Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

π-Extended diphosphonium-bridged ladder stilbenes: water-soluble fluorophores with up to eight annulated rings

Sebastian Senn a, Jean-Marc Mörsdorf *a, Maria-Sophie Bertrams b, Christoph Kerzig *b and Joachim Ballmann a
aAnorganisch-Chemisches Institut Universität Heidelberg, Im Neuenheimer Feld 276, D-69120 Heidelberg, Germany. E-mail: jean-marc.moersdorf@aci.uni-heidelberg.de
bDepartment of Chemistry, Johannes Gutenberg University Mainz, Duesbergweg 10-14, D-55128 Mainz, Germany. E-mail: ckerzig@uni-mainz.de

Received 23rd May 2025 , Accepted 26th September 2025

First published on 29th September 2025


Abstract

Diphosphapentalene-derived P-heterocyclic materials with two directly fused phospholes are fairly scarce, at least in comparison to their simpler congeners containing only one phosphole entity. To fill that void, π-conjugated naphtho-fused phospholo[3,2-b]phosphole dications were prepared via in situ oxidation of the corresponding diphosphines. In the case of one specific naphtho-annulation pattern, a hitherto unprecedented bis-(Δ2-phosphetene) dication was formed selectively and isolated as a colorless powder. DFT modelling studies revealed that this bis-(Δ2-phosphetenium) salt is produced via single electron transfer steps, while all the phospholo[3,2-b]phosphole salts may either be generated via their P-diylidic counterparts or via similar radical mechanisms. Exploiting this knowledge, the dicationic phospholo[3,2-b]phosphole isomer of the bis-(Δ2-phosphetenium) salt was isolated as well. In view of the high fluorescence quantum yields of these naphtho-fused phospholo[3,2-b]phosphole salts in aqueous solution, linearly π-extended anthraceno-fused derivatives were developed in order to bathochromically shift their emissions into the biological window. While detailed optoelectronic studies confirmed our expectations, the utmost remarkable observation is that even the anthraceno-fused materials were found to be sufficiently soluble in water, despite the fact that these fluorophores comprise up to eight fused rings.


Introduction

Cyclic π-conjugated organophosphorus materials comprising 4-(phosphetenes),1 5-(phospholes),2 6-(phosphinines)3 and 7-membered rings (phosphepines)4 are known for their manifold applications, not only in organic electronics5 and (bio)imaging,6 but also in (photo)catalysis.7 Despite the seemingly endless possibilities to fuse regular arenes to P-containing rings with different ring sizes, π-extended phospholes are targeted most frequently due to the availability of various synthetic protocols for their preparation8 and due to the fact that phospholes are readily converted to P(V)-oxides, P(V)-sulphides and P(V)-phospholium cations, which are known to be intrinsically air-stable.9 The latter property is certainly of crucial importance for most applications, while the presence of two exocyclic P(V)-substituents is commonly considered beneficial considering that hyperconjugative interactions10 between the exocyclic C–P bonds and the endocyclic π-systems are easily exploited in order to fine-tune the electronic properties of the ensuing materials. Hence, π-extended phosphole oxides and phospholium salts have attracted considerable attention over the past years,11 which also led to the development of novel synthetic routes for their preparation.12 A decade ago, phosphines or phospholes had to be isolated, for example for the preparation of A13 and B14, while methods that avoid these potentially air-sensitive intermediates only emerged quite recently.12 Along these lines, a straightforward one-pot procedure for the preparation of remarkably robust P-ylides (C) was reported by Jacquemin and Bouit,15 while Leifert and Studer demonstrated that quinoline-fused phosphole oxides, such as D, may be assembled directly from phosphine oxides via a new radical cyclization protocol.16 These particularly elegant synthetic routes to π-extended mono-phosphole-derived materials, however, are rarely applicable if two mutually fused phospholes need to be incorporated into a given π-framework. Hence, 6π phospholo–phospholes and 8π diphosphapentalenes are best regarded as a compound class of their own, despite the fact that only a few members of this class have been discovered so far. Prior to 2008, solely carboxylate-substituted diphosphapentalenes, such as E (see Scheme 1),17 were known, which changed when Yamaguchi's group prepared the first bis-(PhP[double bond, length as m-dash]O)-bridged ladder stilbenes and succeeded in separating the cis- and trans-isomer of F.18 Separating isomeric mixtures by column chromatography,19 however, may be tedious, which is considered a major drawback, in particular if the P-oxides are introduced in the last step. Starting from simple phosphinous and phosphinic acids, Huang and Xiao recently managed to prepare dioxaphosphorane-fused diphosphacycles, such as G,20 which were shown to reversibly ring-open upon addition of OH. Targeting more robust diphosphacycles, we have set our focus on bis-(R2P+)-bridged ladder stilbenes (H with n = 2) and their use as novel triplet–triplet annihilators in aqueous solution.21 In previous studies, the dication of H (n = 2) has been prepared via oxidation of the corresponding P-diylidic diphosphapentalene (H with n = 0), which also led to the isolation of the intermediate radical cation (H with n = 1).22a In contrast to F, no stereochemical peculiarities are to be expected in compounds of type H, which prompted us to exploit this advantage by seeking for π-extended derivatives with bathochromically shifted absorption and fluorescence bands. Herein, the envisioned naphtho- and anthraceno-fused diphosphonium-bridged ladder stilbenes (see Scheme 1) have been prepared without isolating the aforementioned (air-sensitive) P-diylidic species. Over the course of our study, an unexpected and hitherto unprecedented bis-(Δ2-phosphetene) dication (see Scheme 1) was obtained selectively, which paved the way for a comparative DFT-aided mechanistic analysis. This analysis revealed that the thermodynamically favoured phospholo[3,2-b]phosphole dications may be formed via three different mechanisms (depending on the reaction conditions), while the formation of the bis-(Δ2-phosphetenium) salt relies on a commencing SET (single electron transfer) step to initiate the conversion. In the following, these findings are discussed in more depth and presented in conjunction with detailed photophysical studies.
image file: d5sc03752b-s1.tif
Scheme 1 Selection of π-conjugated mono-phosphole-based (top, A–D) and diphosphapentalene-derived materials (middle, E–H) together with the most important structural motifs presented herein (bottom).

Results and discussion

Given that H has been prepared previously starting from 2,2′-dibromotolane,22a similar π-extended dibromo acetylenes were targeted at first. Although many excellent emitters containing the Ph2P+ unit to connect aromatic systems were reported previously,22b–e we decided to focus on iPr2P+-bridged systems as we observed a slightly higher emission quantum yield and, more importantly, a much higher photostability for the iPr2P+-containing system in a comparative study as well as an improved water solubility (see ref. 22a, Table S9 and related text). Starting from two different regioisomeric bromo-iodonaphthalenes (1a and 1b, see Scheme 2), the di(bromonaphthyl) acetylenes (2a and 2b) were obtained via Stille coupling with Bu3SnC[triple bond, length as m-dash]CSnBu3. Exchange of the bromides in 2a and 2b for lithium, followed by treatment of the resulting dilithiated intermediates with ClPiPr2, led to the expected diphosphines, 3a and 3b, which were detected by 31P{1H} NMR spectroscopy (δ(31P) = 6.6 ppm for 3a, δ(31P) = 2.0 ppm for 3b), but not isolated. Upon oxidation of 3a with C2Cl6 at room temperature, the expected bis-(iPr2P+)-bridged ladder stilbene ([4a]2+ with δ(31P) = 55.9 ppm) was obtained as an air-stable red- orange powder exhibiting a bright orange fluorescence (vide infra). Similarly, the bis-[PF6] salt of [4a]2+ was obtained upon oxidation of 3a with Fc[PF6].
image file: d5sc03752b-s2.tif
Scheme 2 Synthesis of [4a]2+, [4a-Ph]2+, [iso-4b]2+ and [4b-Ph]2+ (with [X] = [Cl] or [PF6]) together with the crystallographically determined molecular structures of the chloride salts (co-crystallized solvents, disorder and hydrogen atoms omitted for clarity, thermal ellipsoids drawn at the 50% probability level).

Treatment of 3b with C2Cl6, however, led to an air-sensitive, nearly colourless powder featuring a significantly downfield-shifted singlet (δ(31P) = 97.5 ppm) in the 31P{1H} NMR spectrum. These observations suggested that a significantly different species was produced, which was also obtained upon oxidation of 3b with Fc[PF6]. Single crystal X-ray diffraction (scXRD) unambiguously confirmed that an unexpected bis-(Δ2-phosphetene) dication ([iso-4b]2+, see Scheme 2) was generated. In [iso-4b]2+, two 4-membered phosphetenium cations are interconnected via a shared C[double bond, length as m-dash]C bond (dC[double bond, length as m-dash]C = 1.336(4) Å), thus forging a planar, but not fully ring-fused bis-(iPr2P+)-containing ladder stilbene. To the best of our knowledge, bis-(Δ2-phosphetenes) akin to [iso-4b]2+ have not been reported previously, while uncharged mono-phosphetenes with an exocyclic double bond are well-known.1c,e,23 To address the question whether each naphthalene unit in 3a and 3b is needed to dictate the reaction outcome ([4a]2+vs. [iso-4b]2+), one naphthyl unit in 2a and 2b was replaced for a phenyl entity (cf.2a-Ph and 2b-Ph). Employing the aforementioned methodology via dilithiation, phosphination and consecutive oxidative cyclization (vide supra), [4a-Ph]2+ (δ(31P) = 58.2 and 56.2 ppm) and [4b-Ph]2+ (δ(31P) = 58.3 and 66.3 ppm) were obtained as orange powders and found to exhibit 31P{1H} NMR shifts akin to the one observed for [4a]2+ (δ(31P) = 55.9 ppm). These NMR data strongly suggested that 5-membered rings ([4a-Ph]2+ and [4b-Ph]2+) were produced, which was confirmed by scXRD (see Scheme 2). On this basis, it is clear that the bis-(Δ2-phosphetene) dication [iso-4b]2+ is truly exceptional and that the specific naphtho-annulation pattern present in 3b is needed at both ends of the central alkyne to selectively forge the 4-membered rings in [iso-4b]2+.

To further elucidate these observations, DFT modelling studies (r2SCAN-3c, D4, def2-mTZVPP, CPCM for CH2Cl2)24 were carried out. Given that [H]0 (see Scheme 1) has been previously isolated in form of its P-diylide and oxidized via two consecutive SET steps, we at first assumed that a similar cyclization → SET → SET mechanism (cf. mechanism A in Fig. 1) may be operative here. In such a scenario, P-diylidic diphosphapentalenes are formed via carbene intermediates,25 which leads to 5-membered rings in all cases ([4a]0, [4b]0, [4a-Ph]0 and [4b-Ph]0). The formation of neutral P-diylidic bis-(Δ2-phosphetenes), however, is thermodynamically and kinetically prohibited (activation barriers of >35 kcal mol−1). Hence, mechanism A was excluded as it fails to explain the formation of [iso-4b]0 (and thus the formation of its experimentally observed dication [iso-4b]2+), at least at room temperature. Yet, it is noted that the barriers along the way to each P-diylidic diphosphapentalene are very reasonable, which suggested that this mechanistic pathway may be exploited in the absence of an oxidant (vide infra). In the search for a mechanism, which actually explains that [iso-4b]2+ is produced selectively in our experiments, two alternative mechanisms were considered, namely a SET → SET → cyclization (mechanism B) and a SET → cyclization → SET sequence (mechanism C). In mechanism B, a chlorophosphorus(V) chloride intermediate ([INT-Cl]+Cl, see Fig. 1) is expected to form after two consecutive SET steps, which has been proposed previously for similar cyclization reactions.26 Starting from [INT-Cl]+Cl, each 5-endo-dig cyclization is thermodynamically favoured and predicted to occur for [4a]2+, [4a-Ph]2+ and [4b-Ph]2+ (see Fig. 1 and SI for details). Upon oxidation of 3b to its chlorophosphorus(V) chloride intermediate, however, the kinetic product [iso-4b]2+ is expected to form at room temperature, while the thermodynamic product [4b]2+ is inaccessible given that a barrier of approximately 30 kcal mol−1 was calculated for the conversion of [iso-4b]2+ to [4b]2+. Hence, mechanism B is in line with the experimental findings and considered plausible, but only if a chlorine synthon (such as C2Cl6) was used to actually produce [INT-Cl]+Cl in the first place.


image file: d5sc03752b-f1.tif
Fig. 1 Mechanistic proposals (mechanism A, B and C) for the formation of [4a]2+, [4a-Ph]2+, [4b-Ph]2+ and [iso-4b]2+ based on DFT calculations (r2SCAN-3c, D4, def2-mTZVPP, CPCM for CH2Cl2). Additional energy profiles and relaxed PES scans are provided in the SI.

Considering that the use of Fc[PF6] in non-chlorinated solvents also led to [4a]2+, [4a-Ph]2+, [4b-Ph]2+ and [iso-4b]2+, yet another mechanism was studied in silico, namely the aforementioned SET → cyclization → SET sequence (mechanism C). In this scenario, radical cations (3˙+) are produced in the first SET step, which were identified as local minima in the case of 3a˙+, 3a-Ph˙+ and 3b-Ph˙+. For these three radical cations, a cyclization to 4-membered and 5-membered rings is plausible as shown for 3a˙+ in Fig. 1. In the case of 3a˙+, [iso-4a+ is formed reversibly given that less than 25 kcal mol−1 are required to open the 4-membered rings, while the formation of [4a+ is energetically favoured by approximately 55 kcal mol−1 suggesting that [4a+ is produced irreversibly at room temperature. While very similar reaction profiles are also found for 3a-Ph˙+ and 3b-Ph˙+, an entirely different situation was encountered during the optimization of 3b˙+, which was found to spontaneously cyclize to [iso-4b+. To further support this finding, a relaxed PES (potential energy surface) scan along both C–P vectors that are involved in a 5-endo-dig cyclization was carried out. The latter scan confirmed that [iso-4b+ is formed in a barrier-free manner, while no energetically reasonable trajectory interconnecting the starting point (approximate geometry of 3b˙+) and [4b+ was found, despite the fact that [4b+ is thermodynamically favoured over [iso-4b+ by approximately 45 kcal mol−1. For the thermal conversion of [iso-4b+ to [4b+, an activation barrier of 29.9 kcal mol−1 was calculated, which is prohibitively high for a reaction at room temperature.

The finding that [iso-4b+ is formed via a barrier-free pathway upon 1e-oxidation of 3b also inferred that a twofold oxidation of 3b (cf. mechanism B) is only possible if the second SET step is faster than intramolecular radical cyclization, which is unlikely for most 1e oxidants. For C2Cl6, however, it is known that image file: d5sc03752b-t1.tif, which is formed after the first SET, is more oxidizing than its parent reagent (C2Cl6),27 thus rendering two consecutive SETs possible. Hence, it is concluded that mechanism B is plausible for highly oxidizing chlorine-synthons (such as C2Cl6), while mechanism C is considered more likely for prototypical 1e oxidants (such as Fc[PF6]).

Despite the good agreement between mechanisms B and C with all our experimental findings, we were puzzled by the fact that mechanism A predicts the formation of P-diylidic diphosphapentalenes, not only for [4a]0, [4a-Ph]0 and [4b-Ph]0, but also for [4b]0. To elucidate whether [4b]0 may be prepared via mechanism A, [iso-4b]2+ was reduced to 3b as shown in Scheme 3.


image file: d5sc03752b-s3.tif
Scheme 3 Photochemical and thermal conversion of [iso-4b]2+ to [4b]2+ (top) together with the crystallographically determined molecular structures of [4b]0 and [4b]2+ (bottom, co-crystallized solvents and hydrogen atoms omitted, thermal ellipsoids drawn at the 50% probability level).

The finding that [iso-4b]0 was not detected during this reduction, is in line with the prediction that P-diylidic bis-(Δ2-phosphetenes) are thermodynamically disfavoured relative to the corresponding diphosphines (cf. mechanism A). Upon heating of 3b in the absence of C2Cl6, however, a fairly selective transformation to [4b]0 (δ(31P) = 32.3 ppm) set in at 60 °C, which is in line with mechanism A. In a subsequent step, the thus obtained P-diylide [4b]0 was oxidized using C2Cl6, which led to the originally envisioned dicationic phospholo[3,2-b]phosphole dication [4b]2+ (δ(31P) = 65.2 ppm). Alternatively, the latter product ([4b]2+) may as well be prepared directly from [iso-4b]2+via irradiation with violet light, while vast decomposition set in upon heating of [iso-4b]2+ to temperatures of ≥70 °C. Taken together, these findings strongly suggested that mechanism A is operative in the absence of an oxidant (as exploited for the thermal conversion of 3b to [4b]0), while one (for Fc[PF6]) or two (for C2Cl6) SET steps are needed for the oxidative cyclization of 3b to [iso-4b]2+ (see Scheme 2) at room temperature. Furthermore, we were able to show that the position of the alkyne on the naphthalene scaffold (connected to C-1 vs. C-2) plays a decisive role in the cyclization process. For the C-1 alkyne-linked naphthalene derivative 2d-Naph (see SI, Scheme S6), treatment with PhICl2 did not result in cyclization but instead led to chlorination at the phosphine moieties, indicating pronounced steric influences on this reaction.

With the knowledge that 4-exo-dig cyclizations are to be expected for annulation patterns akin to the one in 3b, we set out to further expand the π-system in [4a]2+ and [4a-Ph]2+ by targeting linearly fused anthraceno analogues.28 For this purpose, 2-bromo-3-iodo-anthracene (1c, see Scheme 4) was prepared in five steps starting from 5-bromo-6-iodo-phthalimide (see SI for details). With 1c at hand, 2c-Ph and 2c-Anth were obtained in analogy to 2a as shown in Scheme 4. For 2c-Naph, however, a different method had to be established given that multiple attempts to prepare the required stannylated 2-bromo-3-naphthyl acetylene met with failure. Thus, the envisioned dibromo naphthyl-anthracenyl acetylene was eventually assembled via a Julia–Lythgoe-type alkyne synthesis starting from 1d and an appropriately substituted sulfone (1e). With the dibromides 2c-Ph, 2c-Naph and 2c-Anth available, the target compounds [4c-Ph]2+, [4c-Naph]2+ and [4c-Anth]2+ were produced without difficulties (see Scheme 4) and isolated as red-orange powders.


image file: d5sc03752b-s4.tif
Scheme 4 Synthesis of [4c-Ph]2+, [4c-Naph]2+ and [4c-Anth]2+ together with the crystallographically determined molecular structures of [4c-Ph]2+ and [4c-Anth]2+ (co-crystallized solvent, disorder and hydrogen atoms omitted for clarity, thermal ellipsoids drawn at the 50% probability level).

In the molecular structure of [4c-Anth]2+, colinear π-stacking layers were found (see Scheme 4), which are also present in [4a]2+, but absent in all other phospholo[3,2-b]phosphole dications reported herein. This finding, however, seems to be related to crystal packing effects given that all compounds [4]2+ are nearly planar (with respect to the π-system) and therefore supposedly suited to engage in π-stacking interactions.

To gauge the extent of electron delocalization within the π-system of our new chromophores, AICD calculations (anisotropy of the induced current density)29 and NICSzz(1) XY-scans (nucleus-independent chemical shifts)30 were carried out (B3LYP, GD3, def2-TZVPP, SCRF for water). In all cases, diatropic ring currents indicative of aromaticity were found for the arenes on each side of the dicationic phospholo[3,2-b]phosphole core, while global ring currents involving the iPr2P+-bridges were clearly absent (see SI for details). Hence, the individual π-systems at both ends of each molecule are conjugated via the central dicationic core, but independent in terms of aromaticity. In [4c-Ph]2+, [4c-Naph]2+ and [4c-Anth]2+, for example, negative NICSzz(1) values were calculated for the b*, c* and d* rings, while positive NICSzz(1) values of +7(±2) ppm were found for the a and a* rings (see Fig. 2). These positive values for the a and a* rings, which are commonly seen as a sign for antiaromaticity,31 are misleading in the present case: In a geometrically constrained trans-stilbene, which was constructed by replacing the iPr2P+moieties in [H]2+ for hydrogen atoms, almost identical NICSzz(1) values were calculated, suggesting that all the π-extended derivatives of [H]2+ are best interpreted as diphosphonium-bridged ladder stilbenes.


image file: d5sc03752b-f2.tif
Fig. 2 NICSzz(1) XY-scans for [H]2+, [4c-Ph]2+, [4c-Naph]2+, [4c-Anth]2+and for a geometrically constrained trans-stilbene. The positive NICSzz(1) values for the a and a* are not indicative of antiaromaticity given that almost identical values were found for the latter trans-stilbene.

Due to the dicationic charge of all these compounds, energetically low-lying LUMOs are to be expected and indeed found in silico (see SI for details). A closer inspection revealed that the linearly annulated derivatives ([4a-Ph]2+, [4a]2+, [4c-Ph]2+, [4c-Naph]2+ and [4c-Anth]2+) exhibit LUMO energies in the narrow range of −3.48 ± 0.11 eV, while slightly lower energies of −3.70 ± 0.11 eV were calculated for the non-linear compounds ([4b-Ph]2+ and [4b]2+). For that reason, each series (linear vs. non-linear) is expected to display similar reduction potentials, which was confirmed by cyclic voltammetry (see SI for details). As expected, reduction of the non-linear derivatives (to their radical cations) was observed at less negative potentials (E1/2 = −0.71 ± 0.05 V vs. Fc/Fc+ for [4b-Ph]2+ and [4b]2+) in comparison to the corresponding reduction waves for the linear compounds ([4a-Ph]2+, [4a]2+, [4c-Ph]2+, [4c-Naph]2+ and [4c-Anth]2+ with E1/2 = −0.92 ± 0.10 V vs. Fc/Fc+). From exemplary measurements in an extended positive potential range, oxidation waves were observed for [H]Cl2, [4a-Ph]Cl2 and [4c-Ph]Cl2 (E1/2 = +0.47 ± 0.05 V, E1/2 = +0.51 ± 0.05 V and E1/2 = +0.49 ± 0.05 V vs. Fc/Fc+, respectively). Comparative measurements with triflate salts, however, unambiguously showed that these signals do not arise from oxidation of the π-conjugated ladder-stilbene framework, but rather from electrochemical oxidation of the chloride counterion (see SI, Fig. S153).

To further substantiate our findings, photophysical measurements and TD-DFT calculations (B3LYP, def2-TZVPP, GD3, solvation-corrected for H2O) were carried out. In all cases, the Cl salts, which are soluble in water in micromolar concentrations, were found to exhibit high molar absorption coefficients in the visible and ultraviolet spectral range (see Fig. 3). The lowest energy absorption in each compound is interpreted as a predominant π → π*-transition involving the HOMO and the LUMO, which is in line with our TD-DFT analysis (see SI for details). All new chromophores emit from an excited singlet state with fluorescence lifetimes (τ0) in the low nanosecond range (see Table 1 and SI).


image file: d5sc03752b-f3.tif
Fig. 3 Absorption and emission spectra of the chromophores as their chloride salts dissolved in pure water. The decay traces of the excited states are shown in the insets. Details for the measurements are provided in the SI. Strongly diluted dye solutions (lifetime measurements: c < 25 μM; emission spectra: c < 105 μM) were used for emission spectroscopy to avoid self-quenching or filter effects.
Table 1 Photophysical characteristics of the chromophores (see Fig. 3 for spectra, decay traces and structures) in aqueous solution. Details for the measurements are given in the SI
[H]2+a [4a-Ph]2+ [4a]2+ [4b-Ph]2+ [4b]2+ [4c-Ph]2+ [4c-Naph]2+ [4c-Anth]2+
a Ref. 21. b Fluorescein in 0.1 M NaOH as a reference compound.33 c Ruthenium-based complex ([Ru(bpy)3]2+) in H2O as a reference compound.33 d Relative quantum yields, see SI for details. e k FL = ϕFL/τ0. f k nr = (1 − ϕFL)/τ0.
ε/103 M−1 cm−1 (nm) 29.6 ± 1.0 (256) 41.6 ± 0.3 (265) 42.1 ± 1.1 (285) 20.9 ± 0.7 (270) 28.6 ± 0.0 (298) 37.9 ± 0.2 (269) 24.4 ± 0.3 (322) 53.2 ± 0.4 (336)
1.8 ± 0.0 (321) 8.6 ± 0.0 (305) 19.3 ± 0.5 (310) 7.3 ± 0.2 (408) 8.2 ± 0.2 (408) 10.1 ± 0.1 (389) 12.0 ± 0.1 (410) 17.2 ± 0.3 (418)
8.4 ± 0.0 (400) 11.8 ± 0.1 (426) 19.4 ± 0.6 (435) 4.9 ± 0.1 (450) 6.9 ± 0.1 (484) 9.6 ± 0.0 (489) 12.5 ± 0.2 (479) 22.3 ± 0.4 (498)
E 00/eV (nm) 2.81 (441) 2.63 (470) 2.61 (475) 2.46 (503) 2.30 (540) 2.19 (566) 2.23 (555) 2.23 (557)
ΔEcalc/eV (nm) 2.95 (420) 2.76 (449) 2.70 (459) 2.50 (495) 2.31 (537) 2.30 (539) 2.35 (527) 2.29 (542)
f calc 0.2167 0.3839 0.6105 0.1194 0.1858 0.3700 0.5390 0.8911
τ 0/ns (λmax/nm) 14.7 ± 0.09 (494) 9.1 ± 0.05 (520) 6.4 ± 0.05 (519) 14.1 ± 0.08 (583) 10.5 ± 0.00 (605) 1.3 ± 0.00 (670) 2.2 ± 0.04 (646) 2.6 ± 0.04 (631)
ϕ FL 0.87 ± 0.019 0.81 ± 0.020b 0.88 ± 0.000b 0.34 ± 0.005c 0.36 ± 0.026c 0.09 ± 0.010c 0.19 ± 0.008c 0.27 ± 0.014c
k FL /107 s−1 5.9 8.9 14 2.4 3.4 6.9 8.6 10
k nr /107 s−1 0.9 2.1 1.9 4.7 6.1 70 37 28


In the series [H]2+ → [4a]2+ → [4c-Anth]2+, the π-system is extended linearly and in a symmetric fashion, which is accompanied by a red-shift of the low-energy absorption bands and the fluorescence emission bands. It has also been demonstrated that the optical properties of ladder stilbenes can be further modified by incorporation of different heteroatoms into the scaffold (see Table S10 in the SI).22b,c,32 For related π-extended phospholium salts, similar bathochromic shifts have been reported recently.22b–d In our case, this trend is reflected in the energies of the 0-0 transitions E00 (see Table 1 and SI for details), which were determined from the intersections of the normalized absorption and emission spectra. The calculated energies for the S0 → S1 excitation (see Table 1) were found to be in good agreement with the E00 values, thus confirming this trend. In [4a-Ph]2+ and [4a]2+, almost identical E00 energies were determined experimentally (ΔE00 = 0.02 eV), but the molar absorption coefficient at the low-energy absorption band is higher in [4a]2+ by a factor of ∼1.5. The calculated oscillator strengths for the lowest energy absorptions in [4a-Ph]2+ and [4a]2+ are in line with this finding. Furthermore, a decrease in the excited-state lifetimes (increase in kFL) was observed in the series [H]2+ → [4a-Ph]2+ → [4a]2+, which is consistent with the Strickler–Berg equation in the simplified version (constant of radiative decay is proportional to the energy of the lowest transition squared and the corresponding oscillator strength, kradν2f).34 Within the series of anthraceno-fused compounds ([4c-Ph]2+, [4c-Naph]2+ and [4c-Anth]2+) similar trends were evident: Compared to [H]2+ and [4a]2+, the E00 energies are generally lower in all the anthraceno-fused compounds. The molar absorption coefficients for the lowest energy transition were found to increase within the series [4c-Ph]2+ → [4c-Naph]2+ → [4c-Anth]2+, while the E00 energy remains essentially identical (ΔE00 = 0.04 eV). The lifetimes increase when going from [4c-Ph]2+ to [4c-Anth]2+, which is opposite to the trend observed for the linearly naphtho-fused compounds ([4a-Ph]2+ → [4a]2+). This observation might be related to the degeneracies of the orbitals involved in the S1→S0 transition: In [4c-Anth]2+, two anthracene units are fused to the central dicationic core, whereas the symmetry for [4c-Naph]2+ and [4c-Ph]2+ is lower. As a result, the degeneracies of the orbitals involved in the transition may differ, which eventually affects kFL and therefore τ0 as well. Moreover, strongly divergent intersystem crossing rates could contribute to such τ0 differences of seemingly similar compounds. In general, red emitters show significantly lower emission quantum yields compared to their congeners emitting at higher energies as a result of the energy gap law.35 In consequence, higher rate constants for the non-radiative decay (knr, see Table 1) are commonly observed for red emitters. A quantum yield ϕFL close to 30% for [4c-Anth]2+, whose broad emission band is centred around 630 nm, is therefore highly appreciated, in particular in aqueous solution. For the non-symmetric red emitters, a closer look at the frontier orbitals indicates the admixture of charge transfer contributions in addition to pronounced π → π* transitions. Hence, we carried out additional spectroscopic measurements with [4c-Ph]2+, [4c-Naph]2+ and [4c-Anth]2+ for investigating solvatochromic effects (see Fig. S74–S76 and Tables S5–S7 in the SI). The absorption and emission spectra and hence the derived parameters (E00 and Stokes shift) only change to a small extent (<10%) when going from water via polar organic solvents to DCM, substantiating the absence of significant charge transfer characters for the first excited singlet states. The non-linear naphtho-fused chromophores [4b-Ph]2+ and [4b]2+ differ from their linear congeners, not only in terms of their redox potentials (vide supra), but also in their photophysical characteristics: For [4b-Ph]2+ and [4b]2+, comparatively low molar absorption coefficients and red-shifted absorption and emission bands are found. While our TD-DFT analysis correctly predicted these findings (cf. lower oscillator strengths for [4b-Ph]2+ and [4b]2+, see Table 1), the origin of this effect is not obvious, although it is well-known that the electronic transitions of linear and non-linear π-systems may differ significantly: For π-extended pentalenes, for example, the HOMO–LUMO transitions are allowed in linear structures, but symmetry-forbidden (and therefore less intense) in non-linear derivatives.36 For all the diphosphonium bridged ladder stilbenes presented herein, however, the HOMO–LUMO transitions are symmetry-allowed, regardless of the annulation pattern (see SI for details). In the non-linear derivatives, however, significantly lower transition electric dipole moments were calculated (4.9 debye for [4b-Ph]2+ and 8.3 debye for [4b]2+) in comparison to their linear counterparts (14.4 debye for [4a-Ph]2+ and 23.4 debye for [4a]2+, see SI for details). In consequence, lower molar absorption coefficients are to be expected for the non-linear chromophores. Finally, the lifetimes of the non-linear derivatives are longer than those of the linear compounds, which is in line with the Strickler–Berg equation, taking the lower E00 energies and molar absorption coefficients of the non-linear derivatives into account.34

Conclusions

In summary, several π-extended phospholo[3,2-b]phosphole dications with different naphtho and anthraceno ring-fusion patterns have been prepared in order to establish a correlation between their structures and their optoelectronic properties. It was found that the annulation pattern (linear vs. non-linear) is of crucial importance, not only from a photophysical perspective, but also from a synthetic point of view: For the specific naphtho-annulation pattern present in 3b, a dicationic bis-(Δ2-phosphetene) ([iso-4b]2+) and its phospholo[3,2-b]phosphole isomer ([4b]2+) have been prepared, each in a selective manner. On this basis, three distinct mechanistic pathways for the cyclization of diphosphinotolanes have been explored in silico, which revealed that the formation of uncharged diphosphapentalenes (such as [4b]0) via mechanism A plays a role in the absence of oxidants, while SET steps play a crucial role in the presence of suitable oxidants (mechanism B for C2Cl6 or mechanism C for Fc[PF6]). While these findings are expected to guide the way to new phospholo[3,2-b]phosphole dications, it was also shown that all our new chromophores are best interpreted as diphosphonium-bridged ladder stilbenes (no antiaromaticity within the phospholium rings). Detailed photophysical studies in water revealed that both linearly fused centrosymmetric compounds ([4a]2+ and [4c-Anth]2+) display fairly high fluorescence quantum yields, while π-stacking interactions were found in their crystalline samples. Hence, [4a]2+ and [4c-Anth]2+ seem to be particularly suited for sensing applications (e.g. intercalation into DNA), which will be a subject of future studies.

Author contributions

S. S. and J.-M. M. conducted the synthetic work and M.-S. B. conducted the photophysical measurements. J. B. carried out the DFT work and the crystallographic analysis. C. K. and J. B. supervised the project. All authors contributed to writing and manuscript preparation.

Conflicts of interest

There are no conflicts to declare.

Data availability

CCDC 2431371–2431379 and 2470828 contain the supplementary crystallographic data for this paper.37a–j

The data supporting this article have been included as part of the supplementary information (SI). Supplementary information: experimental details, computational data, details on photophysical measurements, electrochemical measurements, crystallographic data. See DOI: https://doi.org/10.1039/d5sc03752b.

Acknowledgements

The authors acknowledge support by the state of Baden-Württemberg through bwHPC and the DFG through grant INST 40/467-1 FUGG (JUSTUS cluster). We thank Malte Hanne, Christian Delavier, Tim Diederich, Jessica Schaaf, Karolin Becker and Teresa van Ghemen for their help with preparing some of the starting materials during research internships. S. S. thanks the Landesgraduiertenförderung (LGF) for a PhD fellowship. S. S. and J.-M. M. gratefully acknowledge the support of Prof. L. H. Gade following the passing of Prof. J. Ballmann.

Notes and references

  1. (a) H. Lauwick, M. P. Duffy, P.-A. Bouit and M. Hissler, Coord. Chem. Rev., 2021, 433, 213759 CrossRef CAS; (b) H. Lauwick, E. Kertész, K. N. Garami, W. Huadsai, M. P. Duffy, R. Foundi, A. Chemin, T. Roisnel, N. Vanthuyne, Z. Benkő, P.-A. Bouit and M. Hissler, Angew. Chem., Int. Ed., 2024, 63, e202409988 CAS; (c) T. Xin and C. C. Cummins, J. Am. Chem. Soc., 2023, 145, 25989–25994 CrossRef CAS PubMed; (d) U. Heim, H. Pritzkow, U. Fleischer, H. Grützmacher, M. Sanchez, R. Réau and G. Bertrand, Chem.–Eur. J., 1996, 2, 68–74 CrossRef CAS; (e) H. Chen, S. Pascal, Z. Wang, P.-A. Bouit, Z. Wang, Y. Zhang, D. Tondelier, B. Geffroy, R. Réau, F. Mathey, Z. Duan and M. Hissler, Chem.–Eur. J., 2014, 20, 9784–9793 CrossRef CAS PubMed.
  2. (a) F. Mathey, Chem. Rev., 1988, 88, 429–453 CrossRef CAS; (b) T. Baumgartner and R. Réau, Chem. Rev., 2006, 106, 4681–4727 CrossRef CAS PubMed; (c) Y. Matano, A. Saito, T. Fukushima, Y. Tokudome, F. Suzuki, D. Sakamaki, H. Kaji, A. Ito, K. Tanaka and H. Imahori, Angew. Chem., Int. Ed., 2011, 50, 8016–8020 CrossRef CAS PubMed; (d) P.-A. Bouit, A. Escande, R. Szűcs, D. Szieberth, C. Lescop, L. Nyulászi, M. Hissler and R. Réau, J. Am. Chem. Soc., 2012, 134, 6524–6527 CrossRef CAS PubMed; (e) M. P. Duffy, W. Delaunay, P. A. Bouit and M. Hissler, Chem. Soc. Rev., 2016, 45, 5296–5310 RSC; (f) P. Hibner-Kulicka, J. A. Joule, J. Skalik and P. Bałczewski, RSC Adv., 2017, 7, 9194–9236 RSC; (g) C. Wang, M. Taki, Y. Sato, A. Fukazawa, T. Higashiyama and S. Yamaguchi, J. Am. Chem. Soc., 2017, 139, 10374–10381 CrossRef CAS PubMed; (h) T. Johannsen, C. Golz and M. Alcarazo, Angew. Chem., Int. Ed., 2020, 59, 22779–22784 CrossRef CAS PubMed; (i) Q. Wu, M. Taki, Y. Tanaka, M. Kesherwani, Q. M. Phung, S. Enoki, Y. Okada, F. Tama and S. Yamaguchi, Angew. Chem., Int. Ed., 2024, 63, e202400711 CrossRef CAS PubMed.
  3. (a) T. Delouche, A. Vacher, E. Caytan, T. Roisnel, B. Le Guennic, D. Jacquemin, M. Hissler and P.-A. Bouit, Chem.–Eur. J., 2020, 26, 8226–8229 CrossRef CAS PubMed; (b) M. Grzybowski, M. Taki, K. Senda, Y. Sato, T. Ariyoshi, Y. Okada, R. Kawakami, T. Imamura and S. Yamaguchi, Angew. Chem., Int. Ed., 2018, 57, 10137–10141 CrossRef CAS PubMed; (c) N. Hashimoto, R. Umano, Y. Ochi, K. Shimahara, J. Nakamura, S. Mori, H. Ohta, Y. Watanabe and M. Hayashi, J. Am. Chem. Soc., 2018, 140, 2046–2049 CrossRef CAS PubMed; (d) P. Hindenberg, M. Busch, A. Paul, M. Bernhardt, P. Gemessy, F. Rominger and C. Romero-Nieto, Angew. Chem., Int. Ed., 2018, 57, 15157–15161 CrossRef CAS PubMed; (e) G. Pfeifer, F. Chahdoura, M. Papke, M. Weber, R. Szűcs, B. Geffroy, D. Tondelier, L. Nyulászi, M. Hissler and C. Müller, Chem.–Eur. J., 2020, 26, 10534–10543 CrossRef CAS PubMed; (f) T. Sangchai, N. Ledos, A. Vacher, M. Cordier, B. Le Guennic, M. Hissler, D. Jacquemin and P.-A. Bouit, Chem.–Eur. J., 2023, 29, e202301165 CrossRef CAS PubMed; (g) R. Sugiyama, R. Okada, T. Noda, N. Meguro, N. Yoshida, K. Hoshi, H. Ohta, M. Hayashi, H. Sasabe and J. Kido, Chem.–Eur. J., 2024, 30, e202304328 CrossRef CAS PubMed; (h) R. Szűcs, P.-A. Bouit, L. Nyulászi and M. Hissler, ChemPhysChem, 2017, 18, 2618–2630 CrossRef PubMed.
  4. (a) K. Andoh, M. Murai, P.-A. Bouit, M. Hissler and S. Yamaguchi, Angew. Chem., Int. Ed., 2024, 63, e202410204 CrossRef CAS PubMed; (b) T. Delouche, R. Mokrai, T. Roisnel, D. Tondelier, B. Geffroy, L. Nyulászi, Z. Benkő, M. Hissler and P.-A. Bouit, Chem.–Eur. J., 2020, 26, 1856–1863 CrossRef CAS PubMed; (c) C. Li, W. Zhou, Z. Liu, R. Gao, Q. Mi, Z. Ning and Y. Ren, Chem. Sci., 2024, 15, 18608–18616 RSC; (d) G. Märkl and W. Burger, Angew. Chem., Int. Ed., 1984, 23, 894–895 CrossRef; (e) K. Padberg, J. D. R. Ascherl, F. Hampel and M. Kivala, Chem.–Eur. J., 2020, 26, 3474–3478 CrossRef CAS PubMed; (f) W. Winter, Chem. Ber., 1976, 109, 2405–2419 CrossRef CAS; (g) F.-P. Zhang, R.-H. Wang, J.-F. Li, H. Chen, M. Hari Babu and M. Ye, Angew. Chem., Int. Ed., 2023, 62, e202314701 CrossRef CAS PubMed; (h) X. He, J. Borau-Garcia, A. Y. Y. Woo, S. Trudel and T. Baumgartner, J. Am. Chem. Soc., 2013, 135, 1137–1147 CrossRef CAS PubMed.
  5. (a) N. Asok, J. R. Gaffen and T. Baumgartner, Acc. Chem. Res., 2023, 56, 536–547 CrossRef CAS PubMed; (b) T. Baumgartner, Acc. Chem. Res., 2014, 47, 1613–1622 CrossRef CAS PubMed; (c) D. Joly, P. A. Bouit and M. Hissler, J. Mater. Chem. C, 2016, 4, 3686–3698 RSC; (d) M. A. Shameem and A. Orthaber, Chem.–Eur. J., 2016, 22, 10718–10735 CrossRef CAS PubMed; (e) K. Zhang, X. Wang, Z. Zhou, J. Guo, H. Liu, Y. Zhai, Y. Yao, K. Yang and Z. Zeng, Angew. Chem., Int. Ed., 2025, e202418520 CAS.
  6. (a) C. Wang, A. Fukazawa, M. Taki, Y. Sato, T. Higashiyama and S. Yamaguchi, Angew. Chem., Int. Ed., 2015, 54, 15213–15217 CrossRef CAS PubMed; (b) J. Wang, M. Taki, Y. Ohba, M. Arita and S. Yamaguchi, Angew. Chem., Int. Ed., 2024, 63, e202404328 CrossRef CAS PubMed; (c) E. Yamaguchi, C. Wang, A. Fukazawa, M. Taki, Y. Sato, T. Sasaki, M. Ueda, N. Sasaki, T. Higashiyama and S. Yamaguchi, Angew. Chem., Int. Ed., 2015, 54, 4539–4543 CrossRef CAS PubMed; (d) E. Regulska and C. Romero-Nieto, Mater. Today Chem., 2021, 22, 100604 CrossRef CAS; (e) M. Schenk, N. König, E. Hey-Hawkins and A. G. Beck-Sickinger, ChemBioChem, 2024, 25, e202300857 CrossRef CAS PubMed.
  7. (a) X. Zhang, Y. Jiang, Q. Ma, S. Hu and S. Liao, J. Am. Chem. Soc., 2021, 143, 6357–6362 CrossRef CAS PubMed; (b) Z. Yang, J. Chen and S. Liao, ACS Macro Lett., 2022, 11, 1073–1078 CrossRef CAS PubMed; (c) J. Ding, S. Luo, Y. Xu, Q. An, Y. Yang and Z. Zuo, Chem. Commun., 2023, 59, 4055–4058 RSC; (d) Y. Xu, W. Chen, R. Pu, J. Ding, Q. An, Y. Yang, W. Liu and Z. Zuo, Nat. Commun., 2024, 15, 9366 CrossRef CAS PubMed; (e) Z. Yang, Y. Liao, Z. Zhang, J. Chen, X. Zhang and S. Liao, J. Am. Chem. Soc., 2024, 146, 6449–6455 CrossRef CAS PubMed.
  8. (a) Y. Ren and T. Baumgartner, Dalton Trans., 2012, 41, 7792–7800 RSC; (b) D. Thakur, Sushmita, S. A. Meena and A. K. Verma, Chem. Rec., 2024, 24, e202400058 CrossRef CAS PubMed.
  9. Note that parent phosphole oxides are unstable, see: Z. Mucsi and G. Keglevich, Eur. J. Org. Chem., 2007, 28, 4765–4771 CrossRef.
  10. (a) O. Larrañaga, C. Romero-Nieto and A. de Cózar, Chem.–Eur. J., 2019, 25, 9035–9044 CrossRef PubMed; (b) A. de Cózar and C. Romero-Nieto, Inorg. Chem., 2023, 62, 4097–4105 CrossRef PubMed.
  11. (a) A. Belyaev, P.-T. Chou and I. O. Koshevoy, Chem.–Eur. J., 2021, 27, 537–552 CrossRef CAS PubMed; (b) S. Yamaguchi, A. Fukazawa and M. Taki, J. Synth. Org. Chem., Jpn., 2017, 75, 1179–1187 CrossRef CAS; (c) Y. Matano and H. Imahori, Org. Biomol. Chem., 2009, 7, 1258–1271 RSC.
  12. (a) S. Nieto, P. Metola, V. M. Lynch and E. V. Anslyn, Organometallics, 2008, 27, 3608–3610 CrossRef CAS; (b) Y. Kuninobu, T. Yoshida and K. Takai, J. Org. Chem., 2011, 76, 7370–7376 CrossRef CAS PubMed; (c) A. Belyaev, Y.-T. Chen, S.-H. Su, Y.-J. Tseng, A. J. Karttunen, S. P. Tunik, P.-T. Chou and I. O. Koshevoy, Chem. Commun., 2017, 53, 10954–10957 RSC; (d) Q. Ge, J. Zong, B. Li and B. Wang, Org. Lett., 2017, 19, 6670–6673 CrossRef CAS PubMed; (e) S. Nilewar, A. Jayaraman and B. T. Sterenberg, Organometallics, 2018, 37, 4699–4710 CrossRef CAS; (f) K. Nishimura, Y. Unoh, K. Hirano and M. Miura, Chem.–Eur. J., 2018, 24, 13089–13092 CrossRef CAS PubMed; (g) Z. Wang, N. Asok, J. Gaffen, Y. Gottlieb, W. Bi, C. Gendy, R. Dobrovetsky and T. Baumgartner, Chem, 2018, 4, 2628–2643 CrossRef CAS; (h) J. Ma, L. Wang and Z. Duan, Org. Lett., 2022, 24, 1550–1555 CrossRef CAS PubMed; (i) L. Ma, S. Mallet-Ladeira, J. Monot, B. Martin-Vaca and D. Bourissou, Org. Chem. Front., 2024, 11, 3376–3383 RSC; (j) N. Saha and B. König, ACS Catal., 2024, 14, 17958–17971 CrossRef CAS; (k) J. Guo, C. Mao, B. Deng, L. Ye, Y. Yin, Y. Gao and S. Tu, J. Org. Chem., 2020, 85, 6359–6371 CrossRef CAS PubMed; (l) H.-L. Ni, Z. Wen, H.-Z. Xue and L. Chen, Adv. Synth. Catal., 2024, 366, 4323–4346 CrossRef CAS; (m) Z. Mei, C. Yuan, C. Chen, M. You, H. Huang and Q. Xiao, Org. Lett., 2025, 27, 3905–3910 CrossRef CAS PubMed.
  13. A. Fukazawa, H. Yamada and S. Yamaguchi, Angew. Chem., Int. Ed., 2008, 47, 5582–5585 CrossRef CAS PubMed.
  14. Y. Ren and T. Baumgartner, J. Am. Chem. Soc., 2011, 133, 1328–1340 CrossRef CAS PubMed.
  15. T. Delouche, E. Caytan, M. Cordier, T. Roisnel, G. Taupier, Y. Molard, N. Vanthuyne, B. Le Guennic, M. Hissler, D. Jacquemin and P.-A. Bouit, Angew. Chem., Int. Ed., 2022, 61, e202205548 CrossRef CAS PubMed.
  16. S. K. Banjare, L. Lezius, E. S. Horst, D. Leifert, C. G. Daniliuc, F. A. Alasmary and A. Studer, Angew. Chem., Int. Ed., 2024, 63, e202404275 CrossRef CAS PubMed.
  17. J. Silberzahn, H. Pritzkow and H. P. Latscha, Angew. Chem., Int. Ed., 1990, 29, 799 CrossRef.
  18. A. Fukazawa, M. Hara, T. Okamoto, E.-C. Son, C. Xu, K. Tamao and S. Yamaguchi, Org. Lett., 2008, 10, 913–916 CrossRef CAS PubMed.
  19. Compounds containing two phosphoranly units are commonly obtained as isomeric mixtures and separated later on by column chromatography; see for example: (a) A. Fukazawa, Y. Ichihashi, Y. Kosaka and S. Yamaguchi, Chem.–Asian J., 2009, 4, 1729–1740 CrossRef CAS PubMed; (b) Y. Hayashi, Y. Matano, K. Suda, Y. Kimura, Y. Nakao and H. Imahori, Chem.–Eur. J., 2012, 18, 15972–15983 CrossRef CAS PubMed; (c) P. Hindenberg, A. Belyaev, F. Rominger, I. O. Koshevoy and C. Romero-Nieto, Org. Lett., 2022, 24, 6391–6396 CrossRef CAS PubMed; (d) E. Regulska, P. Hindenberg, A. Espineira-Gutierrez and C. Romero-Nieto, Chem.–Eur. J., 2023, 29, e202202769 CrossRef CAS PubMed.
  20. C. Li, H. Huang, L. Sun, M. Huang, H. Ding, J. Bai, B.-P. Cao and Q. Xiao, Angew. Chem., Int. Ed., 2023, 62, e202215436 CrossRef CAS PubMed.
  21. M.-S. Bertrams, K. Hermainski, J.-M. Mörsdorf, J. Ballmann and C. Kerzig, Chem. Sci., 2023, 14, 8583–8591 RSC.
  22. (a) P. Federmann, H. K. Wagner, P. W. Antoni, J.-M. Mörsdorf, J. L. Pérez Lustres, H. Wadepohl, M. Motzkus and J. Ballmann, Org. Lett., 2019, 21, 2033–2038 CrossRef CAS PubMed; (b) K. Górski, Ł. W. Ciszewski, A. Wrzosek, A. Szewczyk, A. L. Sobolewski and D. T. Gryko, ChemRxiv, 2025 DOI:10.26434/chemrxiv-2025-whg34; (c) K. Górski, Ł. W. Ciszewski, A. Wrzosek, A. Szewczyk, A. L. Sobolewski and D. T. Gryko, Org. Chem. Front., 2025 10.1039/D5QO00708A , Advance Article; (d) A. Belyaev, Y.-T. Chen, Z.-Y. Liu, P. Hindenberg, C.-H. Wu, P.-T. Chou, C. Romero-Nieto and I. O. Koshevoy, Chem.–Eur. J., 2019, 25, 6332–6341 CrossRef CAS PubMed; (e) S. Arndt, M. M. Hansmann, F. Rominger, M. Rudolph and A. S. K. Hashmi, Chem.–Eur. J., 2017, 23, 5429–5433 CrossRef CAS PubMed.
  23. (a) J. C. Tendyck, H. Klöcker, L. Schürmann, E.-U. Würthwein, A. Hepp, M. Layh and W. Uhl, J. Org. Chem., 2020, 85, 14315–14332 CrossRef CAS PubMed; (b) F. Roesler, M. Kovács, C. Bruhn, Z. Kelemen and R. Pietschnig, Chem.–Eur. J., 2021, 27, 9782–9790 CrossRef CAS PubMed.
  24. (a) F. Neese, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 73–78 CAS; (b) E. Caldeweyher, J.-M. Mewes, S. Ehlert and S. Grimme, Phys. Chem. Chem. Phys., 2020, 22, 8499–8512 RSC; (c) M. Bursch, J.-M. Mewes, A. Hansen and S. Grimme, Angew. Chem., Int. Ed., 2022, 61, e202205735 CrossRef CAS PubMed; (d) T. Gasevic, J. B. Stückrath, S. Grimme and M. Bursch, J. Phys. Chem. A, 2022, 126, 3826–3838 CrossRef CAS PubMed; (e) F. Neese, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2022, 12, e1606 Search PubMed.
  25. Note that the proposed carbene intermediates have been trapped previously via complexation with numerous transition metals, see: (a) H. K. Wagner, N. Ansmann, T. Gentner, H. Wadepohl and J. Ballmann, Organometallics, 2021, 40, 804–812 CrossRef CAS; (b) H. K. Wagner, H. Wadepohl and J. Ballmann, Chem. Sci., 2021, 12, 3693–3701 RSC; (c) B. Rudin, L. Eberle and J. Ballmann, Organometallics, 2023, 42, 933–943 CrossRef CAS.
  26. X. Zhao, Z. Gan, C. Hu, Z. Duan and F. Mathey, Org. Lett., 2017, 19, 5814–5817 CrossRef CAS PubMed.
  27. (a) A. L. Roberts and P. M. Gschwend, Environ. Sci. Technol., 1991, 25, 76–86 CrossRef CAS; (b) E. V. Patterson, C. J. Cramer and D. G. Truhlar, J. Am. Chem. Soc., 2001, 123, 2025–2031 CrossRef CAS PubMed; (c) C. Costentin, M. Robert and J.-M. Savéant, J. Am. Chem. Soc., 2003, 125, 10729–10739 CrossRef CAS PubMed.
  28. Anthraceno-fused mono-phospholes have been reported, see: Y. Kudoh, K. Fujii, Y. Kimura, M. Minoura and Y. Matano, J. Org. Chem., 2022, 87, 10493–10500 CrossRef CAS PubMed.
  29. (a) R. Herges and D. Geuenich, J. Phys. Chem. A, 2001, 105, 3214–3220 CrossRef CAS; (b) D. Geuenich, K. Hess, F. Köhler and R. Herges, Chem. Rev., 2005, 105, 3758–3772 CrossRef CAS PubMed.
  30. (a) Z. Chen, C. S. Wannere, C. Corminboeuf, R. Puchta and P. v. R. Schleyer, Chem. Rev., 2005, 105, 3842–3888 CrossRef CAS PubMed; (b) R. Gershoni-Poranne and A. Stanger, Chem.–Eur. J., 2014, 20, 5673–5688 CrossRef CAS PubMed; (c) Z. Wang, Chemistry, 2024, 6, 1692–1703 CrossRef CAS.
  31. It is known that NICS values may be misleading and/or difficult to interpret; see for example: J. Poater, M. Solà, R. G. Viglione and R. Zanasi, J. Org. Chem., 2004, 69, 7537–7542 CrossRef CAS PubMed.
  32. H. Haiyang, L. Sun, Z. Mei, J. Bai, C. Chen, M. You, Y. Wu, H. Ding and Q. Xiao, ChemRxiv, 2024 DOI:10.26434/chemrxiv-2024-dlx0q.
  33. K. Suzuki, A. Kobayashi, S. Kaneko, K. Takehira, T. Yoshihara, H. Ishida, Y. Shiina, S. Oishi and S. Tobita, Phys. Chem. Chem. Phys., 2009, 11, 9850–9860 RSC.
  34. (a) S. J. Strickler and R. A. Berg, J. Chem. Phys., 1962, 37, 814–822 CrossRef CAS; (b) J. Mohanty and W. M. Nau, Photochem. Photobiol. Sci., 2004, 3, 1026–1031 CrossRef CAS PubMed; (c) M. Montalti, A. Credi, L. Prodi and M. T. Gandolfi, Handbook of Photochemistry, CRC Press, 2006 CrossRef; (d) M.-S. Bertrams and C. Kerzig, Chem. Commun., 2021, 57, 6752–6755 RSC.
  35. (a) J. R. Lakowicz, Principles of Fluorescence Spectroscopy, Springer, Boston, MA, 2006 CrossRef; (b) P. Klán and J. Wirz, Photochemistry of organic compounds: from concepts to practice, John Wiley & Sons, 2009 CrossRef; (c) C. Fischer and C. Sparr, Angew. Chem., Int. Ed., 2018, 57, 2436–2440 CrossRef CAS PubMed; (d) J. Maillard, K. Klehs, C. Rumble, E. Vauthey, M. Heilemann and A. Fürstenberg, Chem. Sci., 2021, 12, 1352–1362 RSC.
  36. (a) T. Kawase, T. Fujiwara, C. Kitamura, A. Konishi, Y. Hirao, K. Matsumoto, H. Kurata, T. Kubo, S. Shinamura, H. Mori, E. Miyazaki and K. Takimiya, Angew. Chem., Int. Ed., 2010, 49, 7728–7732 CrossRef CAS; (b) G. London, M. von Wantoch Rekowski, O. Dumele, W. B. Schweizer, J.-P. Gisselbrecht, C. Boudon and F. Diederich, Chem. Sci., 2014, 5, 965–972 RSC; (c) T. Kawase and J.-i. Nishida, Chem. Rec., 2015, 15, 1045–1059 CrossRef PubMed; (d) P. J. Mayer and G. London, Org. Lett., 2023, 25, 42–46 CrossRef CAS PubMed; (e) J. Shen, D. Yuan, Y. Qiao, X. Shen, Z. Zhang, Y. Zhong, Y. Yi and X. Zhu, Org. Lett., 2014, 16, 4924–4927 CrossRef CAS PubMed.
  37. (a) CCDC 2431371: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mm1bn; (b) CCDC 2431372: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mm1cp; (c) CCDC 2431373: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mm1dq; (d) CCDC 2431374: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mm1fr; (e) CCDC 2431375: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mm1gs; (f) CCDC 2431376: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mm1ht; (g) CCDC 2431377: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mm1jv; (h) CCDC 2431378: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mm1kw; (i) CCDC 2431379: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2mm1lx; (j) CCDC 2470828: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ny34w.

Footnotes

Equal contribution.
J. B. deceased on June 7, 2025.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.