17).
In [Mn(CNPh)6]PF6, the quasi-reversible oxidation of Mn(I) to Mn(II) occurs at 1.01 V vs SCE in CH2Cl2 (Table 2).45,49 The E1/2 value of [Mn(CNPh)6]PF6 is close to the Mn(I/II) potentials reported for our previously investigated [Mn(Ltri)2]PF6 and [Mn(Lbi)3]PF6 compounds (1.00 and 1.05 V vs SCE in CH2Cl2; Ltri = 5,5′-(2-isocyano-5-methyl-1,3-phenylene)bis(2-(3,5-di-tert-butyl-2-isocyanophenyl)thiophene) and Lbi = 2,5-bis(3,5-di-tert-butyl-2-isocyanophenyl)thiophene).53 In [Mn(CNMe)6]PF6, the oxidation of Mn(I) to Mn(II) occurs already at 0.47 V vs SCE in CH2Cl2 due to the stronger electron-donating character of the aliphatic isocyanide ligands compared to arylisocyanides.45,49 For [Mn(CNMe)6]PF6 and [Mn(CNPh)6]PF6, a second one-electron wave was detected at 1.59 and 1.91 V vs SCE, respectively, and assigned to the oxidation of Mn(II) to Mn(III).45 Against this background, it seems clear that the first and the second one-electron oxidation events observed for [Mn(CNdippPhOMe2)6]PF6 are metal-based and furthermore that the arylisocyanide ligand CNdippPhOMe2 makes oxidation easier compared to the other discussed arylisocyanide manganese(I) complexes.
Table 2. Electrochemical Potentials of [Mn(CNdippPhOMe2)6]PF6 in Comparison with Similar Compounds from the Literature.
compound | E1/2 (MnI/II) vs SCE (V) | E1/2 (MnII/III) vs SCE (V) |
---|---|---|
[Mn(CNdippPhOMe2)6]PF6a | 1.21 | 1.74b |
[Mn(Ltri)2]PF653 a,c | 1.00 | |
[Mn(Lbi)3]PF653 a,d | 1.05 | |
[Mn(CNPh)6]PF645 e | 1.01 | 1.91 |
[Mn(CNMe)6]PF645 e | 0.47 | 1.59 |
In (dry and deaerated) CH2Cl2 at 20 °C with TBAPF6 (0.1 M) as the supporting electrolyte.
Only observable during one cycle (irreversible).
Ltri = 5,5′-(2-isocyano-5-methyl-1,3-phenylene)bis(2-(3,5-di-tert-butyl-2-isocyanophenyl)thiophene).
Lbi = 2,5-bis(3,5-di-tert-butyl-2-isocyanophenyl)thiophene.
In CH2Cl2 (5 × 10–3 M) with TBAClO4 (0.1 M) as the supporting electrolyte.
3.4. Optical Spectroscopy
The UV–vis spectrum of [Mn(CNdippPhOMe2)6]PF6 exhibits two absorption band maxima at 278 nm and 370 nm (Figure 4, blue line), similar to [Mn(CNPh)6]Cl.52 Thus, the yellow color stems from the broad absorption band maximizing at 370 nm, which tails into the violet region and which is not present in the UV–vis spectrum of the free ligand (Figure 4B, black trace). In CH2Cl2, the band at 278 nm has an extinction coefficient (ε) of 1.5 × 105 M –1 cm–1, whereas the band at 370 nm shows a slightly lower ε-value of 1.4 × 105M–1 cm–1. For [Mn(CNPh)6]Cl in ethanol/methanol/diethyl ether (8:2:1), an absorption band at 225 nm (ε = 7.5 × 104 M–1 cm–1) with shoulders at 234 (ε = 7.1 × 104 M–1 cm–1) and 249 nm (ε = 5.1 × 104 M–1 cm–1) was assigned to an intraligand (IL) transition (π → π*), whereas a band centered at 322 nm (ε = 6.6 × 104 M–1 cm–1) with a shoulder at 340 nm (ε = 6.1 × 104 M–1 cm–1) was attributed to MLCT (dπ → π*) transitions.99 Our previously reported Mn(I) compounds [Mn(Ltri)2]PF6 and [Mn(Lbi)3]PF6 in CH2Cl2 showed absorption bands near 400 nm with similar extinction coefficients ((2–4) × 104 M–1 cm–1) and were likewise assigned to MLCT transitions.53 Against this background, it seems plausible that the absorption band observable for [Mn(CNdippPhOMe2)6]PF6 at 278 nm is predominantly due to intraligand (IL) transitions (π → π*), whereas the band at 370 nm could be due to a delocalized MLCT or a mixed MLCT/IL transition.
Figure 4.
UV/vis absorption spectra of [Mn(CNdippPhOMe2)6]PF6 in dry and argon-saturated (A) CH2Cl2 and (B) CH3CN before (blue) and after (green) irradiation with the Xenon lamp of (A) a spectrofluorometer or (B) a picosecond laser. The UV/vis spectrum of the free ligand in CH2Cl2 is included in (A). The absorbance axes (colored in green) apply to all measured spectra.
Continuous photoexcitation into the band at 370 nm or into its tail at 410 nm causes a decrease of the absorption band at 370 nm and leads to the appearance of a shoulder near 425 nm, both in CH3CN and in CH2Cl2 (green traces in Figure 4). This behavior is observable under laser irradiation (Figure 4A) as well as under illumination with the Xenon lamp of a commercial fluorimeter (Figure 4B), suggesting that photoinduced ligand dissociation occurs.
In contrast to our previously reported manganese(I)53 and isoelectronic chromium(0)65,66 complexes with chelating arylisocyanide ligands,97 [Mn(CNdippPhOMe2)6]PF6 does not show any detectable photoluminescence, neither in solution at room temperature (both in CH2Cl2 and CH3CN) nor at 77 K in a 2-methyl-THF glass. Consequently, we explored whether there are any dark (i. e., nonluminescent) states that could be detectable prior to the abovementioned photodecomposition. Light-induced ligand dissociation typically occurs from dissociative metal-centered excited states, and those states are usually nonemissive.124,182−184 However, all our attempts to detect any such dark state with ns, ps, and fs laser transient absorption spectroscopy were unsuccessful. These findings are reminiscent of photoCORM behavior, such as observed, for example, in tricarbonyl manganese(I) complexes of the type [Mn(CO)3(Rbpy)Br] (Rbpy = 4,4′-disubstituted-2,2′-bipyridyl ligand, where R = tBu, H or CF3).182 Upon photoexcitation, these complexes undergo loss of one CO ligand on the femtosecond time scale, followed by solvent coordination on the picosecond time scale. The proposed mechanism of CO extrusion involves ultrafast internal conversion from an initially populated MLCT excited state to a dissociative ligand-field (d–d) excited state.182
3.5. Photoinduced Ligand Dissociation
To investigate the photodissociation in more detail, we irradiated [Mn(CNdippPhOMe2)6]PF6 in CH2Cl2 (Figures 5A and S18A) and in CH3CN (Figure 5B) with a blue continuous-wave (cw) laser (405 nm, 526 mW). The absorbance at the excitation wavelength was adjusted close to 0.1 (Table S4), resulting in concentrations of 7.29 μM in CH2Cl2 (7.15 μM for a second measurement) and 7.13 μM in CH3CN. The measurements were done at 25 °C and the absorption spectra were measured each 10 s during continuous irradiation at 405 nm.
Figure 5.
Photodegradation of (A) 7.29 μM [Mn(CNdippPhOMe2)6]PF6 in dry and argon-saturated CH2Cl2 and (B) 7.13 μM of the same compound in dry and argon-saturated CH3CN upon irradiation with a cw-laser (405 nm, 526 mW) for 30 min. Time steps between each spectrum: 10 seconds. The dips at 405 nm in (A) and (B) are artefacts caused by the cw-laser (marked by λexc and purple arrows).
Based on the series of UV–vis spectra in Figure 5, [Mn(CNdippPhOMe2)6]+ undergoes nearly complete photodegradation within 30 min. Aside from the very prominent absorbance decrease at 370 nm, the absorption band shift from 278 nm to 273 nm seems important, because the free ligand has an absorption band maximum at 273 nm (black trace in Figure 4A). Thus, the spectral evolution over time is compatible with the loss of one or more CNdippPhOMe2 ligands. Furthermore, in CH3CN, a new absorption band at 425 nm becomes increasingly prominent with increasing irradiation time. When CH2Cl2 and CH3CN solutions of [Mn(CNdippPhOMe2)6]PF6 are kept under ambient conditions without direct 405 nm irradiation, the UV–vis absorption spectra show comparatively minor changes over much longer time periods (Figure S19), compatible with the view that ligand loss is phototriggered. MS-ESI-HR measurements of a CH3CN sample prior and after photoirradiation (Figure S10 and Table S3) show that the initially dominant [M – PF6]+ mass peak (relative intensity of 85% with respect to 100% for [M – PF6 – L + CH3CN]+; L = CNdippPhOMe2) vanishes with continuing irradiation, finally resulting in relative intensities of 4.3% for [M – PF6]+ and 64% for [M – PF6 – L + CH3CN]+. Furthermore, a new mass peak attributable to [M – PF6 − 2L + 2CH3CN]+ with a relative intensity of 15.7% becomes observable. Collectively, the irradiation experiments and HR-ESI-MS measurements demonstrate that light-induced ligand dissociation from an electronically excited state is much faster than spontaneous ligand dissociation in the electronic ground state.
The prototypical d6 metal compound [Ru(bpy)3]2+ is known to undergo photodegradation as well,4,185−187 and consequently it seemed meaningful to compare the photostabilities of [Mn(CNdippPhOMe2)6]PF6 and [Ru(bpy)3](PF6)2 in CH3CN under similar conditions (SI pages S27 to S31 and Figures S16B and S20 to S22). The key finding is that in CH3CN our manganese(I) complex undergoes photodegradation far more efficiently than [Ru(bpy)3](PF6)2 under these conditions of very intense cw-laser irradiation.
4. Conclusions
The primary coordination sphere of the [Mn(CNdippPhOMe2)6]+ complex comes close to Oh symmetry according to single-crystal X-ray diffraction, 55Mn NMR spectroscopy, IR spectroscopy, and vibrational analysis with DFT calculations. A detailed vibrational analysis permitted the identification of the IR-active and asymmetrical υ(C≡N), υ(Mn–C), δ(C–Mn–C), and δ(Mn–C≡N) vibrational modes (T1u), and furthermore reveals substantial differences in the π-backbonding interaction between the two previously investigated Mn(I) complexes with chelating isocyanide ligands and the new compound studied here (Table 3). Based on cyclic voltammetry, the Mn(I) center of [Mn(CNdippPhOMe2)6]PF6 is substantially more electron-rich than in previously reported hexakis(arylisocyanide)manganese(I) complexes.45−51 Unlike our recently reported manganese(I) complexes with tris(bidendate) and bis(tridentate) coordination environments (Table 3),53,97 and unlike its third-row congener [W(CNdippPhOMe2)6],100−105 the [Mn(CNdippPhOMe2)6]+ complex is nonemissive. Instead, photoirradiation into its lowest energy UV–vis absorption band leads to ligand dissociation, similar to what is well-known from manganese(I) carbonyl compounds.75,117−119,121,123−130 The much weaker ligand field in our 3d6 complex compared to that experienced by W(0) in the same coordination environment likely leads to fast and efficient population of a dissociative metal-centered excited state, which represents a common challenge in first-row transition metal compounds.56,59 The photophysical properties of the new Mn(I) complex differ substantially from those of the recently reported analogous complexes with bi- and tridentate isocyanide ligands (Table 3).53 Aside from the chelating nature of the previously used ligands, their thiophene units in the ligand skeleton likely affect the electronic structures of the resulting Mn(I) complexes.
Table 3. Comparison of [Mn(CNdippPhOMe2)6]PF6 to Related Mn(I) Compounds with Tris(bidendate) and Bis(tridentate) Coordination Environmentsa.
In CH2Cl2 at 20 °C.
Thus, our study suggests that monodentate isocyanide ligands are not well suited for obtaining manganese(I) complexes with emissive MLCT excited states. Chelating arylisocyanides, as used previously for chromium(0),65,66 manganese(I) (Table 3),53,97 and molybdenum(0),156,188,189 seem better suited for this purpose.
Acknowledgments
S.O. acknowledges funding from the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)—460752300 and by the University of Basel. O.S.W. thanks the Swiss National Science Foundation for funding through grant number 200021_178760. The authors thank S. Mittelheisser and J. Zurflüh for CHN measurements and M. Pfeffer for ESI mass spectrometry measurements. Calculations were performed at sciCORE (http://scicore.unibas.ch/) scientific computing center at the University of Basel.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.2c01438.
Synthesis procedures and characterization of the CNdippPhOMe2 ligand; 1H, 13C, 19F, 31P, and 55Mn NMR spectroscopy of [Mn(CNdippPhOMe2)6]PF6; crystallographic data and crystal structure of the complex; XRPD pattern; HR-ESI-MS measurements of the complex; IR spectra and vibrational analysis; and cyclic voltammetry measurements, photostability studies, and UV–vis absorption spectroscopy of the complex, as well as cartesian coordinates of the optimized structure of the complex (PDF)
Full dynamic figures of the corresponding metal–ligand vibrations (ZIP)
Author Contributions
The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.
The authors declare no competing financial interest.
Notes
Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)—460752300. Swiss National Science Foundation—200021_178760.
Supplementary Material
References
- Juris A.; Balzani V.; Barigelletti F.; Campagna S.; Belser P.; von Zelewsky A. Ru(II) Polypyridine Complexes: Photophysics, Photochemistry, Eletrochemistry, and Chemiluminescence. Coord. Chem. Rev. 1988, 84, 85–277. 10.1016/0010-8545(88)80032-8. [DOI] [Google Scholar]
- Hammarström L.; Johansson O. Expanded Bite Angles in Tridentate Ligands. Improving the Photophysical Properties in Bistridentate RuII Polypyridine Complexes. Coord. Chem. Rev. 2010, 254, 2546–2559. 10.1016/j.ccr.2010.01.006. [DOI] [Google Scholar]
- Hu K.; Sampaio R. N.; Schneider J.; Troian-Gautier L.; Meyer G. J. Perspectives on Dye Sensitization of Nanocrystalline Mesoporous Thin Films. J. Am. Chem. Soc. 2020, 142, 16099–16116. 10.1021/jacs.0c04886. [DOI] [PubMed] [Google Scholar]
- Soupart A.; Alary F.; Heully J. L.; Elliott P. I. P.; Dixon I. M. Theoretical Study of the Full Photosolvolysis Mechanism of [Ru(Bpy)3]2+: Providing a General Mechanistic Roadmap for the Photochemistry of [Ru(N∧N)3]2+-Type Complexes toward Both Cis and Trans Photoproducts. Inorg. Chem. 2020, 59, 14679–14695. 10.1021/acs.inorgchem.0c01843. [DOI] [PubMed] [Google Scholar]
- Arias-Rotondo D. The Fruit Fly of Photophysics. Nat. Chem. 2022, 14, 716. 10.1038/s41557-022-00955-8. [DOI] [PubMed] [Google Scholar]
- Amemori S.; Sasaki Y.; Yanai N.; Kimizuka N. Near-Infrared-to-Visible Photon Upconversion Sensitized by a Metal Complex with Spin-Forbidden yet Strong S0-T1 Absorption. J. Am. Chem. Soc. 2016, 138, 8702–8705. 10.1021/jacs.6b04692. [DOI] [PubMed] [Google Scholar]
- Sasaki Y.; Yanai N.; Kimizuka N. Osmium Complex-Chromophore Conjugates with Both Singlet-to-Triplet Absorption and Long Triplet Lifetime through Tuning of the Heavy-Atom Effect. Inorg. Chem. 2021, 61, 5982–5990. 10.1021/acs.inorgchem.1c03129. [DOI] [PubMed] [Google Scholar]
- Smitten K. L.; Scattergood P. A.; Kiker C.; Thomas J. A.; Elliott P. I. P. Triazole-Based Osmium(II) Complexes Displaying Red/near-IR Luminescence: Antimicrobial Activity and Super-Resolution Imaging. Chem. Sci. 2020, 11, 8928–8935. 10.1039/d0sc03563g. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ravetz B. D.; Tay N. E. S.; Joe C. L.; Sezen-Edmonds M.; Schmidt M. A.; Tan Y.; Janey J. M.; Eastgate M. D.; Rovis T. Development of a Platform for Near-Infrared Photoredox Catalysis. ACS Cent. Sci. 2020, 6, 2053–2059. 10.1021/acscentsci.0c00948. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Haruki R.; Sasaki Y.; Masutani K.; Yanai N.; Kimizuka N. Leaping across the Visible Range: Near-Infrared-to-Violet Photon Upconversion Employing a Silyl-Substituted Anthracene. Chem. Commun. 2020, 56, 7017–7020. 10.1039/d0cc02240c. [DOI] [PubMed] [Google Scholar]
- Favale J. M.; Danilov E. O.; Yarnell J. E.; Castellano F. N. Photophysical Processes in Rhenium(I) Diiminetricarbonyl Arylisocyanides Featuring Three Interacting Triplet Excited States. Inorg. Chem. 2019, 58, 8750–8762. 10.1021/acs.inorgchem.9b01155. [DOI] [PubMed] [Google Scholar]
- Larsen C. B.; Wenger O. S. Photophysics and Photoredox Catalysis of a Homoleptic Rhenium(I) Tris(Diisocyanide) Complex. Inorg. Chem. 2018, 57, 2965–2968. 10.1021/acs.inorgchem.7b03258. [DOI] [PubMed] [Google Scholar]
- Chakraborty I.; Jimenez J.; Sameera W. M. C.; Kato M.; Mascharak P. K. Luminescent Re(I) Carbonyl Complexes as Trackable PhotoCORMs for CO Delivery to Cellular Targets. Inorg. Chem. 2017, 56, 2863–2873. 10.1021/acs.inorgchem.6b02999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pierri A. E.; Pallaoro A.; Wu G.; Ford P. C. A Luminescent and Biocompatible PhotoCORM. J. Am. Chem. Soc. 2012, 134, 18197–18200. 10.1021/ja3084434. [DOI] [PubMed] [Google Scholar]
- Xue W. M.; Chan M. C. W.; Su Z. M.; Cheung K. K.; Liu S. T.; Che C. M. Spectroscopic and Excited-State Properties of Luminescent Rhenium(I) N-Heterocyclic Carbene Complexes Containing Aromatic Diimine Ligands. Organometallics 1998, 17, 1622–1630. 10.1021/om9709042. [DOI] [Google Scholar]
- Lo K. K. W. Luminescent Rhenium(I) and Iridium(III) Polypyridine Complexes as Biological Probes, Imaging Reagents, and Photocytotoxic Agents. Acc. Chem. Res. 2015, 48, 2985–2995. 10.1021/acs.accounts.5b00211. [DOI] [PubMed] [Google Scholar]
- Chan K. C.; Tong K. M.; Cheng S. C.; Ng C. O.; Yiu S. M.; Ko C. C. Design of Luminescent Isocyano Rhenium(I) Complexes: Photophysics and Effects of the Ancillary Ligands. Inorg. Chem. 2018, 57, 13963–13972. 10.1021/acs.inorgchem.8b02536. [DOI] [PubMed] [Google Scholar]
- Stout M. J.; Skelton B. W.; Sobolev A. N.; Raiteri P.; Massi M.; Simpson P. V. Synthesis and Photochemical Properties of Re(I) Tricarbonyl Complexes Bound to Thione and Thiazol-2-Ylidene Ligands. Organometallics 2020, 39, 3202–3211. 10.1021/acs.organomet.0c00381. [DOI] [Google Scholar]
- Nicholls T. P.; Burt L. K.; Simpson P. V.; Massi M.; Bissember A. C. Tricarbonyl Rhenium(I) Tetrazolato and N-Heterocyclic Carbene Complexes: Versatile Visible-Light-Mediated Photoredox Catalysts. Dalton Trans. 2019, 48, 12749–12754. 10.1039/c9dt02533b. [DOI] [PubMed] [Google Scholar]
- Schmid L.; Glaser F.; Schaer R.; Wenger O. S. High Triplet Energy Iridium(III) Isocyanoborato Complex for Photochemical Upconversion, Photoredox and Energy Transfer Catalysis. J. Am. Chem. Soc. 2022, 144, 963–976. 10.1021/jacs.1c11667. [DOI] [PubMed] [Google Scholar]
- Diluzio S.; Mdluli V.; Connell T. U.; Lewis J.; Vanbenschoten V.; Bernhard S. High-Throughput Screening and Automated Data-Driven Analysis of the Triplet Photophysical Properties of Structurally Diverse, Heteroleptic Iridium(III) Complexes. J. Am. Chem. Soc. 2021, 143, 1179–1194. 10.1021/jacs.0c12290. [DOI] [PubMed] [Google Scholar]
- Coppo P.; Plummer E. A.; De Cola L. Tuning Iridium(III) Phenylpyridine Complexes in the “Almost Blue” Region. Chem. Commun. 2004, 4, 1774–1775. 10.1039/b406851c. [DOI] [PubMed] [Google Scholar]
- Cañada L. M.; Kölling J.; Wen Z.; Wu J. I.; Teets T. S. Cyano-Isocyanide Iridium(III) Complexes with Pure Blue Phosphorescence. Inorg. Chem. 2021, 60, 6391–6402. 10.1021/acs.inorgchem.1c00103. [DOI] [PubMed] [Google Scholar]
- Pal A. K.; Krotkus S.; Fontani M.; Mackenzie C. F. R.; Cordes D. B.; Slawin A. M. Z.; Samuel I. D. W.; Zysman-Colman E. High-Efficiency Deep-Blue-Emitting Organic Light-Emitting Diodes Based on Iridium(III) Carbene Complexes. Adv. Mater. 2018, 30, 1–10. 10.1002/adma.201804231. [DOI] [PubMed] [Google Scholar]
- Bevernaegie R.; Wehlin S. A. M.; Piechota E. J.; Abraham M.; Philouze C.; Meyer G. J.; Elias B.; Troian-Gautier L. Improved Visible Light Absorption of Potent Iridium(III) Photo-Oxidants for Excited-State Electron Transfer Chemistry. J. Am. Chem. Soc. 2020, 142, 2732–2737. 10.1021/jacs.9b12108. [DOI] [PubMed] [Google Scholar]
- Shon J. H.; Kim D.; Rathnayake M. D.; Sittel S.; Weaver J.; Teets T. S. Photoredox Catalysis on Unactivated Substrates with Strongly Reducing Iridium Photosensitizers. Chem. Sci. 2021, 12, 4069–4078. 10.1039/d0sc06306a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Topics in Current Chemistry. In Photochemistry and Photophysics of Coordination Compounds I; Balzani V.; Campagna S., Eds.; Springer: Berlin Heidelberg: Berlin, Heidelberg, 2007; Vol. 280. [Google Scholar]
- Topics in Current Chemistry. In Photochemistry and Photophysics of Coordination Compounds II; Balzani V.; Campagna S., Eds.; Springer: Berlin Heidelberg: Berlin, Heidelberg, 2007; Vol. 281. [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]
- Chábera P.; Kjaer K. S.; Prakash O.; Honarfar A.; Liu Y.; Fredin L. A.; Harlang T. C. B.; Lidin S.; Uhlig J.; Sundström V.; Lomoth R.; Persson P.; Wärnmark K. Fe II 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]
- Mengel A. K. C. 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]
- Liu Y.; 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]
- Liu Y.; Harlang T.; Canton S. E.; Chábera P.; Suárez-Alcántara K.; Fleckhaus A.; Vithanage D. A.; Göransson 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. 10.1039/c3cc43833c. [DOI] [PubMed] [Google Scholar]
- Zimmer P.; Burkhardt L.; Friedrich A.; Steube J.; Neuba A.; Schepper R.; Müller P.; Flörke 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]
- 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]
- 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]
- Vöhringer P. Vibrations Tell the Tale. A Time-Resolved Mid-Infrared Perspective of the Photochemistry of Iron Complexes. Dalton Trans. 2020, 49, 256–266. 10.1039/c9dt04165f. [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]
- 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]
- Paulus B. C.; Adelman S. L.; Jamula L. L. L.; McCusker J. K. 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]
- 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]
- 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]
- Young E. R.; Oldacre A. Iron Hits the Mark. Science. 2019, 363, 225–226. 10.1126/science.aav9866. [DOI] [PubMed] [Google Scholar]
- Juban E. A.; Smeigh A. L.; Monat J. E.; McCusker J. K. Ultrafast Dynamics of Ligand-Field Excited States. Coord. Chem. Rev. 2006, 250, 1783–1791. 10.1016/j.ccr.2006.02.010. [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]
- Treichel P. M.; Dirreen G. E. Electrochemical Oxidation of Isocyanide Complexes of Manganese and Chromium. J. Organomet. Chem. 1972, 39, C20–C22. 10.1016/S0022-328X(00)88893-2. [DOI] [Google Scholar]
- Treichel P. M.; Dirreen G. E.; Mueh H. J. Manganese(I) Isocyanide Complexes. J. Organomet. Chem. 1972, 44, 339–352. 10.1016/S0022-328X(00)82923-X. [DOI] [Google Scholar]
- Treichel P. M. Transition Metal-Lsocyanide Complexes. Adv. Organomet. Chem. 1973, 21–86. 10.1016/S0065-3055(08)60159-8. [DOI] [Google Scholar]
- Treichel P. M.; Mueh H. J. Synthesis and Electrochemistry of C5H5Mn(CNR)3 Compounds. Inorg. Chim. Acta 1977, 22, 265–268. 10.1016/S0020-1693(00)90929-0. [DOI] [Google Scholar]
- Treichel P. M.; Mueh H. J. Electrochemical Studies on [Mn(CNR)6]+ Complexes. Inorg. Chem. 1977, 16, 1167–1169. 10.1021/ic50171a038. [DOI] [Google Scholar]
- Treichel P. M.; Firsich D. W.; Essenmacher G. P. Manganese(I) and Chromium(0) Complexes of Phenyl Isocyanide. Inorg. Chem. 1979, 18, 2405–2409. 10.1021/ic50199a015. [DOI] [Google Scholar]
- Treichel P. M.Manganese. In Comprehensive Organometallic Chemistry; Elsevier, 1982; Vol. 4, pp 1–159. [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 Cr(0), Mo(0), W(0), Mn(I), and Mn(II). Inorg. Chim. Acta 1976, 16, 97–101. 10.1016/S0020-1693(00)91697-9. [DOI] [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]
- 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]
- Melcher F.; Wilken H. Die Verfügbarkeit von Hochtechnologie-Rohstoffen. Chem. Unserer Zeit 2013, 47, 32–49. 10.1002/ciuz.201300597. [DOI] [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]
- 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]
- Wenger O. S. Photoactive Complexes with Earth-Abundant Metals. J. Am. Chem. Soc. 2018, 140, 13522–13533. 10.1021/jacs.8b08822. [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]
- Zobel J. P.; González L. The Quest to Simulate Excited-State Dynamics of Transition Metal Complexes. JACS Au 2021, 1, 1116–1140. 10.1021/jacsau.1c00252. [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. A.; Bings N. H.; 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]
- 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]
- Fataftah M. S.; Bayliss S. L.; Laorenza D. W.; Wang X.; Phelan B. T.; Wilson C. B.; Mintun P. J.; Kovos B. D.; Wasielewski M. R.; Han S.; Sherwin M. S.; Awschalom D. D.; Freedman D. E. Trigonal Bipyramidal V3+ Complex as an Optically Addressable Molecular Qubit Candidate. J. Am. Chem. Soc. 2020, 142, 20400–20408. 10.1021/jacs.0c08986. [DOI] [PubMed] [Google Scholar]
- Zobel J. P.; Knoll T.; González L. Ultrafast and Long-Time Excited State Kinetics of an NIR-Emissive Vanadium(III) Complex II. Elucidating Triplet-to-Singlet Excited-State Dynamics. Chem. Sci. 2021, 12, 10791–10801. 10.1039/d1sc02149d. [DOI] [PMC free article] [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]
- 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]
- 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]
- Wang C.; Otto S.; Dorn M.; Kreidt E.; Lebon J.; Sršan 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]
- Jiménez J.-R.; Doistau B.; Cruz C. M.; Besnard C.; Cuerva J. M.; Campaña 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]
- Jiménez J.; 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]
- Conti C.; Castelli F.; Forster L. S. Photophysics of Hexakis(Cyano)Chromate(3-) and Hexakis(Cyano)Cobaltate(3-) in Polyalcohol-Water Solutions at Room Temperature. J. Phys. Chem. A 1979, 83, 2371–2376. 10.1021/j100481a013. [DOI] [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]
- Farran R.; Le-Quang L.; Mouesca J. M.; Maurel V.; Jouvenot D.; Loiseau F.; Deronzier A.; Chauvin J. [Cr(ttpy)2]3+ as a Multi-Electron Reservoir for Photoinduced Charge Accumulation. Dalton Trans. 2019, 48, 6800–6811. 10.1039/c9dt00848a. [DOI] [PubMed] [Google Scholar]
- Qin Y.; She P.; Huang X.; Huang W.; Zhao Q. Luminescent Manganese(II) Complexes: Synthesis, Properties and Optoelectronic Applications. Coord. Chem. Rev. 2020, 416, 213331 10.1016/j.ccr.2020.213331. [DOI] [Google Scholar]
- Jimenez J.; Chakraborty I.; Dominguez A.; Martinez-Gonzalez J.; Sameera W. M. C.; Mascharak P. K. A Luminescent Manganese PhotoCORM for CO Delivery to Cellular Targets under the Control of Visible Light. Inorg. Chem. 2018, 57, 1766–1773. 10.1021/acs.inorgchem.7b02480. [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]
- Chábera P.; Liu Y.; Prakash O.; Thyrhaug E.; El Nahhas A.; Honarfar A.; Essén S.; Fredin L. A.; Harlang T. C. B.; Kjær K. S.; Handrup K.; Ericson F.; Tatsuno H.; 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 Fe(III) 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. 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.; Wärnmark K. 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]
- Lindh L.; Chábera 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]
- Pal A. K.; Li C.; Hanan G. S.; Zysman-Colman E. Blue-Emissive Cobalt(III) Complexes and Their Use in the Photocatalytic Trifluoromethylation of Polycyclic Aromatic Hydrocarbons. Angew. Chem. 2018, 130, 8159–8163. 10.1002/ange.201802532. [DOI] [PubMed] [Google Scholar]
- Kaufhold S.; Rosemann N. W.; Chábera P.; Lindh L.; Bolaño Losada I.; 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]
- 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]
- Sinha N.; Pfund B.; Wegeberg C.; Prescimone A.; Wenger O. S. Cobalt Carbene(III) 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]
- 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]
- 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]
- Cope J. D.; Denny J. A.; Lamb R. W.; McNamara L. E.; Hammer N. I.; Webster C. E.; Hollis T. K. Synthesis, Characterization, Photophysics, and a Ligand Rearrangement of CCC-NHC Pincer Nickel Complexes: Colors, Polymorphs, Emission, and Raman Spectra. J. Organomet. Chem. 2017, 845, 258–265. 10.1016/j.jorganchem.2017.05.046. [DOI] [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 Cu(I) Emitters:>99% Quantum Efficiency and Microsecond Lifetime. Science 2019, 363, 601–606. 10.1126/science.aav2865. [DOI] [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]
- Shi S.; Jung M. C.; Coburn C.; Tadle A.; Sylvinson D. M. R.; 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]
- Nicholls T. P.; Bissember A. C. Developments in Visible-Light-Mediated Copper Photocatalysis. Tetrahedron Lett. 2019, 60, 150883 10.1016/j.tetlet.2019.06.042. [DOI] [Google Scholar]
- Rentschler M.; Schmid M. A.; Frey W.; Tschierlei S.; Karnahl M. Multidentate Phenanthroline Ligands Containing Additional Donor Moieties and Their Resulting Cu(I) and Ru(II) Photosensitizers: A Comparative Study. Inorg. Chem. 2020, 59, 14762–14771. 10.1021/acs.inorgchem.9b03687. [DOI] [PubMed] [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]
- 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]
- Hölzel T.; Belyaev A.; Terzi M.; Stenzel L.; Gernert M.; Marian C. M.; Steffen A.; Ganter C. Linear Carbene Pyridine Copper Complexes with Sterically Demanding N, N′-Bis(Trityl)Imidazolylidene: Syntheses, Molecular Structures, and Photophysical Properties. Inorg. Chem. 2021, 60, 18529–18543. 10.1021/acs.inorgchem.1c03082. [DOI] [PubMed] [Google Scholar]
- Schinabeck A.; Chen J.; Kang L.; Teng T.; Homeier H. H. H.; Suleymanova A. F.; Shafikov M. Z.; Yu R.; Lu C. Z.; Yersin H. Symmetry-Based Design Strategy for Unprecedentedly Fast Decaying Thermally Activated Delayed Fluorescence (TADF). Application to Dinuclear Cu(I) Compounds. Chem. Mater. 2019, 31, 4392–4404. 10.1021/acs.chemmater.9b00671. [DOI] [Google Scholar]
- Hofbeck T.; Monkowius U.; Yersin H. Highly Efficient Luminescence of Cu(I) Compounds: Thermally Activated Delayed Fluorescence Combined with Short-Lived Phosphorescence. J. Am. Chem. Soc. 2015, 137, 399–404. 10.1021/ja5109672. [DOI] [PubMed] [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]
- 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]
- 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]
- 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]
- 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]
- Kvapilová H.; Sattler W.; Sattler A.; Sazanovich I. V.; Clark I. P.; Towrie M.; Gray H. B.; Záliš S.; Vlček A. Electronic Excited States of Tungsten(0) Arylisocyanides. Inorg. Chem. 2015, 54, 8518–8528. 10.1021/acs.inorgchem.5b01203. [DOI] [PubMed] [Google Scholar]
- Takematsu K.; Wehlin S. A. M.; Sattler W.; Winkler J. R.; Gray H. B. Two-Photon Spectroscopy of Tungsten(0) Arylisocyanides Using Nanosecond-Pulsed Excitation. Dalton Trans. 2017, 46, 13188–13193. 10.1039/C7DT02632C. [DOI] [PubMed] [Google Scholar]
- Fajardo J.; Schwan J.; Kramer W. W.; Takase M. K.; Winkler J. R.; Gray H. B. Third-Generation W(CNAr)6 Photoreductants (CNAr = Fused-Ring and Alkynyl-Bridged Arylisocyanides). Inorg. Chem. 2021, 60, 3481–3491. 10.1021/acs.inorgchem.0c02912. [DOI] [PubMed] [Google Scholar]
- Fajardo J.; 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]
- 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]
- Cannizzo A.; Milne C. J.; Consani C.; Gawelda W.; Bressler C.; van Mourik F.; Chergui M. Light-Induced Spin Crossover in Fe(II)-Based Complexes: The Full Photocycle Unraveled by Ultrafast Optical and X-Ray Spectroscopies. Coord. Chem. Rev. 2010, 254, 2677–2686. 10.1016/j.ccr.2009.12.007. [DOI] [Google Scholar]
- Zhang W.; Alonso-Mori R.; Bergmann U.; Bressler C.; Chollet M.; Galler A.; Gawelda W.; Hadt R. G.; Hartsock R. W.; Kroll T.; Kjær K. S.; Kubiek K.; Lemke H. T.; Liang H. W.; Meyer D. A.; Nielsen M. M.; Purser C.; Robinson J. S.; Solomon E. I.; Sun Z.; Sokaras D.; Van Driel T. B.; Vankó G.; Weng T. C.; Zhu D.; Gaffney K. J. Tracking Excited-State Charge and Spin Dynamics in Iron Coordination Complexes. Nature 2014, 509, 345–348. 10.1038/nature13252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Harlang T. C. B.; Liu Y.; Gordivska O.; Fredin L. A.; Ponseca C. S.; Huang P.; Chábera 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]
- Duchanois T.; Etienne T.; Cebrián 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]
- Zhang W.; Kjær K. S.; Alonso-Mori R.; Bergmann U.; Chollet M.; Fredin L. A.; Hadt R. G.; Hartsock R. W.; Harlang T.; Kroll T.; Kubiček K.; Lemke H. T.; Liang H. W.; Liu Y.; Nielsen M. M.; Persson P.; Robinson J. S.; Solomon E. I.; Sun Z.; Sokaras D.; van Driel T. B.; Weng T.-C.; Zhu D.; 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]
- Chen J.; Browne W. R. Photochemistry of Iron Complexes. Coord. Chem. Rev. 2018, 374, 15–35. 10.1016/j.ccr.2018.06.008. [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]
- Francés-Monerris A.; Gros P. C.; Assfeld X.; Monari A.; Pastore M. Toward Luminescent Iron Complexes: Unravelling the Photophysics by Computing Potential Energy Surfaces. ChemPhotoChem 2019, 3, 666–683. 10.1002/cptc.201900100. [DOI] [Google Scholar]
- Benaicha B.; Van Do K.; Yangui A.; Pittala N.; Lusson A.; Sy M.; Bouchez G.; Fourati H.; Gómez-Garciá C. J.; Triki S.; Boukheddaden K. Interplay between Spin-Crossover and Luminescence in a Multifunctional Single Crystal Iron(II) Complex: Towards a New Generation of Molecular Sensors. Chem. Sci. 2019, 10, 6791–6798. 10.1039/c9sc02331c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kjær K. S.; Van Driel T. B.; Harlang T. C. B.; Kunnus K.; Biasin E.; Ledbetter K.; Hartsock R. W.; Reinhard M. E.; Koroidov S.; Li L.; Laursen M. G.; Hansen F. B.; Vester P.; Christensen M.; Haldrup K.; Nielsen M. M.; Dohn A. O.; Pápai M. I.; Møller K. B.; Chabera P.; Liu Y.; Tatsuno H.; Timm C.; Jarenmark M.; Uhlig J.; Sundstöm V.; Wärnmark K.; Persson P.; Németh Z.; Szemes D. S.; Bajnóczi É.; Vankó G.; Alonso-Mori R.; Glownia J. M.; Nelson S.; Sikorski M.; Sokaras D.; Canton S. E.; Lemke H. T.; Gaffney K. J. Finding Intersections between Electronic Excited State Potential Energy Surfaces with Simultaneous Ultrafast X-Ray Scattering and Spectroscopy. Chem. Sci. 2019, 10, 5749–5760. 10.1039/c8sc04023k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kottelat E.; Zobi F. Visible Light-Activated PhotoCORMs. Inorganics 2017, 5, 24 10.3390/inorganics5020024. [DOI] [Google Scholar]
- Ji X.; Damera K.; Zheng Y.; Yu B.; Otterbein L. E.; Wang B. Toward Carbon Monoxide-Based Therapeutics: Critical Drug Delivery and Developability Issues. J. Pharm. Sci. 2016, 105, 406–416. 10.1016/j.xphs.2015.10.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kottelat E.; Ruggi A.; Zobi F. Red-Light Activated PhotoCORMs of Mn(I) Species Bearing Electron Deficient 2,2′-Azopyridines. Dalton Trans. 2016, 45, 6920–6927. 10.1039/c6dt00858e. [DOI] [PubMed] [Google Scholar]
- Stout M. J.; Stefan A.; Skelton B. W.; Sobolev A. N.; Massi M.; Hochkoeppler A.; Stagni S.; Simpson P. V. Synthesis and Photochemical Properties of Manganese(I) Tricarbonyl Diimine Complexes Bound to Tetrazolato Ligands. Eur. J. Inorg. Chem. 2020, 2020, 292–298. 10.1002/ejic.201900987. [DOI] [Google Scholar]
- Tang M.; Cameron L.; Poland E. M.; Yu L.-J.; Moggach S. A.; Fuller R. O.; Huang H.; Sun J.; Thickett S. C.; Massi M.; Coote M. L.; Ho C. C.; Bissember A. C. Photoactive Metal Carbonyl Complexes Bearing N-Heterocyclic Carbene Ligands: Synthesis, Characterization, and Viability as Photoredox Catalysts. Inorg. Chem. 2022, 61, 1888–1898. 10.1021/acs.inorgchem.1c02964. [DOI] [PubMed] [Google Scholar]
- Cole-Filipiak N. C.; Troß J.; Schrader P.; McCaslin L. M.; Ramasesha K. Ultrafast Infrared Transient Absorption Spectroscopy of Gas-Phase Ni(CO)4 Photodissociation at 261 Nm. J. Chem. Phys. 2022, 156, 144306 10.1063/5.0080844. [DOI] [PubMed] [Google Scholar]
- Marhenke J.; Trevino K.; Works C. The Chemistry, Biology and Design of Photochemical CO Releasing Molecules and the Efforts to Detect CO for Biological Applications. Coord. Chem. Rev. 2016, 306, 533–543. 10.1016/j.ccr.2015.02.017. [DOI] [Google Scholar]
- Schatzschneider U. PhotoCORMs: Light-Triggered Release of Carbon Monoxide from the Coordination Sphere of Transition Metal Complexes for Biological Applications. Inorg. Chim. Acta 2011, 374, 19–23. 10.1016/j.ica.2011.02.068. [DOI] [Google Scholar]
- Chakraborty I.; Carrington S. J.; Mascharak P. K. Design Strategies to Improve the Sensitivity of Photoactive Metal Carbonyl Complexes (PhotoCORMs) to Visible Light and Their Potential as CO-Donors to Biological Targets. Acc. Chem. Res. 2014, 47, 2603–2611. 10.1021/ar500172f. [DOI] [PubMed] [Google Scholar]
- Rimmer R. D.; Pierri A. E.; Ford P. C. Photochemically Activated Carbon Monoxide Release for Biological Targets. Toward Developing Air-Stable PhotoCORMs Labilized by Visible Light. Coord. Chem. Rev. 2012, 256, 1509–1519. 10.1016/j.ccr.2011.12.009. [DOI] [Google Scholar]
- Rudolf P.; Kanal F.; Knorr J.; Nagel C.; Niesel J.; Brixner T.; Schatzschneider U.; Nuernberger P. Ultrafast Photochemistry of a Manganese-Tricarbonyl CO-Releasing Molecule (CORM) in Aqueous Solution. J. Phys. Chem. Lett. 2013, 4, 596–602. 10.1021/jz302061q. [DOI] [PubMed] [Google Scholar]
- Heinemann S. H.; Hoshi T.; Westerhausen M.; Schiller A. Carbon Monoxide – Physiology, Detection and Controlled Release. Chem. Commun. 2014, 50, 3644–3660. 10.1039/c3cc49196j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schatzschneider U. Novel Lead Structures and Activation Mechanisms for CO-Releasing Molecules (CORMs). Br. J. Pharmacol. 2015, 172, 1638–1650. 10.1111/bph.12688. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wright M. A.; Wright J. A. PhotoCORMs: CO Release Moves into the Visible. Dalton Trans. 2016, 45, 6801–6811. 10.1039/c5dt04849d. [DOI] [PubMed] [Google Scholar]
- Gottlieb H. E.; Kotlyar V.; Nudelman A. NMR Chemical Shifts of Common Laboratory Solvents as Trace Impurities. J. Org. Chem. 1997, 3263, 7512–7515. [DOI] [PubMed] [Google Scholar]
- Fulmer G. R.; Miller A. J. M.; Sherden N. H.; Gottlieb H. E.; Nudelman A.; Stoltz B. M.; Bercaw J. E.; Goldberg K. I. NMR Chemical Shifts of Trace Impurities: Common Laboratory Solvents, Organics, and Gases in Deuterated Solvents Relevant to the Organometallic Chemist. Organometallics 2010, 29, 2176–2179. 10.1021/om100106e. [DOI] [Google Scholar]
- Babij N. R.; McCusker E. O.; Whiteker G. T.; Canturk B.; Choy N.; Creemer L. C.; Amicis C. V. D.; Hewlett N. M.; Johnson P. L.; Knobelsdorf J. A.; Li F.; Lorsbach B. A.; Nugent B. M.; Ryan S. J.; Smith M. R.; Yang Q. NMR Chemical Shifts of Trace Impurities: Industrially Preferred Solvents Used in Process and Green Chemistry. Org. Process Res. Dev. 2016, 20, 661–667. 10.1021/acs.oprd.5b00417. [DOI] [Google Scholar]
- Harris R. K.; Becker E. D.; Cabral de Menezes S. M.; Goodfellow R.; Granger P. NMR Nomenclature. Nuclear Spin Properties and Conventions for Chemical Shifts (IUPAC Recommendations 2001). Pure Appl. Chem. 2001, 73, 1795–1818. 10.1351/pac200173111795. [DOI] [PubMed] [Google Scholar]
- Aranzaes J. R.; Daniel M.-C.; Astruc D. Metallocenes as References for the Determination of Redox Potentials by Cyclic Voltammetry - Permethylated Iron and Cobalt Sandwich Complexes, Inhibition by Polyamine Dendrimers, and the Role of Hydroxy-Containing Ferrocenes. Can. J. Chem. 2006, 84, 288–299. 10.1139/v05-262. [DOI] [Google Scholar]
- Bourhis L. J.; Dolomanov O. V.; Gildea R. J.; Howard J. A. K.; Puschmann H. The Anatomy of a Comprehensive Constrained, Restrained Refinement Program for the Modern Computing Environment - Olex2 Dissected. Acta Crystallogr., Sect. A: Found. Crystallogr. 2015, 71, 59–75. 10.1107/S2053273314022207. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dolomanov O. V.; Bourhis L. J.; Gildea R. J.; Howard J. A. K.; Puschmann H. OLEX2: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Crystallogr. 2009, 42, 339–341. 10.1107/S0021889808042726. [DOI] [Google Scholar]
- Sheldrick G. M. Crystal Structure Refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3–8. 10.1107/S2053229614024218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Macrae C. F.; Edgington P. R.; McCabe P.; Pidcock E.; Shields G. P.; Taylor R.; Towler M.; van de Streek J. Mercury: Visualization and Analysis of Crystal Structures. J. Appl. Crystallogr. 2006, 39, 453–457. 10.1107/S002188980600731X. [DOI] [Google Scholar]
- Macrae C. F.; Bruno I. J.; Chisholm J. A.; Edgington P. R.; McCabe P.; Pidcock E.; Rodriguez-Monge L.; Taylor R.; van de Streek J.; Wood P. A. Mercury CSD 2.0 - New Features for the Visualization and Investigation of Crystal Structures. J. Appl. Crystallogr. 2008, 41, 466–470. 10.1107/S0021889807067908. [DOI] [Google Scholar]
- CYL; Legault C. Y. Université de Sherbrooke, 2020. (https://www.cylview.org).
- Allen F. H. The Cambridge Structural Database: A Quarter of a Million Crystal Structures and Rising. Acta Crystallogr., Sect. B: Struct. Sci. 2002, 58, 380–388. 10.1107/S0108768102003890. [DOI] [PubMed] [Google Scholar]
- Groom C. R.; Allen F. H. The Cambridge Structural Database in Retrospect and Prospect. Angew. Chem., Int. Ed. 2014, 53, 662–671. 10.1002/anie.201306438. [DOI] [PubMed] [Google Scholar]
- Groom C. R.; Bruno I. J.; Lightfoot M. P.; Ward S. C. The Cambridge Structural Database. Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater. 2016, 72, 171–179. 10.1107/S2052520616003954. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Neese F. The ORCA Program System. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012, 2, 73–78. 10.1002/wcms.81. [DOI] [Google Scholar]
- Neese F. Software Update: The ORCA Program System, Version 4.0. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2018, 8, e1327 10.1002/wcms.1327. [DOI] [Google Scholar]
- Neese F.; Wennmohs F.; Becker U.; Riplinger C. The ORCA Quantum Chemistry Program Package. J. Chem. Phys. 2020, 152, 224108 10.1063/5.0004608. [DOI] [PubMed] [Google Scholar]
- Grimme S.; Brandenburg J. G.; Bannwarth C.; Hansen A. Consistent Structures and Interactions by Density Functional Theory with Small Atomic Orbital Basis Sets. J. Chem. Phys. 2015, 143, 054107 10.1063/1.4927476. [DOI] [PubMed] [Google Scholar]
- Brandenburg J. G.; Bannwarth C.; Hansen A.; Grimme S. B97-3c: A Revised Low-Cost Variant of the B97-D Density Functional Method. J. Chem. Phys. 2018, 148, 064104 10.1063/1.5012601. [DOI] [PubMed] [Google Scholar]
- Grimme S.; Antony J.; Ehrlich S.; Krieg H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
- Grimme S.; Ehrlich S.; Goerigk L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456–1465. 10.1002/jcc.21759. [DOI] [PubMed] [Google Scholar]
- Kruse H.; Grimme S. A Geometrical Correction for the Inter- and Intra-Molecular Basis Set Superposition Error in Hartree-Fock and Density Functional Theory Calculations for Large Systems. J. Chem. Phys. 2012, 136, 154101 10.1063/1.3700154. [DOI] [PubMed] [Google Scholar]
- Weigend F. Accurate Coulomb-Fitting Basis Sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. 10.1039/B515623H. [DOI] [PubMed] [Google Scholar]
- Libint; Valeev E. F.. http://libint.valeyev.net.
- Chemcraft - Graphical Software for Visualization of Quantum Chemistry Computations.
- 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]
- Patil P.; Ahmadian-Moghaddam M.; Dömling A. Isocyanide 2.0. Green Chem. 2020, 22, 6902–6911. 10.1039/D0GC02722G. [DOI] [Google Scholar]
- Nielson R. M.; Wherland S. Multinuclear NMR Studies of Mn(CNR)6+,2+: R = Methyl, Ethyl, Isopropyl, Tert-Butyl, Cyclohexyl, Benzyl, Phenyl, p-Tolyl, and p-Anisyl. Inorg. Chem. 1985, 24, 3458–3464. 10.1021/ic00215a031. [DOI] [Google Scholar]
- Iasco O.; Boillot M.-L.; Bellec A.; Guillot R.; Rivière E.; Mazerat S.; Nowak S.; Morineau D.; Brosseau A.; Miserque F.; Repain V.; Mallah T. The Disentangling of Hysteretic Spin Transition, Polymorphism and Metastability in Bistable Thin Films Formed by Sublimation of Bis(Scorpionate) Fe(II) Molecules. J. Mater. Chem. C 2017, 5, 11067–11075. 10.1039/C7TC03276E. [DOI] [Google Scholar]
- Guionneau P.; Marchivie M.; Bravic G.; Létard J.-F.; Chasseau D.. Structural Aspects of Spin Crossover. Example of the [FeIILn(NCS)2] Complexes. In Spin Crossover in Transition Metal Compounds II; Topics in Current Chemistry; Springer; Berlin Heidelberg, 2004; Vol. 234 pp 97–128. [Google Scholar]
- Halcrow M. A.Structure:Function Relationships in Molecular Spin-Crossover Materials. In Spin-Crossover Materials; John Wiley and Sons Ltd: Oxford, UK, 2013; pp 147–169. [Google Scholar]
- Ericsson M.-S.; Jagner S.; Ljungström E.; Tørneng E.; Woldbæk T.; Strand T. G.; Sukhoverkhov V. F. The Crystal Structure of Hexakis(Phenylisocyanide)Manganese(I) Tri-Iodide, [Mn(CNC6H5)6]I3. Acta Chem. Scand. 1980, 34a, 535–540. 10.3891/acta.chem.scand.34a-0535. [DOI] [Google Scholar]
- Cotton F. A.; Zingales F. The Donor-Acceptor Properties of Isonitriles as Estimated by Infrared Study. J. Am. Chem. Soc. 1961, 83, 351–355. 10.1021/ja01463a022. [DOI] [Google Scholar]
- Hahn F. E. The Coordination Chemistry of Multidentate Isocyanide Ligands. Angew. Chem., Int. Ed. 1993, 32, 650–665. 10.1002/anie.199306501. [DOI] [Google Scholar]
- Margulieux G. W.; Weidemann N.; Lacy D. C.; Moore C. E.; Rheingold A. L.; Figueroa J. S. Isocyano Analogues of [Co(CO)4]n: A Tetraisocyanide of Cobalt Isolated in Three States of Charge. J. Am. Chem. Soc. 2010, 132, 5033–5035. 10.1021/ja1012382. [DOI] [PubMed] [Google Scholar]
- Ditri T. B.; Carpenter A. E.; Ripatti D. S.; Moore C. E.; Rheingold A. L.; Figueroa J. S. Chloro- and Trifluoromethyl-Substituted Flanking-Ring m-Terphenyl Isocyanides: H6-Arene Binding to Zero-Valent Molybdenum Centers and Comparison to Alkyl-Substituted Derivatives. Inorg. Chem. 2013, 52, 13216–13229. 10.1021/ic402130p. [DOI] [PubMed] [Google Scholar]
- Ditri T. B.; Moore C. E.; Rheingold A. L.; Figueroa J. S. Oxidative Decarbonylation of m-Terphenyl Isocyanide Complexes of Molybdenum and Tungsten: Precursors to Low-Coordinate Isocyanide Complexes. Inorg. Chem. 2011, 50, 10448–10459. 10.1021/ic2015868. [DOI] [PubMed] [Google Scholar]
- Carpenter A. E.; Margulieux G. W.; Millard M. D.; Moore C. E.; Weidemann N.; Rheingold A. L.; Figueroa J. S. Zwitterionic Stabilization of a Reactive Cobalt Tris-Isocyanide Monoanion by Cation Coordination. Angew. Chem., Int. Ed. 2012, 51, 9412–9416. 10.1002/anie.201205058. [DOI] [PubMed] [Google Scholar]
- Claude G.; Salsi F.; Hagenbach A.; Gembicky M.; Neville M.; Chan C.; Figueroa J. S.; Abram U. Structural and Redox Variations in Technetium Complexes Supported by m-Terphenyl Isocyanides. Organometallics 2020, 39, 2287–2294. 10.1021/acs.organomet.0c00238. [DOI] [Google Scholar]
- Pižl M.; Hunter B. M.; Sazanovich I. V.; Towrie M.; Gray H. B.; Záliš S.; Vlček A. Excitation-Wavelength-Dependent Photophysics of d8d8 Di-Isocyanide Complexes. Inorg. Chem. 2022, 61, 2745–2759. 10.1021/acs.inorgchem.1c02645. [DOI] [PubMed] [Google Scholar]
- Nielson R. M.; Wherland S. Reduction Potential and Bonding Trends in Manganese(I) and Manganese(II) Hexakis(Aryl and Alkyl Isocyanides). Inorg. Chem. 1985, 24, 1803–1808. 10.1021/ic00206a021. [DOI] [Google Scholar]
- Sutton G. D.; Olumba M. E.; Nguyen Y. H.; Teets T. S. The Diverse Functions of Isocyanides in Phosphorescent Metal Complexes. Dalton Trans. 2021, 50, 17851–17863. 10.1039/D1DT03312C. [DOI] [PubMed] [Google Scholar]
- Joshi K. K.; Pauson P. L.; Stubbs W. H. Isocyanide Metal Complexes I. Reactions of Halocarbonyls with Phenylisocyanide. J. Organomet. Chem. 1963, 1, 51–57. 10.1016/S0022-328X(00)80045-5. [DOI] [Google Scholar]
- Fantucci P. C.; Valenti V.; Cariati F. Electronic Spectra of Some Isocyanide Complexes. Inorg. Chim. Acta 1971, 5, 425–428. 10.1016/S0020-1693(00)95957-7. [DOI] [Google Scholar]
- Fantucci P.; Naldini L.; Cariati F.; Valenti V.; Bussetto C.; Snam Progetti L. S. R. The Electronic Structure of Isocyanide Ligands and the Spectroscopic Behaviour of MnII Octahedral Complexes. J. Organomet. Chem. 1974, 64, 109–124. 10.1016/S0022-328X(00)93034-1. [DOI] [Google Scholar]
- Sarapu A. C.; Fenske R. F. Transition Metal-Isocyanide Bond. Approximate Molecular Orbital Study. Inorg. Chem. 1975, 14, 247–253. 10.1021/ic50144a006. [DOI] [Google Scholar]
- Verdonck L.; Tulun T.; van der Kelen G. P. Far-I.R. and Raman Spectra of (RNC)6Mn+ Compounds (R = CH3, C6H5, 4-Cl-C6H4). Spectrochim. Acta, Part A 1979, 35, 867–870. 10.1016/0584-8539(79)80049-5. [DOI] [Google Scholar]
- Cameron C. J.; Walton R. A.; Edwards D. A. The Vibrational Spectra of the Homoleptic Isocyanide Complexes [Re(CNR)6]PF6, (R = Me, Ph, and 4-MeC6H4). J. Organomet. Chem. 1984, 262, 335–346. 10.1016/S0022-328X(00)99161-7. [DOI] [Google Scholar]
- Wagner N. L.; Wüest A.; Christov I. P.; Popmintchev T.; Zhou X.; Murnane M. M.; Kapteyn H. C. Monitoring Molecular Dynamics Using Coherent Electrons from High Harmonic Generation. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 13279–13285. 10.1073/pnas.0605178103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chapados C.; Birnbaum G. Infrared Absorption of SF6 from 32 to 3000 cm–1 in the Gaseous and Liquid States. J. Mol. Spectrosc. 1988, 132, 323–351. 10.1016/0022-2852(88)90329-3. [DOI] [Google Scholar]
- Elgrishi N.; Rountree K. J.; McCarthy B. D.; Rountree E. S.; Eisenhart T. T.; Dempsey J. L. A Practical Beginner’s Guide to Cyclic Voltammetry. J. Chem. Educ. 2018, 95, 197–206. 10.1021/acs.jchemed.7b00361. [DOI] [Google Scholar]
- Henke W. C.; Otolski C. J.; Moore W. N. G.; Elles C. G.; Blakemore J. D. Ultrafast Spectroscopy of [Mn(CO)3] Complexes: Tuning the Kinetics of Light-Driven CO Release and Solvent Binding. Inorg. Chem. 2020, 59, 2178–2187. 10.1021/acs.inorgchem.9b02758. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kottelat E.; Lucarini F.; Crochet A.; Ruggi A.; Zobi F. Correlation of MLCTs of Group 7 fac-[M(CO)3]+ Complexes (M = Mn, Re) with Bipyridine, Pyridinylpyrazine, Azopyridine, and Pyridin-2-ylmethanimine Type Ligands for Rational PhotoCORM Design. Eur. J. Inorg. Chem. 2019, 2019, 3758–3768. 10.1002/ejic.201900568. [DOI] [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]
- Krishnan C. V.; Creutz C.; Mahajan D.; Schwarz H. A.; Sutin N. Homogeneous Catalysis of the Photoreduction of Water by Visible Light. 3. Mediation by Polypyridine Complexes of Ruthenium(II) and Cobalt(II). Isr. J. Chem. 1982, 22, 98–106. 10.1002/ijch.198200020. [DOI] [Google Scholar]
- Vaidyalingam A.; Dutta P. K. Analysis of the Photodecomposition Products of Ru(bpy)32+ in Various Buffers and upon Zeolite Encapsulation. Anal. Chem. 2000, 72, 5219–5224. 10.1021/ac000408e. [DOI] [PubMed] [Google Scholar]
- Soupart A.; Alary F.; Heully J. L.; Elliott P. I. P.; Dixon I. M. Recent Progress in Ligand Photorelease Reaction Mechanisms: Theoretical Insights Focusing on Ru(II) 3MC States. Coord. Chem. Rev. 2020, 408, 213184 10.1016/j.ccr.2020.213184. [DOI] [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]
- 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]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.