DOI:
10.1039/D5QI02274F
(Research Article)
Inorg. Chem. Front., 2026, Advance Article
Polycyclic aromatic hydrocarbon chromophores tune photo-physical properties in bichromophoric Cu(I) photosensitizers
Received
8th November 2025
, Accepted 5th January 2026
First published on 9th January 2026
Abstract
A modular design approach for bichromophoric Cu(I) complexes is reported, combining a [Cu(N^N)(P^P)]+ scaffold that supports metal-to-ligand charge transfer with two polycyclic aromatic hydrocarbon chromophores covalently attached to the diimine ligand. Three new systems incorporating 9,9-dimethyl-9H-fluorene, phenanthrene, and anthracene units at the 4,7-positions of neocuproine were synthesized and fully characterized. Structural analysis reveals twisted geometries of the attached chromophores, which reduce conjugation with the diimine core. Electrochemical measurements indicate only moderate shifts in reduction potentials, indicating preservation of the Cu(I)-centered redox behavior. Across the series, the visible MLCT band is retained without pronounced red-shifts, while the molar attenuation coefficients increase markedly (×1.7–3.4 vs. the reference), in line with calculated oscillator strengths. Fragment comparisons and computations attribute the additional UV intensity to intraligand charge transfer (9,9-dimethyl-9H-fluorene, phenanthrene) and to vibronically structured πA* ← πA transitions (anthracene), consistent with the non-coplanar geometries. Complexes bearing 9,9-dimethyl-9H-fluorene and phenanthrene substituents exhibit yellow emission involving the Cu(I)-based metal-to-ligand charge transfer state (λem ≈ 565 nm, τ up to 20.8 μs), whereas the anthracene-substituted system populates a non-emissive 3π–π* triplet state localized on an anthracene moiety (τ > 35 μs), as evidenced by transient absorption spectroscopy and time-dependent density functional theory. These results establish clear structure–property relationships in bichromophoric Cu(I) systems and illustrate how polycyclic aromatic hydrocarbon chromophores influence excited-state character and dynamics.
Introduction
The class of heteroleptic Cu(I) complexes has experienced a rise in popularity due to their diverse and successful applications in the field of photochemistry, including photoredox catalysis,1–3 dye-sensitized solar cells (DSSCs),4–7 light-emitting diodes8,9 or solar fuels production.10–16 In comparison to established and commercially available standards such as Ru(II) or Ir(III) complexes,17–19 the attractiveness of Cu(I)-based photosensitizers (CuPSs) can be attributed to different factors: (i) the significantly lower cost of copper due to its much higher abundance in the Earth's crust (Cu: 27 ppm vs. Ru: 37 × 10−6 ppm),20,21 (ii) the d10 shell, which prevents low-energy metal-centered (MC) states that can cause undesirably rapid excited-state deactivation, and (iii) the potential to outperform their noble-metal competitors.1,10,22,23 Numerous studies have already demonstrated that noble-metal-free systems for the reduction of water or carbon dioxide can be achieved using a heteroleptic CuPS along with a suitable earth-abundant-metal catalyst (e.g. based on Fe or Mn).11,14,24–27 Thus, further research and development of improved CuPSs can contribute to a more sustainable conversion of solar energy.
In general, heteroleptic CuPSs of the type [Cu(N^N)(P^P)]+ combine a diimine (N^N) and diphosphine (P^P) ligand that can both be tuned separately. The diimine ligand, e.g. neocuproine (2,9-dimethyl-1,10-phenanthroline, neo), has strong π-accepting properties that allow for the formation of low-lying excited metal-to-ligand charge transfer (MLCT) states.28 Hence, the diimine ligand is typically modified to alter the photophysical properties of the CuPS.3,29,30 One strategy to improve the relevant MLCT states, i.e. absorptivity and lifetime, is to attach an additional organic chromophore to the diimine to create a bichromophoric system.31–40 The attachment of the chromophore can be achieved via a single covalent bond with or without a linking group. The variety of linker groups is large and some common examples include alkyl, phenyl, triazole, thiophenyl, or alkynyl moieties.33,41–44 Moreover, the use of linking groups adds another dimension to the ligand design, because each bridge has its own properties e.g. rigidity and size.41,44,45 The advantages of combining chromophores can be numerous: (i) increased absorptivity (antenna effect), (ii) potential red-shifted absorption through π-conjugation, (iii) exploitation of high spin–orbit coupling on the sensitizer for efficient triplet generation and low spin–orbit coupling for the stabilization of long-lived triplet states on the organic chromophore, (iv) individual tunability, and (v) prolonged excited-state lifetimes through the reservoir effect.34–36,45–47 The latter is particularly attractive as the excited-state lifetime is a crucial factor for photocatalytic applications.48,49 However, synthesizing and analyzing such bichromophoric systems can be challenging and typically requires an interdisciplinary approach.
In a previous study, Doettinger et al. revealed by steady-state and time-resolved spectroscopy together with quantum-chemical computations that additional pyrene chromophores act as an energy sink for CuPS MLCT states.50 This concept enabled efficient singlet oxygen sensitization, with yields depending on the position (4,7 vs. 5,6) where the pyrene was attached to the neo ligand.50 In another study, Takeda et al. attached a variety of phenyl, thiophenyl and furan units to the 4,7-positions of neo.15 Most complexes retained their MLCT behavior, but heterocycles bearing a heteroatom in 2-position displayed altered excited states and reactivities.15
To expand the catalog of diimines and bichromophoric systems, three new ligands were developed in this study (Fig. 1) in which the 4,7-positions of neo are symmetrically disubstituted with polycyclic aromatic hydrocarbon (PAH) chromophores: 9,9-dimethyl-9H-fluorene (F), phenanthrene (P) and anthracene (A). To suppress potential deactivation channels, e.g. through n–π* states, heteroatoms were avoided in the substituents. According to the El-Sayed rule, these states allow higher crossing rates between systems.28,51 Furthermore, F, P and A are rigid and modifiable chromophores with well-defined absorption characteristics and sufficient photostability.52–58 For these reasons, they were used in diverse applications including fluorescence microscopy dyes,58 dye lasers,59 DSSCs,60 two-photon absorption59,61 and photon upconversion.54,57,62,63
 |
| | Fig. 1 Schematic representation of the diimine ligands and their respective heteroleptic Cu(I) complexes investigated in this study. The influence of different PAH chromophores (F = 9,9-dimethyl-9H-fluorene, P = phenanthrene, A = anthracene) on key photophysical and electrochemical properties is assessed. neo = neocuproine, also 2,9-dimethyl-1,10-phenanthroline, bcp = bathocuproine, also 2,9-dimethyl-4,7-diphenyl-1,10-phenanthroline, nF = 4,7-di-(9,9-dimethyl-9H-fluoren-2-yl)-2,9-dimethyl-1,10-phenanthroline, nP = 2,9-dimethyl-4,7-di(phenanthren-9-yl)-1,10-phenanthroline and nA = 4,7-di(anthracen-2-yl)-2,9-dimethyl-1,10-phenanthroline. | |
While PAHs have many desirable applications, this work focuses on the synthesis and in-depth characterization of the novel neo derivatives 4,7-bis(9,9-dimethyl-9H-fluoren-2-yl)-neocuproine (nF), 4,7-di(phenanthren-9-yl)-neocuproine (nP), 4,7-di(anthracen-2-yl)-neocuproine (nA) and their respective heteroleptic Cu(I) complexes CunF, CunP, and CunA (Fig. 1, P^P: xantphos, anion: PF6−). As references, neo and bathocuproine (bcp) as well as the complexes Cuneo and Cubcp were selected.
The following sections present structural, electrochemical and spectroscopic analysis supported by density functional theory (DFT) and time-dependent DFT (TDDFT) calculations, providing insights into the interplay between the different chromophores.
Results and discussion
Synthesis
The reference ligands neo and bcp were acquired commercially. The syntheses of the three novel ligands nF, nP and nA were conducted via Suzuki–Miyaura cross-coupling reactions64,65 from the common substrate 4,7-dichloro-2,9-dimethyl-1,10-phenanthroline66 and their corresponding boronic acids (for nF: 9,9-dimethyl-9H-fluoren-2-yl-2-boronic acid, for nP: phenanthrene-9-boronic acid, for nA: anthracene-2-boronic acid; further details see SI Chapter 2). The precursor 4,7-dichloro-2,9-dimethyl-1,10-phenanthroline was synthesized on gram-scale as described elsewhere.65,66 Subsequent cross-coupling reactions were conducted in a THF/water mixture using K3PO4 as base (0.5 M aqueous solution). To ensure efficient cross coupling, the XPhos-Pd-G2 precatalyst was employed, which under basic conditions releases a highly reactive XPhos-Pd0 species.67 Compared to the more commonly used [Pd(PPh3)4], the XPhos-Pd-G2 precatalyst is often a more convenient and reliable choice for transformations involving 4,7-dichloro- or 5,6-dibromo-1,10-phenanthroline derivatives, as shown in earlier studies.23,50,65 The target ligands nF, nP and nA were obtained after column chromatography on basic aluminum oxide in yields of 91%, 77% and 80%, respectively (for further details SI Chapter 2, 3 and 4).
The complexes were synthesized via a well-established two-step, one-pot approach,10,68,69 in which the xantphos ligand is first coordinated to the Cu(I) precursor [Cu(CH3CN)4]PF6 in deaerated dichloromethane at elevated temperatures (SI Chapter 2). Subsequently, the respective diimine ligand (nF, nP or nA) was added dropwise at 0 °C for at least one hour using an automatic syringe pump. Adding the reagent at low temperature prevents formation of the thermodynamically favored homoleptic [Cu(N^N)2]PF6 side product.3,70–73 Typically, isolation of the complexes proceeds via slow precipitation using n-hexane to afford crystalline solids. However, due to the bulky π-extended chromophores, the complexes synthesized herein formed deep red oils rather than solids upon n-hexane addition. This behavior was also observed for pyrenyl-substituted analogs and is likely caused by competing intermolecular π–π-interactions hindering crystal formation.50 Nevertheless, the target complexes CunF, CunP and CunA were successfully obtained in yields of 66%, 59% and 50%, respectively (see SI Chapter 2).
The identities and purities of all ligands and complexes were confirmed by nuclear magnetic resonance (NMR: 1H, 13C{1H}, DEPT 135, and 31P{1H}, see SI Chapter 3) and electrospray ionization high-resolution mass spectrometry (ESI-HRMS, SI Chapter 4). nP exhibited distinct behavior in the 1H NMR spectrum, revealing an unexpected pair of singlets for the methyl protons (3.08 ppm, Δ = 0.002 ppm), whereas a single singlet is typically observed. Combined with duplicate signals in the aromatic region of the 13C{1H} and DEPT135 spectra, this strongly suggests the presence of stereoisomers. This phenomenon is also observed for CunP, but is absent in all other ligands or complexes. However, due to the minor structural differences and the absence of distinctive spectroscopic or electrochemical effects, the presence of these isomers was considered negligible for further discussions.
Structural characterization
To elucidate the main structural similarities and differences between the designed complexes, DFT calculations were performed and the predicted molecular structures were analyzed. Additionally, single crystals of a solvate of CunA were obtained by slow diffusion of n-hexane into a concentrated dichloromethane/diethyl ether solution at room temperature, enabling a direct comparison between the experimental solid-state structure and the calculated geometry (Table 1, for experimental details: SI Chapter 5). The experimental (Fig. 2) and computed molecular structures (Fig. S5.1 and S5.2) are in good agreement and consistent with previous reports on heteroleptic Cu(I) complexes.15,16,29,50,65,74 Cu–N bond lengths (206.5–212.3 pm) are consistently shorter than Cu–P bond lengths (224.3–232.6 pm), and bite angles fall within expected pseudo-tetrahedral geometry ranges (N–Cu–N: 78.3–80.4°; P–Cu–P: 116.5–119.7°). The interplane angles θ between the planes formed by the N–Cu–N and P–Cu–P atoms (83.2–87.7°) are typical for such heteroleptic Cu(I) complexes.15,16,29,50,65,75,76
 |
| | Fig. 2 Molecular crystal structure of CunA. Nitrogen atoms are depicted in blue, oxygen in red, phosphorus in pink, copper in orange, and hydrogen as well as solvent molecules are omitted for clarity. The carbon atoms in 4,7-position are labeled with C4 and C7. Torsion angles ϑ are measured along C3–C4–CSub1,1–CSub1,2 and C8–C7–CSub2,1–CSub2,2. The second substituent CSub−2 is on the same side as the stacking phenyl ring. | |
Table 1 Selected bond lengths (pm), bond angles (°), N–Cu–N and P–Cu–P interplane angle θ (°) and torsion angles ϑ (°) from DFT (calc.) and X-ray (exp.) data for all complexes. The Cneo–Csub bond lengths describe the bond length between the respective substituent and the C4/C7 of neo. Torsion angles ϑsub1 and ϑsub2 express torsion angles along the substituent linkages from C3–C4–CSub1,1–CSub1,2 and C8–C7–CSub2,1–CSub2,2, where the second substituent carbon atom is on the same side as the stacking phenyl ring. Cubcp and CunA were analyzed both experimentally and computationally for validation (experimental data for Cubcp taken from ref. 16). Standard deviations are given in parentheses. Calculations at PBE0/def2-TZVP level (CPCM: CH2Cl2)
| |
|
Cuneocalc. |
Cubcpexp. |
Cubcpcalc. |
CunFcalc. |
CunPcalc. |
CunAexp. |
CunAcalc. |
| Cu–N1 |
(pm) |
210.9 |
207.26(19) |
210.3 |
210.7 |
211.5 |
210.62(33) |
211.5 |
| Cu–N2 |
212.3 |
209.63(19) |
211.4 |
211.1 |
211.1 |
211.14(38) |
210.3 |
| Cu–P1 |
226.3 |
223.99(6) |
226.0 |
226.3 |
226.7 |
226.29(13) |
225.5 |
| Cu–P2 |
229.6 |
230.19(6) |
229.9 |
229.2 |
229.1 |
231.68(12) |
232.6 |
| C4–Csub1 |
— |
147.5(3) |
147.6 |
147.4 |
148.1 |
149.16(61) |
147.3 |
| C7–Csub2 |
— |
148.4(3) |
147.6 |
147.5 |
148.1 |
149.77(65) |
147.4 |
| |
| N–Cu–N |
(°) |
79.7 |
80.41(7) |
79.2 |
79.2 |
79.4 |
78.253(142) |
79.0 |
| P–Cu–P |
|
118.5 |
119.95(2) |
118.7 |
118.8 |
119.0 |
116.789(47) |
116.5 |
| θ |
|
85.4 |
85.280(6) |
85.2 |
86.7 |
87.7 |
86.695(99) |
83.8 |
| ϑsub1 |
|
— |
−129.414(246) |
−122.3 |
−125.6 |
−82.4 |
50.330(613) |
−126.5 |
| ϑsub2 |
|
— |
124.560(278) |
55.0 |
55.1 |
104.6 |
50.183(618) |
127.1 |
In general, the geometry around the central Cu(I) atom is largely unaffected by the substituents, which is reasonable as the 4,7-positions are para to the copper-binding nitrogens. The C4–CSub1,1 and C7–CSub2,1 bond lengths (Table 1) confirm a carbon–carbon C–C single bond, with average values of 149.2 pm (exp.) and 147.7 pm (calc.). The observed torsion angles ϑsub1,2 along atoms C3–C4–CSub1,1–CSub1,2 and C8–C7–CSub2,1–CSub2,2 resemble those of biphenyl-like systems (ϑsolid = 0–10°, ϑvapor = 44.4 ± 1.2°, θsolute = 30–40°), where conjugation (ϑ ≈ 0°) and steric hindrance (ϑ ≈ 90°) compete.77–80 For the present series, the effective torsion between the neo plane and the substituent plane is 50.2–57.7° for Cubcp, CunF, and CunA, i.e. moderate twisting and partial conjugation limited by steric hindrance. This is primarily due to the repulsion between the ortho-hydrogens and the 5,6-hydrogens of the neo backbone, as also noted by Takeda et al.81
In contrast, with ϑsub1,2 = 82.4 and 75.4° the phenanthrene units in CunP are nearly perpendicular to the diimine plane. This can be explained by the repulsion between the ortho-hydrogen and the adjacent methine bridge with the 5,6-hydrogen atoms of neo. As a result, through-bond electronic communication between neo and phenanthrene in CunP is significantly reduced, because the π-systems and p-orbitals cannot effectively overlap. Overall, torsion increases in the following order: Cubcp ≈ CunF ≈ CunA (50.2–57.7°) < CunP (75.4–82.4°) and the expected π-communication follows: Cubcp ≈ CunF ≈ CunA > CunP.
Molecular orbital trends
To better understand the electronic communication and redox behavior, the molecular orbitals (MO) of the Cu(I) complexes (Fig. 3 and SI Chapter 6) were investigated (PBE0/def2-TZVP, CPCM: dichloromethane). The frontier molecular orbitals of Cuneo and Cubcp (Fig. S6.12 and S6.14) serve as reference systems: their highest occupied molecular orbital (HOMO) is predominantly dCu- and pP-based, and the lowest unoccupied molecular orbital (LUMO) corresponds to a πneo* orbital on the neo ligand, consistent with literature.73,75,76,82–84
 |
| | Fig. 3 Molecular orbital energy diagram calculated with PBE0/def2-TZVP. Energy differences are given relative to the Cuneo HOMO (−6.15 eV) for the respective molecule columns (in eV). Encircled markers denote the dCu-HOMO, diamonds denote the πneo*-LUMOs. Representative molecular orbitals of the latter are shown on the lefthand side for Cuneo, while the HOMO, LUMO and LUMO+2 of CunA are shown on the righthand side. | |
The HOMO (SI Chapter 6) is only marginally affected (<0.1 eV) upon 4,7-disubstitution of neo, and spin density analysis of the singly-reduced species (Fig. S6.22) corroborate a neo-centered LUMO. CunP and CunF retain πneo* LUMOs (Fig. S6.16 and S6.18) comparable in energy and character to the references Cuneo and Cubcp (Fig. 3, S6.12 and S6.14), reflecting minimal substituent contribution to πneo*-MOs (SI Chapter 6).
Rather than forming one uniform π-system, the MOs of neo and the substituents (phenyl in bcp, phenanthrene in nP and 9,9-dimethyl-9H-fluorene in nF) remain largely separate. This is consistent with the non-coplanar geometries (see above) and an energetic mismatch between the π-systems. The substituent orbitals of the latter are outside the shown frontier orbital window (Fig. 3). These findings highlight the bichromophoric nature of CunF and CunP, where electronic independence of the two chromophores is largely preserved. A similar separation is observed for the ligands (SI Chapter 6). Such behavior has also been observed for related systems containing carbazole, aniline or pyrene substituents.13,50,85
In contrast, CunA displays a unique MO behavior. Four additional MOs appear within the frontier orbital region, differentiating it from the other complexes (Fig. 3). Their assignment is complicated due to the energetically close MO energies of anthracene and neo as well as the torsion between the anthracene and diimine units, allowing some electronic interaction. The two highest occupied MOs (HOMO, HOMO−1) lie above the dCu MO (HOMO−2) and are clearly anthracene-centered (Fig. S6.20). The LUMOs, however, cannot unambiguously be assigned to either neo or A (Fig. S6.20). Thus, CunA differs from the other complexes in possessing notable LUMO stabilization and partial delocalization across both fragments, consistent with its distinct excited-state behavior discussed below.
Electrochemical behavior
Cyclic voltammetry (CV) was performed to evaluate the redox behavior and support the previously discussed MO assignments. For the unsubstituted complex Cuneo (Fig. S7.9), a reversible one-electron reduction is observed at −2.09 V vs. Fc/Fc+ in deaerated dimethylformamide (see Table 2).11,50,69 The localization of the reduction at the neo ligand was confirmed by spin density analysis of the singly reduced species (Fig. S6.22). Oxidative events were not analyzed in detail, as oxidation in heteroleptic Cu(I) complexes typically leads to irreversible cleavage of the Cu–P bond at ≈0.9 V vs. Fc/Fc+.23,24,86–88
Table 2 Reduction potentials of the substituents, ligands and complexes measured in deaerated dimethylformamide and referenced against the ferrocene/ferrocenium couple (Fc/Fc+). All irreversible events are simply listed at their peak (minimum) current without special notation. Reversible signals are given as the half wave potential (Ered,1/2)
| |
— |
— |
F |
P |
A |
| Ered/V |
— |
— |
−3.13b |
−2.73 |
−2.41b |
| −2.91b |
| |
neo |
bcp |
nF |
nP |
nA |
| Ered/V |
ca. −2.7a |
−2.60 |
−2.42 |
−2.46 |
−2.23b |
| −2.46 |
−2.53 |
−2.68 |
| −2.70 |
−2.75 |
|
| −2.76 |
|
|
| |
Cuneo |
Cubcp |
CunF |
CunP |
CunA |
| No distinct minimum determinable. reversible. |
| Ered/V |
−2.09b |
−2.01b |
−1.99b |
−2.00b |
−1.94b |
| −2.83 |
−2.63 |
−2.47b |
−2.67 |
−2.25 |
| |
|
−2.86 |
−2.76 |
−2.50 |
| |
|
−2.94 |
|
−2.70 |
The introduction of PAH chromophores at the 4,7-positions of neo leads to a slight anodic shift in the reduction potential (Table 2). For Cubcp, a small shift to −2.01 V (Δ = +80 mV) is observed (Fig. S7.10), in line with the computations (Table S6.11). For CunF and CunP, the reduction waves are shifted by ≈ +100 mV (Table 2 and Fig. S7.11, S7.12) and correspond to the same neo-centered reduction. The computed shift for CunF (+80 mV) matches well with the experiment, while the shift for CunP (+40 mV) is underestimated, likely due to the lack of computed contribution from the phenanthrene unit (Fig. S6.22).
CunA exhibits a more complex reduction behavior, with multiple reductions between −1.9 V and −2.7 V (Fig. S7.13). A comparison with related systems and its ligand nA (Fig. S7.8) suggests that the reduction at −1.94 V corresponds to the neo based reduction (Δ = +150 mV vs. Cuneo). The calculated shift (Δ = +230 mV) is overestimated due to strong orbital contributions from anthracene, but the overall trend is reproduced. Furthermore, CunA has two weaker reversible reductions around −1.8 V and −1.2 V, which are also present in nA (Fig. S7.8), but not in A itself (Fig. S7.3). These features were not predicted by DFT in the present setup and may arise from intermolecular and solvent interactions or side reactions.
In summary, the observed anodic shifts in neo-based reductions correlate with the size of the attached π-system (Δ ≈ +100 mV for F and P, +150 mV for A), but remain modest overall. This supports the concept that the Cu(I) redox core is largely decoupled from the chromophoric PAH substituents. The calculated reduction potentials follow the same trend (see SI Table S6.11). The disagreement with experiment likely arises from the minimal orbital overlap in CunP and significant orbital mixing in CunA (Fig. S6.22).
UV/vis absorption
To explore the bichromophoric character of the novel complexes, UV/vis absorption spectra were recorded. Substitution at the 4,7-positions of neo with PAH chromophores results in only small red-shifts, but a drastic increase in molar attenuation coefficients (Fig. 4 and Table 3). This trend has also been observed for related substitution patterns.3,12,89
 |
| | Fig. 4 UV/vis absorption spectra of CunF (top), CunP (middle), and CunA (bottom) compared to their individual fragments. Top: CunF (blue), nF (light blue), neo (black), F (dark blue). Middle: CunP (green), nP (light green), neo (black), P (dark green). Bottom: CunA (red), nA (light red), neo (black), A (dark red). Difference density plots in the top right of the insets show the migration of the electron density from blue (−) to yellow (+) during the MLCT excitations (S1 ← S0). An additional anthracene-localized πA*–πA excitation (S8 ← S0) is shown for CunA. | |
Table 3 Experimental absorption wavelengths λabs,exp (nm), extinction coefficients ε (103 M−1cm−1), emission wavelengths λem,exp (nm), emission quantum yields Φem,exp (%), emission lifetime τem,exp (μs), transient absorption lifetime τtA,exp (μs), compared to calculated fluorescence wavelengths λfl,calc (relaxed S1 → unrelaxed S0, TDDFT) and phosphorescence wavelengths (relaxed T1 → unrelaxed S0, DFT) λph,calc. All measurements were carried out in dichloromethane under inert conditions at 298 K. Computations were performed with implicit dichloromethane. The singlet–triplet energy gap ΔES1–T1,calc (eV) and interplane angles (θ, °) of the optimized singlet S1 and triplet T1 are given. Optimization of the S1 state in CunA yielded two different excited states depending on the sophistication of the method (see computational details). The first and lowest one is the 1MLCT and the second one was a 1πA–πA*. The latter ended up as S2 after convergence. Experimental data for Cubcp referenced from ref. 92
| |
λabs/nm (ε/103 M−1 cm−1) |
λem,exp./nm |
Φem,exp./% |
τem,exp./μs |
τtA,exp./μs |
λfl,calc./nm |
λph,calc./nm |
ΔES1–T1,calc./eV |
S1 θcalc./° |
T1 θcalc./° |
| Cuneo |
276 (47.4), 400 (3.29) |
559 |
27.9 |
7.04 |
6.51 |
599.4 |
741.9 |
0.388 |
70.1 |
70.4 |
| Cubcp |
287(59.8), 400(5.58) |
568 |
38.0 (ref. 92) |
8.38 |
8.04 |
580.3 |
756.2 |
0.404 |
74.3 |
71.0 |
| CunF |
269(56.6), 290(57.7) |
570 |
44.2 |
20.78 |
21.62 |
576.9 |
745.5 |
0.385 |
73.7 |
71.6 |
| 325(49.2), 400 (10.3) |
| CunP |
255(143.6), 278(83.2) |
564 |
28.1 |
9.74 |
8.82 |
600.0 |
752.5 |
0.384 |
70.4 |
70.7 |
| 325 (18.9), 400 (6.32) |
| CunA |
252(114.6), 285(75.8) |
442 |
0.4 |
n.d. |
37.28 |
608.0/476.5 |
866.6 |
0.722/1.036 |
75.5/82.8 |
83.8 |
| 335(24.6), 400 (11.1) |
174.36 |
All complexes exhibit a broad, structureless MLCT band between 350–450 nm, with ε ≈ 3–15 × 103 M−1 cm−1. This assignment is supported by: (i) previous reports about Cuneo, Cubcp and related complexes,29,71,87,90,92 (ii) TDDFT calculations, which consistently predict that the S1 state is an MLCT state (SI Chapter 6), and (iii) the absence of this band in the ligands (Fig. 4, SI Chapter 6 and Fig. S8.2).
As expected, Cuneo has the weakest MLCT absorption (Table 3), followed by Cubcp.87,93 CunP exhibits a similar enhancement, suggesting that phenanthrene has a comparable influence to phenyl. In contrast, CunF and CunA both possess significantly higher attenuation coefficients.
TDDFT-derived difference densities confirm for all complexes that the lowest MLCT excitations (states 1–4, except CunA: 1, 4, 7, 11) have negligible substituent contribution (SI Chapter 6), consistent with the minimal red-shift. This indicates that the MLCT transitions are localized on the Cu(I)/diimine core, electronically decoupled from the substituents.15,81 A comparable trend can be seen for the oscillator strength of the S1 ← S0 transition, suggesting that enhanced absorptivity stems from the elongated diimine system and a larger transition dipole moment.75
In the UV region (250–350 nm), the absorption is dominated by π* ← π transitions within the diimine ligand and the substituents.29,71,87,90,91 Cuneo and Cubcp exhibit typical ligand-centered πneo* ← πneo transitions, corroborated by TDDFT (Fig. S6.2, S6.4, S6.13 and S6.15) and fragment spectrum (Fig. S8.4). Additional excitations involving xantphos and nonbonding MOs are present, but not discussed further.
Introducing PAH chromophores significantly alters the band shape and increases absorption (Fig. 4). A new shoulder at 300–350 nm appears experimentally (Fig. 4 and S8.5–8.7) and computationally (SI Chapter 6).
For nP and CunP, TDDFT indicates intraligand charge transfer (ILCT) from P to neo (Fig. S6.8 transition 2 and Fig. S6.19 transitions 5,6) and an enhancement of the weak S1 ← S0 transition in P and neo (Fig. S6.8 state 1 vs. S6.2 state 3). The MLCT transitions remain centered on the core complex, supporting an electronic decoupling of the two chromophores.
A similar observation is made for nF and CunF: between 250–350 nm, strong deviations from the fragments F and neo occur (Fig. 4 or S8.5). TDDFT reveals enhanced πneo* ← πneo transitions (Fig. S6.6 states 1,2) and strong ILCT transitions from F to neo (πneo* ← πF, see Fig. S6.17 states 5–8,10). The bands at ≈290 and ≈310 nm in CunF and nF likely reflect fluorene-specific transitions, though an unambiguous assignment is not possible. Electronic communication between neo and F is somewhat stronger than in CunP, but a continuous π-system is still not formed.
For CunA, a subtle but distinct set of features is found: the MLCT band has two overlapping maxima at 385 and 366 nm. These are also present in nA and are attributed to vibronically structured anthracene-centered πA* ← πA transitions (Fig. 4 and S6.10 states 1,2). Additional transitions near the MLCT region display mixed local (πA* ← πA) and charge transfer (πneo* ← πA) character (S6.21 states 2, 3, 5, 6). Similar to the other systems, no significant red-shift of the MLCT band is observed, indicating negligible anthracene participation in the MLCT excitations. In the 300–350 nm range, however, CunA and nA display similar ILCT/π–π* patterns, suggesting an electronic communication between neo and A comparable to CunF.
Emission behavior
The ligands neo, bcp, nF, nP and nA emit broadly in the 350 to 500 nm range (Fig. S9.1–3) and their quantum yields (vs. pyrene93) follow the order: nA (428 nm, ϕem = 46.4%) > nF (396 nm; ϕem = 29.6%) > nP (379 nm; ϕem = 6.2%) > bcp (385 nm; ϕem = 4.6%) > neo (364 nm; ϕem = 0.7%).
Upon excitation of the MLCT band at 390 nm, all complexes except CunA display yellow MLCT emission with a typically broad and featureless peak around 565 nm in deaerated dichloromethane (Fig. 5) as has been reported for Cuneo87 and Cubcp.11,29,76 CunF and CunP possess nearly identical emission wavelengths and band shapes compared to the references, differing mainly in emission intensity and quantum yield (referenced vs. Cubcp).92 The quantum yield ϕem order is: CunF (44.2%) > Cubcp (38.0%92) > CunP (28.1%) ≈ Cuneo (27.9%, Table 3). A plausible explanation for this order combines different torsions (reduced overlap for CunP) with increased substituent rigidity in CunF (fluorene bridge), which disfavors non-radiative decay. Moreover, CunF exceeds Cubcp possibly due to the additional phenyl ring, which is forced in plane by the dimethylmethylene bridge.
 |
| | Fig. 5 Emission spectra of Cuneo (black), Cubcp (gray), CunF (blue), CunP (green) and CunA (red) at excitation wavelength of 390 nm in deaerated dichloromethane. Samples were excited at an optical density of 0.1. | |
Time-resolved emission (SI Chapter 10) gives lifetimes of 8.38 µs for Cubcp (cf. 7.24 µs, Forero-Cortés et al.92), 7.08 µs for Cuneo, 9.70 µs for CunP, indicating a negligible effect, and a significantly longer lifetime of 20.78 µs for CunF. Nevertheless, it is assumed that the final emissive state has MLCT characteristics, because the emission wavelength and shape are comparable to the references. Furthermore, computations reinforce a 1MLCT assignment (Table 3): (i) TDDFT-optimized S1 states converged to the flattened 1MLCT state with fluorescence wavelengths closely matching experimentally observed values (Table 3); (ii) DFT-optimized T1 states also converged to the flattened MLCT state, but the calculated phosphorescence wavelengths are strongly red-shifted due to underestimated triplet energies, which can be seen in the large calculated singlet–triplet gaps ΔES1–T1 (>0.3 eV, Table 3).3,29,87,91,94,95 Lastly, both singlet S1 and triplet T1 geometries demonstrate the characteristic MLCT flattening, where the interplane angle θ decreases from ≈86° to ≈72° on average (Table S5.2 and S5.3).29
In CunA the 1MLCT emission is absent and only a very weak band at 442 nm (Φem < 1%) attributed to the nA ligand is detected under deaerated conditions. TDDFT indicates two closely spaced singlet manifolds, a dCu–πneo* 1MLCT state and an anthracene-localized 1πA–πA* state, whereby both can be obtained. The MLCT state yielded a theoretical emission wavelength of 608.0 nm with similar flattening of θ, while the anthracene-localized state yielded 476.5 nm with no geometric distortions to the Cu(I) core. Triplet–state ΔSCF optimizations reveal an anthracene-centered 3πA–πA* level lying well below the 3MLCT manifold, creating an energetically favorable, non-emissive energy sink. This enables efficient triplet–triplet energy transfer from the 3MLCT state, thereby quenching the 1MLCT emission and explaining its absence in CunA. Alternatively, the closely-spaced 1πA–πA* state may benefit from increased spin–orbit-coupling because of the Cu(I) chromophore to populate 3πA–πA* levels more easily.
To sum it up, the emission data establish a clear structure–property relationship: 9,9-dimethyl-9H-fluorene and phenanthrene substitution retain bright, long-lived MLCT emission, whereas anthracene diverts excitation into a low-lying, non-radiative triplet. This provides a practical design guideline: choose F/P substituents to retain a long-lived MLCT excited state with high radiative yield, and A when efficient triplet generation/storage is sought.
Photostability and transient absorption spectroscopy
Prolonged irradiation with an intense, UV-rich light source (e.g. a 150 W Xe arc) facilitates a slow, steady photodecomposition (Fig. S12.1–S12.10). Both complexes and ligands in dichloromethane yield the same emissive photoproducts (Fig. S12.11, λem = 488 nm for Cuneo; 444 nm for Cubcp; 508 nm for CunF; 520 nm for CunP; 617 nm for CunA). This suggests transformations within the diimine ligand rather than the Cu/xantphos fragment. Two non-exclusive hypotheses are consistent with literature: (i) light-induced dimerization/oligomerization enlarging the π-system (reported for bcp dimers with green fluorescence),96,97 or (ii) reactions with radicals generated from chlorinated solvents.98,99
The transient absorption (tA, 355 nm pump) of Cuneo and Cubcp are characterized by very broad peaks around 521 and 534 nm (Fig. 6) with tA lifetimes (τtA) of 6.51 μs and 8.04 μs, respectively. These lifetimes are monoexponential and comparable to the emissive lifetime. The tA spectrum has been attributed to the triplet–triplet absorption of the 3MLCT state, which matches the spin density of the lowest computed triplet (Fig. S6.23) and the computed triplet–triplet absorption spectrum (Fig. S6.24).16,65 The band at about 525 nm is characterized by two excitation types, namely one local πneo* ← πneo transition and ligand-to-metal charge transfers reducing the formal Cu(II) center (Fig. S6.24).
The PAH chromophore F could not be excited with 355 nm, but according to literature data of its derivative, 9H-fluorene has one sharp characteristic triplet–triplet absorption peak at around 375–390 nm.100 However, the tA spectrum of CunF (Fig. 6) displays a broad maximum at 645 nm with a shoulder at 460 nm, which do not match the fluorene signature. Instead, the computed triplet–triplet spectrum of the 3MLCT state fits the spectral signature (Fig. S6.25). Compared to the reference complexes, the 3MLCT tA lifetime is effectively doubled to 21.62 μs with a new spectral signature.
 |
| | Fig. 6 Nanosecond transient absorption spectra (top) of Cuneo (black), Cubcp (gray), CunF (blue), CunP (green) and CunA (red) in deaerated dichloromethane approximately 10 ns after excitation at 355 nm (absorbance of ≈0.3) with a detection gate-width of 3 μs. The decay kinetics (bottom) have been recorded at the transient absorption maxima: Cuneo (black, 521 nm), Cubcp (gray, 530 nm), CunF (blue, 645 nm), CunP (green, 505 nm) and CunA (red, 436 nm). See Fig S11.2 for full details on CunA. All complexes with the exception of CunA display a broad featureless transient spectrum that is attributed to the 3MLCT absorption and a representative spin density of the 3MLCT state of CunF is shown. CunA exhibits a sharp peak with a shoulder at 415 and 437 nm that is attributed to the absorption of the 3πA–πA* state of A as also shown by the spin density in the top left.100 | |
The tA spectroscopy of P reveals two distinct peaks at 460 and 492 nm with a shoulder at 432 nm100 (Fig. S11.1). CunP (τtA = 8.82 μs), however, does not display these peaks and exhibits a similar spectral signature (Fig. 6) as Cuneo and Cubcp with one broad feature at ≈508 nm and another at around 700 nm. This, together with the matching computed 3MLCT-triplet absorption (SI Fig. S6.25), suggests that the triplet state of P is not populated. This is reasonable as the 3MLCT is energetically lower than the triplet states of P and F.101–103
A also possesses two sharp characteristic triplet–triplet absorption bands at roughly 404 and 428 nm.100 In contrast to both CunF and CunP, these characteristic peaks can also be found slightly shifted within CunA at 415 nm and 437 nm (Fig. 6 and S11.2). Furthermore, the tA spectrum of CunA agrees with the computed triplet–triplet absorption spectrum of the 3πA–πA* state (Fig. S6.26). This differentiates the PAH chromophores F and P from A, which efficiently funnels the excited state population into the low-lying anthracene-based 3πA–πA* T1 state leading to vastly different photophysical properties. This is further evidenced in the significantly longer biexponential excited state lifetime (τtA = 37.28 and 174.36 μs).
In agreement with tA and the aforementioned findings (i.e. HOMO–LUMO gaps, electrochemistry, absorption and emission spectra) an energy level diagram of the novel complexes is proposed in Fig. S13.1 to illustrate possible photophysical decay pathways. To summarize, the lowest excited state in the four systems Cuneo, Cubcp, CunF and CunP has been identified to be of 3MLCT character. From a purely energetic perspective, the 3π–π* states of the substituents phenyl, F and P lie too high in energy in order to function as a triplet reservoir. Despite the latter, a significant extension of the lifetime has been observed for CunF which is about twice as large as in Cubcp, whereas that of CunP stayed similar to Cuneo. It is likely that this difference is attributed to geometric differences (e.g. more favorable dihedral angles). CunA on the other hand possesses very low-lying 3πA–πA* states that act as an energy sink, so that a repopulation of the 3MLCT state is no longer possible.
Conclusions and outlook
In this study, three novel bichromophoric Cu(I) complexes CunF, CunP, and CunA were successfully synthesized and fully characterized. The diimine ligands are based on neocuproine (neo) and selectively functionalized at the 4,7-positions with polycyclic aromatic hydrocarbon chromophores – 9,9-dimethyl-9H-fluorene, phenanthrene, and anthracene. Structural analysis by X-ray diffraction (CunA) and DFT optimization confirms pseudo-tetrahedral geometries around the copper center. Due to steric hindrance, the attached organic chromophores adopt twisted, non-coplanar conformations relative to the diimine core. Electrochemical and spectroscopic analyses, supported by (TD-)DFT computations, evidence that the MLCT characteristics of the Cu(I) scaffold are largely retained for CunF and CunP, while CunA exhibits pronounced electronic mixing in the excited-state manifold. All complexes display strong MLCT absorption in the visible range with significantly increased molar attenuation coefficients compared to the references (Cuneo and Cubcp). CunF and CunP exhibit MLCT-based emission in the yellow region (λem ≈ 565 nm) with quantum yields of 44.2% and 28.1%, respectively, and excited-state lifetimes up to 20.8 μs (CunF). In strong contrast, CunA has no detectable MLCT emission. Instead, a non-emissive, anthracene-localized 3πA–πA* state is populated, as confirmed by transient absorption spectroscopy and supported by TDDFT. Cyclic voltammetry revealed only moderate anodic shifts in the first reduction upon substitution with PAH chromophores. This indicates that the LUMO, at least for CunF and CunP, remains neo-centered and that the ground-state redox properties are largely preserved.
Three concise conclusions emerge: (i) structural torsion limits conjugation and preserves the Cu(I) redox core while increasing absorption strength; (ii) CunF/CunP retain MLCT excited states with long lifetimes, (iii) anthracene redirects the excited state into a low-lying, dark 3πA–πA* state that suppresses MLCT emission. Thus, substitution with PAH chromophores enables systematic tuning of absorption properties and excited-state dynamics without fundamentally altering the redox behavior, if the correct chromophore is chosen. CunF combines strong absorption, long lifetime and favorable redox potential, making it a promising candidate for photoredox catalysis. Similarly, CunP's excited-state and redox properties compare directly to Cuneo and Cubcp, but with significantly stronger UV absorption. The unique photophysical behavior of CunA enabled by a very long-lived locally excited triplet state opens up perspectives for its application in energy transfer catalysis and photon upconversion, e.g. via triplet–triplet annihilation (TTA).
In future studies, both directions will be explored. Moreover, the modular approach of combining Cu(I)-MLCT chromophores with rigid, non-coplanar π-systems may be extended to other platforms, offering a versatile strategy to design photoactive metal complexes with tailored excited-state landscapes.
Conflicts of interest
The authors declare no competing financial interests.
Data availability
The data supporting this article have been included as part of the supplementary information (SI). Supplementary information: additional experimental, synthetic, computational, spectroscopic and electrochemical details. See DOI: https://doi.org/10.1039/d5qi02274f.
CCDC 2491672 (CunA) contains the supplementary crystallographic data for this paper.104
Acknowledgements
We are grateful to Cengiz Özmen (TU Braunschweig) for help with solvent purification. L. E. B. is thankful for a doctoral scholarship by the Deutsche Bundesstiftung Umwelt (DBU, AZ: 20024/029). J. P. Z. expresses gratitude to the Cusanuswerk e.V. for a doctoral scholarship.
References
- C. B. Larsen and O. S. Wenger, Photoredox Catalysis with Metal Complexes Made from Earth-Abundant Elements, Chem. – Eur. J., 2018, 24, 2039–2058 Search PubMed.
- C. Sandoval-Pauker, G. Molina-Aguirre and B. Pinter, Status report on copper(I) complexes in photoredox catalysis; photophysical and electrochemical properties and future prospects, Polyhedron, 2021, 199, 115105 Search PubMed.
- J. Beaudelot, S. Oger, S. Peruško, T.-A. Phan, T. Teunens, C. Moucheron and G. Evano, Photoactive Copper Complexes: Properties and Applications, Chem. Rev., 2022, 122, 16365–16609 Search PubMed.
- C. E. Housecroft and E. C. Constable, The emergence of copper(I)-based dye sensitized solar cells, Chem. Soc. Rev., 2015, 44, 8386–8398 Search PubMed.
- A. Lennert and D. M. Guldi, Homoleptic and Heteroleptic Copper Complexes as Redox Couples in Dye-Sensitized Solar Cells, ChemPhotoChem, 2019, 3, 636–644 Search PubMed.
- D. Franchi, V. Leandri, A. R. P. Pizzichetti, B. Xu, Y. Hao, W. Zhang, T. Sloboda, S. Svanström, U. B. Cappel, L. Kloo, L. Sun and J. M. Gardner, Effect of the Ancillary Ligand on the Performance of Heteroleptic Cu(I) Diimine Complexes as Dyes in Dye-Sensitized Solar Cells, ACS Appl. Energy Mater., 2022, 5, 1460–1470 Search PubMed.
- L. E. Burmeister, F. Doettinger, K. J. Haseloff, C. Kleeberg, M. Boujtita, S. Pascal, F. Odobel, S. Tschierlei, Y. Pellegrin and M. Karnahl, Dye-Sensitized Solar Cells Based on Cu(I) Complexes Containing Catechol Anchor Groups That Operate with Aqueous Electrolytes, JACS Au, 2025, 5, 3960–3973 Search PubMed.
- Q. Zhang, J. Ding, Y. Cheng, L. Wang, Z. Xie, X. Jing and F. Wang, Novel Heteroleptic CuI Complexes with Tunable Emission Color for Efficient Phosphorescent Light-Emitting Diodes, Adv. Funct. Mater., 2007, 17, 2983–2990 Search PubMed.
- D. M. Zink, D. Volz, T. Baumann, M. Mydlak, H. Flügge, J. Friedrichs, M. Nieger and S. Bräse, Dinuclear Copper(I) Complexes for Application in Organic Light-Emitting Diodes, Chem. Mater., 2013, 25, 4471–4486 Search PubMed.
- S.-P. Luo, E. Mejía, A. Friedrich, A. Pazidis, H. Junge, A.-E. Surkus, R. Jackstell, S. Denurra, S. Gladiali, S. Lochbrunner and M. Beller, Photocatalytic Water Reduction with Copper-Based Photosensitizers: A Noble-Metal-Free System, Angew. Chem., Int. Ed., 2013, 52, 419–423 Search PubMed.
- E. Mejía, S.-P. Luo, M. Karnahl, A. Friedrich, S. Tschierlei, A.-E. Surkus, H. Junge, S. Gladiali, S. Lochbrunner and M. Beller, A Noble-Metal-Free System for Photocatalytic Hydrogen Production from Water, Chem. – Eur. J., 2013, 19, 15972–15978 Search PubMed.
- N.-Y. Chen, L.-M. Xia, A. J. J. Lennox, Y.-Y. Sun, H. Chen, H.-M. Jin, H. Junge, Q.-A. Wu, J.-H. Jia, M. Beller and S.-P. Luo, Structure-Activated Copper Photosensitisers for Photocatalytic Water Reduction, Chem. – Eur. J., 2017, 23, 3631–3636 Search PubMed.
- J. Kim, D. R. Whang and S. Y. Park, Designing Highly Efficient CuI Photosensitizers for Photocatalytic H2 Evolution from Water, ChemSusChem, 2017, 10, 1883–1886 Search PubMed.
- H. Takeda, H. Kamiyama, K. Okamoto, M. Irimajiri, T. Mizutani, K. Koike, A. Sekine and O. Ishitani, Highly Efficient and Robust Photocatalytic Systems for CO2 Reduction Consisting
of a Cu(I) Photosensitizer and Mn(I) Catalysts, J. Am. Chem. Soc., 2018, 140, 17241–17254 Search PubMed.
- H. Takeda, Y. Monma, H. Sugiyama, H. Uekusa and O. Ishitani, Development of Visible-Light Driven Cu(I) Complex Photosensitizers for Photocatalytic CO2 Reduction, Front. Chem., 2019, 7, 418 Search PubMed.
- J.-W. Wang, X. Zhang, L. Velasco, M. Karnahl, Z. Li, Z.-M. Luo, Y. Huang, J. Yu, W. Hu, X. Zhang, K. Yamauchi, K. Sakai, D. Moonshiram and G. Ouyang, Precious-Metal-Free CO2 Photoreduction Boosted by Dynamic Coordinative Interaction between Pyridine-Tethered Cu(I) Sensitizers and a Co(II) Catalyst, JACS Au, 2023, 3, 1984–1997 Search PubMed.
- R. D. Costa, E. Ortí, H. J. Bolink, F. Monti, G. Accorsi and N. Armaroli, Luminescent Ionic Transition-Metal Complexes for Light-Emitting Electrochemical Cells, Angew. Chem., Int. Ed., 2012, 51, 8178–8211 Search PubMed.
- Y.-J. Yuan, Z.-T. Yu, D.-Q. Chen and Z.-G. Zou, Metal-complex chromophores for solar hydrogen generation, Chem. Soc. Rev., 2017, 46, 603–631 Search PubMed.
- T. Mede, M. Jäger and U. S. Schubert, “Chemistry-on-the-complex”: functional RuII polypyridyl-type sensitizers as divergent building blocks, Chem. Soc. Rev., 2018, 47, 7577–7627 Search PubMed.
- O. S. Wenger, Photoactive Complexes with Earth-Abundant Metals, J. Am. Chem. Soc., 2018, 140, 13522–13533 Search PubMed.
- B. M. Hockin, C. Li, N. Robertson and E. Zysman-Colman, Photoredox catalysts based on earth-abundant metal complexes, Catal. Sci. Technol., 2019, 9, 889–915 Search PubMed.
- S. Paria and O. Reiser, Copper in Photocatalysis, ChemCatChem, 2014, 6, 2477–2483 Search PubMed.
- F. Doettinger, C. Kleeberg, C. Queffélec, S. Tschierlei, Y. Pellegrin and M. Karnahl, Rich or poor: the impact of electron donation and withdrawal on the photophysical and photocatalytic properties of copper(I) complexes, Catal. Sci. Technol., 2023, 13, 4092–4106 Search PubMed.
- S. Fischer, D. Hollmann, S. Tschierlei, M. Karnahl, N. Rockstroh, E. Barsch, P. Schwarzbach, S.-P. Luo, H. Junge, M. Beller, S. Lochbrunner, R. Ludwig and A. Brückner, Death and Rebirth: Photocatalytic Hydrogen Production by a Self-Organizing Copper–Iron System, ACS Catal., 2014, 4, 1845–1849 Search PubMed.
- H. Takeda, K. Ohashi, A. Sekine and O. Ishitani, Photocatalytic CO2 Reduction Using Cu(I) Photosensitizers with a Fe(II) Catalyst, J. Am. Chem. Soc., 2016, 138, 4354–4357 Search PubMed.
- C. Steinlechner, A. F. Roesel, E. Oberem, A. Päpcke, N. Rockstroh, F. Gloaguen, S. Lochbrunner, R. Ludwig, A. Spannenberg, H. Junge, R. Francke and M. Beller, Selective Earth-Abundant System for CO2 Reduction: Comparing Photo- and Electrocatalytic Processes, ACS Catal., 2019, 9, 2091–2100 Search PubMed.
- X. Zhang, M. Cibian, A. Call, K. Yamauchi and K. Sakai, Photochemical CO2 Reduction Driven by Water-Soluble Copper(I) Photosensitizer with the Catalysis Accelerated by Multi-Electron Chargeable Cobalt Porphyrin, ACS Catal., 2019, 9, 11263–11273 Search PubMed.
- G. Accorsi, A. Listorti, K. Yoosaf and N. Armaroli, 1,10-Phenanthrolines: versatile building blocks for luminescent molecules, materials and metal complexes, Chem. Soc. Rev., 2009, 38, 1690–1700 Search PubMed.
- Y. Zhang, M. Schulz, M. Wächtler, M. Karnahl and B. Dietzek, Heteroleptic diimine–diphosphine
Cu(I) complexes as an alternative towards noble-metal based photosensitizers: Design strategies, photophysical properties and perspective applications, Coord. Chem. Rev., 2018, 356, 127–146 Search PubMed.
- P. A. Forero Cortés, M. Marx, M. Trose and M. Beller, Heteroleptic copper complexes with nitrogen and phosphorus ligands in photocatalysis: Overview and perspectives, Chem. Catal., 2021, 1, 298–338 Search PubMed.
- C. Weinheimer, Y. Choi, T. Caldwell, P. Gresham and J. Olmsted III, Effect of a steric spacer on chromophoric interactions of ruthenium complexes containing covalently bound anthracene, J. Photochem. Photobiol., A, 1994, 78, 119–126 Search PubMed.
- G. J. Wilson, W. H. F. Sasse and A. W.-H. Mau, Singlet and triplet energy transfer processes in ruthenium(II) bipyridine complexes containing covalently bound arenes, Chem. Phys. Lett., 1996, 250, 583–588 Search PubMed.
- G. J. Wilson, A. Launikonis, W. H. F. Sasse and A. W.-H. Mau, Excited-State Processes in Ruthenium(II) Bipyridine Complexes Containing Covalently Bound Arenes, J. Phys. Chem. A, 1997, 101, 4860–4866 Search PubMed.
- N. D. McClenaghan, Y. Leydet, B. Maubert, M. T. Indelli and S. Campagna, Excited-state equilibration: a process leading to long-lived metal-to-ligand charge transfer luminescence in supramolecular systems, Coord. Chem. Rev., 2005, 249, 1336–1350 Search PubMed.
- H. Li, C. Wang, F. Glaser, N. Sinha and O. S. Wenger, Metal-Organic Bichromophore Lowers the Upconversion Excitation Power Threshold and Promotes UV Photoreactions, J. Am. Chem. Soc., 2023, 145, 11402–11414 Search PubMed.
- F. N. Castellano, Altering Molecular Photophysics by Merging Organic and Inorganic Chromophores, Acc. Chem. Res., 2015, 48, 828–839 Search PubMed.
- M.-S. Bertrams, K. Hermainski, J.-M. Mörsdorf, J. Ballmann and C. Kerzig, Triplet quenching pathway control with molecular dyads enables the identification of a highly oxidizing annihilator class, Chem. Sci., 2023, 14, 8583–8591 Search PubMed.
- J. E. Yarnell, J. C. Deaton, C. E. McCusker and F. N. Castellano, Bidirectional “Ping-Pong” Energy Transfer and 3000-Fold Lifetime Enhancement in a Re(I) Charge Transfer Complex, Inorg. Chem., 2011, 50, 7820–7830 Search PubMed.
- N. Sinha, C. Wegeberg, D. Häussinger, A. Prescimone and O. S. Wenger, Photoredox-active Cr(0) luminophores featuring photophysical properties competitive with Ru(II) and Os(II) complexes, Nat. Chem., 2023, 15, 1730–1736 Search PubMed.
- F. N. Castellano and M. C. Rosko, Steric and Electronic Influence of Excited-State Decay in Cu(I) MLCT Chromophores, Acc. Chem. Res., 2024, 57, 2872–2886 Search PubMed.
- G. E. Shillito, C. B. Larsen, J. R. W. McLay, N. T. Lucas and K. C. Gordon, Effect of Bridge Alteration on Ground- and Excited-State Properties of Ruthenium(II) Complexes with Electron-Donor-Substituted Dipyrido[3,2-a:2′,3′-c]phenazine Ligands, Inorg. Chem., 2016, 55, 11170–11184 Search PubMed.
- P. Dierks, A. Päpcke, O. S. Bokareva, B. Altenburger, T. Reuter, K. Heinze, O. Kühn, S. Lochbrunner and M. Bauer, Ground- and Excited-State Properties of Iron(II) Complexes Linked to Organic Chromophores, Inorg. Chem., 2020, 59, 14746–14761 Search PubMed.
- S. Neumann, O. S. Wenger and C. Kerzig, Controlling Spin-Correlated Radical Pairs with Donor–Acceptor Dyads: A New Concept to Generate Reduced Metal Complexes for More Efficient Photocatalysis, Chem. – Eur. J., 2021, 27, 4115–4123 Search PubMed.
- G. E. Shillito, D. Preston, J. D. Crowley, P. Wagner, S. J. Harris, K. C. Gordon and S. Kupfer, Controlling Excited State Localization in Bichromophoric Photosensitizers via the Bridging Group, Inorg. Chem., 2024, 63, 4947–4956 Search PubMed.
- A. Lavie-Cambot, C. Lincheneau, M. Cantuel, Y. Leydet and N. D. McClenaghan, Reversible electronic energy transfer: a means to govern excited-state properties of supramolecular systems, Chem. Soc. Rev., 2010, 39, 506–515 Search PubMed.
- D. S. Tyson and F. N. Castellano, Intramolecular Singlet and Triplet Energy Transfer in a Ruthenium(II) Diimine Complex Containing Multiple Pyrenyl Chromophores, J. Phys. Chem. A, 1999, 103, 10955–10960 Search PubMed.
- F. Odobel and H. Zabri, Preparations and Characterizations of Bichromophoric Systems Composed of a Ruthenium Polypyridine Complex Connected to a Difluoroborazaindacene or a Zinc Phthalocyanine Chromophore, Inorg. Chem., 2005, 44, 5600–5611 Search PubMed.
- F. Glaser, C. Kerzig and O. S. Wenger, Multi-Photon Excitation in Photoredox Catalysis: Concepts, Applications, Methods, Angew. Chem., Int. Ed., 2020, 59, 10266–10284 Search PubMed.
- D. M. Arias-Rotondo and J. K. McCusker, The photophysics of photoredox catalysis: a roadmap for catalyst design, Chem. Soc. Rev., 2016, 45, 5803–5820 Search PubMed.
- F. Doettinger, Y. Yang, M. Karnahl and S. Tschierlei, Bichromophoric Photosensitizers: How and Where to Attach Pyrene Moieties to Phenanthroline to Generate Copper(I) Complexes, Inorg. Chem., 2023, 62, 8166–8178 Search PubMed.
- M. Baba, Intersystem Crossing in the 1nπ* and 1ππ* States, J. Phys. Chem. A, 2011, 115, 9514–9519 Search PubMed.
- A. J. Floyd, S. F. Dyke and S. E. Ward, The synthesis of phenanthrenes, Chem. Rev., 1976, 76, 509–562 Search PubMed.
- M. Yoshizawa and J. K. Klosterman, Molecular architectures of multi-anthracene assemblies, Chem. Soc. Rev., 2014, 43, 1885–1898 Search PubMed.
- V. Gray, D. Dzebo, A. Lundin, J. Alborzpour, M. Abrahamsson, B. Albinsson and K. Moth-Poulsen, Photophysical characterization of the 9,10-disubstituted anthracene chromophore and its applications in triplet–triplet annihilation photon upconversion, J. Mater. Chem. C, 2015, 3, 11111–11121 Search PubMed.
- J. Shaya, F. Fontaine-Vive, B. Y. Michel and A. Burger, Rational Design of Push–Pull Fluorene Dyes: Synthesis and Structure-Photophysics Relationship, Chem. – Eur. J., 2016, 22, 10627–10637 Search PubMed.
- G. S. Baviera and P. M. Donate, Recent advances in the syntheses of anthracene derivatives, Beilstein J. Org. Chem., 2021, 17, 2028–2050 Search PubMed.
- A. Olesund, V. Gray, J. Mårtensson and B. Albinsson, Diphenylanthracene Dimers for Triplet–Triplet Annihilation Photon Upconversion: Mechanistic Insights for Intramolecular Pathways and the Importance of Molecular Geometry, J. Am. Chem. Soc., 2021, 143, 5745–5754 Search PubMed.
- J. Shaya, P. R. Corridon, B. Al-Omari, A. Aoudi, A. Shunnar, M. I. H. Mohideen, A. Qurashi, B. Y. Michel and A. Burger, Design, photophysical properties, and applications of fluorene-based fluorophores in two-photon fluorescence bioimaging: A review, J. Photochem. Photobiol., C, 2022, 52, 100529 Search PubMed.
- K. D. Belfield, M. V. Bondar, C. O. Yanez, F. E. Hernandez and O. V. Przhonska, Two-photon absorption and lasing properties of new fluorene derivatives, J. Mater. Chem., 2009, 19, 7498–7502 Search PubMed.
- K. Lim, C. Kim, J. Song, T. Yu, W. Lim, K. Song, P. Wang, N. Zu and J. Ko, Enhancing the Performance of Organic Dye-Sensitized Solar Cells via a Slight Structure Modification, J. Phys. Chem. C, 2011, 115, 22640–22646 Search PubMed.
- M. Pawlicki, H. A. Collins, R. G. Denning and H. L. Anderson, Two-Photon Absorption and the Design of Two-Photon Dyes, Angew. Chem., Int. Ed., 2009, 48, 3244–3266 Search PubMed.
- T. N. Singh-Rachford and F. N. Castellano, Photon upconversion based on sensitized triplet-triplet annihilation, Coord. Chem. Rev., 2010, 254, 2560–2573 Search PubMed.
- F. Deng, J. Blumhoff and F. N. Castellano, Annihilation Limit of a Visible-to-UV Photon Upconversion Composition Ascertained from Transient Absorption Kinetics, J. Phys. Chem. A, 2013, 117, 4412–4419 Search PubMed.
- A. Suzuki, Carbon-carbon bonding made easy, Chem. Commun., 2005, 4759–4763 Search PubMed.
- F. Doettinger, Y. Yang, M.-A. Schmid, W. Frey, M. Karnahl and S. Tschierlei, Cross-Coupled Phenyl- and Alkynyl-Based Phenanthrolines and Their Effect on the Photophysical and Electrochemical Properties of Heteroleptic Cu(I) Photosensitizers, Inorg. Chem., 2021, 60, 5391–5401 Search PubMed.
- J. E. Nycz, J. Wantulok, R. Sokolova, L. Pajchel, M. Stankevič, M. Szala, J. G. Malecki and D. Swoboda, Synthesis and Electrochemical and Spectroscopic Characterization of 4,7-diamino-1,10-phenanthrolines and Their Precursors, Molecules, 2019, 24, 4102 Search PubMed.
- T. Kinzel, Y. Zhang and S. L. Buchwald, A New Palladium Precatalyst Allows for the Fast Suzuki−Miyaura Coupling Reactions of Unstable Polyfluorophenyl and 2-Heteroaryl Boronic Acids, J. Am. Chem. Soc., 2010, 132, 14073–14075 Search PubMed.
- Y. Zhang, M. Heberle, M. Wächtler, M. Karnahl and B. Dietzek, Determination of side products in the photocatalytic generation of hydrogen with copper photosensitizers by resonance Raman spectroelectrochemistry, RSC Adv., 2016, 6, 105801–105805 Search PubMed.
- M. Heberle, S. Tschierlei, N. Rockstroh, M. Ringenberg, W. Frey, H. Junge, M. Beller, S. Lochbrunner and M. Karnahl, Heteroleptic Copper Photosensitizers: Why an Extended π-System Does Not Automatically Lead to Enhanced Hydrogen Production, Chem. – Eur. J., 2017, 23, 312–319 Search PubMed.
- M. S. Lazorski and F. N. Castellano, Advances in the light conversion properties of Cu(I)-based photosensitizers, Polyhedron, 2014, 82, 57–70 Search PubMed.
- M. Sandroni, Y. Pellegrin and F. Odobel, Heteroleptic bis-diimine copper(I) complexes for applications in solar energy conversion, C. R. Chim., 2016, 19, 79–93 Search PubMed.
- V. Leandri, A. R. P. Pizzichetti, B. Xu, D. Franchi, W. Zhang, I. Benesperi, M. Freitag, L. Sun, L. Kloo and J. M. Gardner, Exploring the Optical and Electrochemical Properties of Homoleptic versus Heteroleptic Diimine Copper(I) Complexes, Inorg. Chem., 2019, 58, 12167–12177 Search PubMed.
- J. Beaudelot, G. Evano and C. Moucheron, Structure-Property Relationships in a New Family of Photoactive Diimine-Diphosphine Copper(I) Complexes, Chem. – Eur. J., 2023, 29, e202300758 Search PubMed.
- J. R. Kirchhoff, D. R. McMillin, W. R. Robinson, D. R. Powell, A. T. McKenzie and S. Chen, Steric effects and the behavior of Cu(NN)(PPh3)2+ systems in fluid solution. Crystal and molecular structures of [Cu(dmp)(PPh3)2]NO3 and [Cu(phen)(PPh3)2]NO3·1.5EtOH, Inorg. Chem., 1985, 24, 3928–3933 Search PubMed.
- T. Tsubomura, K. Kimura, M. Nishikawa and T. Tsukuda, Structures and photophysical properties of copper(I) complexes bearing diphenylphenanthroline and bis(diphenylphosphino)alkane: the effect of phenyl groups on the phenanthroline ligand, Dalton Trans., 2015, 44, 7554–7562 Search PubMed.
- D. Moonshiram, P. Garrido-Barros, C. Gimbert-Suriñach, A. Picón, C. Liu, X. Zhang, M. Karnahl and A. Llobet, Elucidating the Nature of the Excited State of a Heteroleptic Copper Photosensitizer by using Time-Resolved X-ray Absorption Spectroscopy, Chem. – Eur. J., 2018, 24, 6464–6472 Search PubMed.
- F. Grein, Twist Angles and Rotational Energy Barriers of Biphenyl and Substituted Biphenyls, J. Phys. Chem. A, 2002, 106, 3823–3827 Search PubMed.
- B. Milián-Medina and J. Gierschner, π-Conjugation, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 513–524 Search PubMed.
- J. Jia, H.-S. Wu, Z. Chen and Y. Mo, Elucidation of the Forces Governing the Stereochemistry of Biphenyl, Eur. J. Org. Chem., 2013, 611–616 Search PubMed.
- B. Landeros-Rivera and J. Hernández-Trujillo, Control of Molecular Conformation and Crystal Packing of Biphenyl Derivatives, ChemPlusChem, 2022, 87, e202100492 Search PubMed.
- H. Takeda, M. Shimo, G. Yasuhara, K. Kobori, M. S. Asano and Y. Amao, Heteroleptic Cu(I) Phenanthroline Complexes Bearing Benzoxazole and Benzothiazole Moieties for Visible Light Absorption, Chem. Lett., 2021, 51, 69–72 Search PubMed.
- M. Iwamura, S. Takeuchi and T. Tahara, Ultrafast Excited-State Dynamics of Copper(I) Complexes, Acc. Chem. Res., 2015, 48, 782–791 Search PubMed.
- C. Li, R. Dickson, N. Rockstroh, J. Rabeah, D. B. Cordes, A. M. Z. Slawin, P. Hünemörder, A. Spannenberg, M. Bühl, E. Mejía, E. Zysman-Colman and P. C. J. Kamer, Ligand electronic fine-tuning and its repercussion on the photocatalytic activity and mechanistic pathways of the copper-photocatalysed aza-Henry reaction, Catal. Sci. Technol., 2020, 10, 7745–7756 Search PubMed.
- Y. Yamazaki, T. Onoda, J. Ishikawa, S. Furukawa, C. Tanaka, T. Utsugi and T. Tsubomura, Photocatalytic CO2 Reduction Using Various Heteroleptic Diimine-Diphosphine Cu(I) Complexes as Photosensitizers, Front. Chem., 2019, 7, 288 Search PubMed.
- Z.-J. Yu, H. Chen, A. J. Lennox, L.-J. Yan, X.-F. Liu, D.-D. Xu, F. Chen, L.-X. Xu, Y. Li, Q.-A. Wu and S.-P. Luo, Heteroleptic copper(I) photosensitizers with carbazole-substituted phenanthroline ligands: Synthesis, photophysical properties and application to photocatalytic H2 generation, Dyes Pigm., 2019, 162, 771–775 Search PubMed.
- A. Kaeser, M. Mohankumar, J. Mohanraj, F. Monti, M. Holler, J.-J. Cid, O. Moudam, I. Nierengarten, L. Karmazin-Brelot, C. Duhayon, B. Delavaux-Nicot, N. Armaroli and J.-F. Nierengarten, Heteroleptic Copper(I) Complexes Prepared from Phenanthroline and Bis-Phosphine Ligands, Inorg. Chem., 2013, 52, 12140–12151 Search PubMed.
- C. Li, C. F. R. Mackenzie, S. A. Said, A. K. Pal, M. A. Haghighatbin, A. Babaei, M. Sessolo, D. B. Cordes, A. M. Z. Slawin, P. C. J. Kamer, H. J. Bolink, C. F. Hogan and E. Zysman-Colman, Wide-Bite-Angle Diphosphine Ligands in Thermally Activated Delayed Fluorescent Copper(I) Complexes: Impact on the Performance of Electroluminescence Applications, Inorg. Chem., 2021, 60, 10323–10339 Search PubMed.
- M. Rentschler, M.-A. Schmid, W. Frey, S. Tschierlei and M. Karnahl, 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 Search PubMed.
- F. Doettinger, M. Obermeier, V. Caliskanyürek, L. E. Burmeister, C. Kleeberg, M. Karnahl, M. Schwalbe and S. Tschierlei, Bright or Dark: Unexpected Photocatalytic Behavior of Closely Related Phenanthroline-based Copper(I) Photosensitizers, ChemCatChem, 2023, 15, e202300452 Search PubMed.
- N. Armaroli, Photoactive mono- and polynuclear Cu(I)–phenanthrolines. A viable alternative to Ru(I)–polypyridines?, Chem. Soc. Rev., 2001, 30, 113–124 Search PubMed.
- C. Sandoval-Pauker, M. Santander-Nelli and P. Dreyse, Thermally activated delayed fluorescence in luminescent cationic copper(I) complexes, RSC Adv., 2022, 12, 10653–10674 Search PubMed.
- P. A. Forero-Cortés, M. Marx, N. G. Moustakas, F. Brunner, C. E. Housecroft, E. C. Constable, H. Junge, M. Beller and J. Strunk, Transferring photocatalytic CO2 reduction mediated by Cu(N^N)(P^P)+ complexes from organic solvents into ionic liquid media, Green Chem., 2020, 22, 4541–4549 Search PubMed.
- D. S. Karpovich and G. J. Blanchard, Relating the polarity-dependent fluorescence
response of pyrene to vibronic coupling. Achieving a fundamental understanding of the py polarity scale, J. Phys. Chem., 1995, 99, 3951–3958 Search PubMed.
- M. R. Silva-Junior, M. Schreiber, S. P. A. Sauer and W. Thiel, Benchmarks for electronically excited states: Time-dependent density functional theory and density functional theory based multireference configuration interaction, J. Chem. Phys., 2008, 129, 104103 Search PubMed.
- R. Grotjahn, T. M. Maier, J. Michl and M. Kaupp, Development of a TDDFT-Based Protocol with Local Hybrid Functionals for the Screening of Potential Singlet Fission Chromophores, J. Chem. Theory Comput., 2017, 13, 4984–4996 Search PubMed.
- S. Scholz, C. Corten, K. Walzer, D. Kuckling and K. Leo, Photochemical reactions in organic semiconductor thin films, Org. Electron., 2007, 8, 709–717 Search PubMed.
- M. Miśnik, K. Falkowski, W. Mróz and W. Stampor, Electromodulation of photoluminescence in vacuum-evaporated films of bathocuproine, Chem. Phys., 2013, 410, 45–54 Search PubMed.
- S. Yamada, H. Kuroki, M. Maeda, M. Takagi and T. Yamashita, Characterization of Pyrene Photoproducts in Chloroform by Gas-Chromatography and Gas-Chromatography Mass-Spectrometry, Microchem. J., 1994, 49, 117–125 Search PubMed.
- J. Aguilera-Sigalat, J. Sanchez-SanMartín, C. E. Agudelo-Morales, E. Zaballos, R. E. Galian and J. Pérez-Prieto, Further Insight into the Photostability of the Pyrene Fluorophore in Halogenated Solvents, ChemPhysChem, 2012, 13, 835–844 Search PubMed.
- I. Carmichael and G. L. Hug, Triplet–Triplet Absorption Spectra of Organic Molecules in Condensed Phases, J. Phys. Chem. Ref. Data, 1986, 15, 1–250 Search PubMed.
- D. S. McClure, Triplet-Singlet Transitions in Organic Molecules. Lifetime Measurements of the Triplet State, J. Chem. Phys., 1949, 17, 905–913 Search PubMed.
- J. Langelaar, R. P. H. Rettschnick and G. J. Hoijtink, Studies on Triplet Radiative Lifetimes, Phosphorescence, and Delayed Fluorescence Yields of Aromatic Hydrocarbons in Liquid Solutions, J. Chem. Phys., 1971, 54, 1–7 Search PubMed.
- S. Scypinski and L. J. C. Love, Room-temperature phosphorescence of polynuclear aromatic hydrocarbons in cyclodextrins, Anal. Chem., 1984, 56, 322–327 Search PubMed.
- CCDC 2491672: Experimental Crystal Structure Determination, 2026, DOI:10.5517/ccdc.csd.cc2pmsjn.
Footnote |
| † These authors share the first authorship of this paper. |
|
| This journal is © the Partner Organisations 2026 |
Click here to see how this site uses Cookies. View our privacy policy here.