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

Dual fluorescent phenanthridinones and crinasiadine derivatives by consecutive palladium-catalyzed three-component syntheses

Regina Kohlbecher and Thomas J. J. Müller*
Heinrich Heine University Duesseldorf, Faculty of Mathematics and Natural Sciences, Institute of Organic Chemistry and Macromolecular Chemistry, Universitätstrasse 1, Düsseldorf D-40225, Germany. E-mail: ThomasJJ.Mueller@hhu.de

Received 13th September 2025 , Accepted 4th December 2025

First published on 10th December 2025


Abstract

The sequential concatenation of Buchwald–Hartwig amination, Suzuki coupling, and lactamization in a consecutive palladium-catalyzed three-component synthesis provides direct access to functionalized N-arylsubstituted phenanthridinones and alkaloid-analogous crinasiadines starting from simple, readily available ortho-bromoanilines. A modified reaction sequence, consisting of Suzuki coupling, amide bond formation (lactamization), and subsequent alkylation at the amide nitrogen, provides N-alkylsubstituted phenanthridinones and crinasiadines, including two natural products, which are known for their cytotoxic activity against various cancer cell lines. A comprehensive investigation of the photophysical properties reveals dual emission of the N-aryl substituted derivatives, characterized by locally excited states (LE band) and intramolecular charge transfer states (CT band). Quantum chemical calculations rationalize the dual emission and suggest that the LE band derives from the pseudo-N-intra conformation, whereas the CT band arises from the pseudo-N-extra conformation in the excited state.


Introduction

Phenanthridinones are major heterocyclic building blocks that occur in a variety of complex natural products and pharmaceuticals, exhibiting a broad spectrum of pharmacological activities.1,2 The significance of their scaffolds has aroused great interest in synthetic and medicinal chemistry to design new methodologies and novel derivatives of pharmaceutical relevance.3

The access to phenanthridinones is facilitated through a wide range of methods, encompassing both traditional and innovative strategies.3 Gräbe and Wander first synthesized 6(5H)-phenanthridinone in 1893, starting from 2′-aminobiphenyl-2-carboxylic acid.4 Later, numerous additional methods, such as the alkaline hydroxylation of phenanthridine,5 Beckmann rearrangement,6 or anomalous Schmidt reaction,7 were reported. However, these classical methods have limitations as they often require multistep syntheses of the starting materials and generally result in low to moderate overall yields. Therefore, the quest for methods using environmentally benign chemicals, energy efficient microwave-assisted reactions, and catalytic processes has become increasingly interesting.3 In 2013, Tanimoto et al. established an efficient one-step synthesis of N-unsubstituted phenanthridinones, including the natural product phenaglydon.8 This method implements 2-halobenzoates and 2-aminophenylboronic acids in a Pd-catalyzed Suzuki coupling reaction.3,8 The phenanthridinone scaffold is a frequently occurring structural motif in bioactive plant alkaloids.3,9 For instance, the phenanthridinone-based alkaloid oxynitidine was isolated from the methanol extracts of the root bark and wood of the climbing plant Zanthoxylum nitidum (Fig. 1, left).10–12 These dried plant extracts are known for their anti-inflammatory, analgesic, antimicrobial, and anticancer properties in traditional Chinese medicine.11,13 The scaffold is also found in various Amaryllidaceae alkaloids, which exhibit antiproliferative effects on cancer cells.14–18 For instance, three crinasiadines (Fig. 1) were isolated from the Amaryllidaceae plant Zephyranthes candida, where N-phenethylcrinasiadine exhibits the strongest cytotoxicity against leukemia, lung, and colon cancer cell lines.19


image file: d5ra06934c-f1.tif
Fig. 1 Phenanthridinone-based natural products.

Recent research highlights the broad pharmacological potential of phenanthridinones, including anticancer activity,20 PARP1 inhibition,21,22 cardioprotective effects,23,24 regulation of cholesterol metabolism,25–28 topoisomerase I inhibition,3,29 and estrogen receptor modulation.30

Phenanthridinone derivatives are also gaining increasing interest in the field of organic electronics.31–33 In particular, the photophysical properties of novel phenanthridinones have only scarcely been recognized so far, as they can provide valuable information on SAR, especially since some are known for their dual emission behavior.34,35

Multicomponent reactions (MCRs) of functional chromophores enable rapid and diversity-oriented syntheses of compound libraries.36–38 By definition, MCRs allow for the synthesis of target structures from three or more reactants in a single reaction vessel, ensuring that significant parts of the employed starting materials are incorporated into the final product while drastically reducing the operational effort required for isolation.36–41

Recently, the concatenation of Suzuki coupling and Buchwald–Hartwig amination in a consecutive sequentially palladium-catalyzed MCR has successfully paved avenues to libraries of functionalized phenothiazines, carbazoles, indoles, and triarylamines.42 Herein, we report a novel, diversified, and sequentially palladium-catalyzed MCR that allows direct access to functionalized phenanthridinones as well as alkaloid-analogous crinasiadines starting from simple, readily available starting materials. Additionally, investigations of the photophysical properties and quantum chemical calculations open up a deeper understanding of the electronic structure of phenanthridinone derivatives.

Results and discussion

Syntheses

Starting from ortho-bromoaniline 1, a palladium-catalyzed three-component synthesis was envisioned to provide direct access to phenanthridinones and crinasiadines. After comprehensive optimization (for details see SI, chpt. 5.1), a novel Buchwald-Hartwig amination-Suzuki coupling-lactamization sequence (BHSL-sequence) was developed, based on a previously optimized, highly diversified, sequentially palladium-catalyzed one-pot process for the synthesis of para-biaryl-substituted triarylamines (p-bTAA).43 This approach aims to construct a substance library of diverse phenanthridinone derivatives in a modular manner. By switching the Buchwald-Hartwig amination and the Suzuki coupling compared to the p-bTAA synthesis,43 the formation of poorly soluble 6(5H)-phenanthridinone intermediates could be circumvented, enabling the successful synthesis of the target molecules. Due to their reactivity, the first step of the Buchwald–Hartwig amination requires an aryl iodide 2 and an ortho-bromoaniline 1 to exclude homocoupling of the aniline.44 Additionally, the aryl iodide should only be employed in one equivalent to avoid twofold arylation of the aniline nitrogen. Furthermore, anhydrous conditions have to be ensured in this step, as traces of water might convert NaOtBu into sodium hydroxide, thereby reducing reaction efficiency due to its lower basicity.45 In the next step, cyclization between the aniline intermediate and a 2-(ethoxycarbonyl)arylboronic acid 3 has to be initiated. In addition to the Suzuki coupling, the N-arylaniline as a nucleophile should attack the ethyl ester to form the lactam. The addition of water and Cs2CO3 in this step suppresses the formation of the undesired dihydrophenazine byproduct (see SI). Ultimately, ortho-bromoanilines 1 and aryl iodides 2 are successfully subjected to Buchwald–Hartwig amination in the presence of catalytic amounts of Pd(dba)2 and [tBu3PH]BF4, as well as an excess of NaOtBu, at 35 °C for 15 h. The subsequent addition of 2-(methoxycarbonyl)arylboronic acids 3, Cs2CO3, and water at 120 °C for 15 h induces cyclization, affording after flash chromatography three N-arylphenanthridinones 4 and three N-arylcrinasiadines 5 in moderate to good yields ranging from 15 to 68% (Scheme 1). This corresponds to a yield per bond formation of up to 89%. The presence of a fluoro substituent and a tert-butyl group improves solubility and consequently leads to higher yields. The benzodioxole boronic acid ester 3b required for the preparation of N-arylcrinasiadines 5 can be successfully synthesized from piperonal according to a literature protocol (see SI).46–48 In comparison to N-arylphenanthridinones 4, the 1,3-benzodioxole causes an increase in yield, possibly due to improved solubility and easier purification of N-arylcrinasiadines 5.
image file: d5ra06934c-s1.tif
Scheme 1 BHSL synthesis of N-arylphenanthridinones 4 and N-arylcrinasiadines 5 (all reactions were performed on a 0.5 mmol scale in 1,4-dioxane (3.5 mL) and yields are given after purification by flash chromatography on silica gel).

Furthermore, the optimized Suzuki coupling can be successfully combined with lactamization and nucleophilic substitution in a consecutive three-component reaction (see SI). Here, a solvent mixture of 1,4-dioxane and DMF in a ratio of 2[thin space (1/6-em)]:[thin space (1/6-em)]1 is used for the Suzuki coupling, which increases the solubility of the 6(5H)-phenanthridinone intermediate and allows the subsequent alkylation of the lactam nitrogen with alkyl halides 2 to proceed more efficiently. In addition, an iodoaniline 1 is used in the first step, as iodides are more reactive than bromides.44 The Suzuki coupling-lactamization-substitution (SLS) sequence finally provides after flash chromatography, access to three N-alkylphenanthridinones 6 and two N-alkylcrinasiadines 7 with yields ranging from 7 to 44% (Scheme 2). These include the two natural products N-methylcrinasiadin 7a and N-phenethylcrinasiadin 7b, which are known for their cytotoxicity against leukemia, lung, and colon cancer cells.19 The literature known syntheses to date for the preparation of 7a and 7b are multi-step syntheses, but no one-pot methodologies, i.e., consecutive MCR.19,49 The incorporation of a fluorine substituent in derivative 6b improves solubility and consequently leads to the highest yield. In addition to methyl iodide (2c), (2-bromoethyl)benzene (2d) can also be employed in the terminal alkylation. However, the inclusion of 2d leads to a lower yield. It cannot be excluded that the presence of the palladium catalyst triggers the dehydrobromination of the substrate by competing β-H-elimination.50


image file: d5ra06934c-s2.tif
Scheme 2 SLS sequence of N-alkylphenanthridinones 6 and N-alkylcrinasiadines 7 (all reactions were performed on a 0.5 mmol scale in 1,4-dioxane (3.5 mL) and yields are given after purification by flash chromatography on silica gel).

The structures of all phenanthridinone analogues 4–7 were assigned by extensive NMR spectroscopy, mass spectrometry, and IR spectroscopy (see SI). In addition, the molecular composition of the novel compounds was determined by HPLC-HRMS (see SI).

Photophysical properties

Although N-arylphenanthridinones have been shown to display dual emission,34,35 the ease of their accessibility through our one-pot sequence prompted us to take a closer look at the photophysics, in particular, to scrutinize the origin of the observed dual emission. This phenomenon was attributed to a conformational change of the N-aryl substituent as supported by solvatochromism and time-resolved fluorescence studies.35 Alternatively, an anti-Kasha behavior was discussed, which involves emission from higher excited states as substantiated by quantum chemical calculations.34 Therefore, the photophysical properties of all compounds were investigated by UV/vis absorption and fluorescence spectroscopy in dichloromethane or acetonitrile (Table 1).
Table 1 Selected photophysical properties (absorption maxima in solution with absorption coefficients ε and emission maxima in solution with fluorescence quantum yields Φem and Stokes shifts Δ[small nu, Greek, tilde]s) of all phenanthridinones 4–7

image file: d5ra06934c-u1.tif

Compound R1 R2 Solvent λmax,absa [nm] (ε [M−1 cm−1]) λmax,emb [nm],/Δ[small nu, Greek, tilde]sd [cm−1] (Φem)c
a Recorded at T = 293[thin space (1/6-em)]K, c = 10−5 M.b Recorded at T = 293 K, c = 10−7 M.c Absolute over all quantum yields recorded at T = 293 K, c = 10−6 M.d Δ[small nu, Greek, tilde]s = 1/λmax,abs − 1/λmax,em.
4a H Ph CH2Cl2 232 (35[thin space (1/6-em)]300), 237 (35[thin space (1/6-em)]000), 260 (13[thin space (1/6-em)]200), 306 (7200), 322 (7300), 336 (5700) 372/2900, 524/10[thin space (1/6-em)]700
4b F p-tBuC6H4 CH2Cl2 237 (34[thin space (1/6-em)]700), 261 (13[thin space (1/6-em)]000), 273 (sh, 8100), 319 (5400), 331 (7200), 345 (sh, 6100) 382/2800, 530/10[thin space (1/6-em)]100 (0.01)
      MeCN 243 (sh, 40[thin space (1/6-em)]700), 248 (44[thin space (1/6-em)]100), 255 (sh, 41[thin space (1/6-em)]900), 281 (sh, 14[thin space (1/6-em)]100), 291 (sh, 9400), 329 (sh, 8300), 348 (8400), 363 (sh, 7100) 376/2600, 579/10[thin space (1/6-em)]300
4c F Ph CH2Cl2 236 (41[thin space (1/6-em)]000), 260 (15[thin space (1/6-em)]100), 272 (sh, 9600), 314 (6100), 329 (8900), 343 (7400) 374/2400, 548/10[thin space (1/6-em)]900
5a H Ph CH2Cl2 249 (4800), 270 (15[thin space (1/6-em)]700), 286 (sh, 10[thin space (1/6-em)]900), 296 (sh, 9700), 311 (sh, 11[thin space (1/6-em)]500), 318 (11[thin space (1/6-em)]500), 339 (6600) 367/2100, 519/12[thin space (1/6-em)]200
5b F p-tBuC6H4 CH2Cl2 249 (43[thin space (1/6-em)]400), 274 (13[thin space (1/6-em)]100), 287 (sh, 10[thin space (1/6-em)]500), 298 (sh, 8200), 317 (sh, 9700), 329 (10[thin space (1/6-em)]300), 346 (8000) 372/2000, 550/10[thin space (1/6-em)]700 (0.02)
      MeCN 247 (35[thin space (1/6-em)]600), 268 (35[thin space (1/6-em)]500), 292 (sh, 10[thin space (1/6-em)]200), 306 (sh, 8100), 316 (sh, 6700), 328 (sh, 8400), 345 (6400), 365 (6300) 372/2100, 562/9600
5c F Ph CH2Cl2 248 (16[thin space (1/6-em)]900), 273 (5200), 287 (4200), 299 (3400), 316 (sh, 3800), 329 (3900), 346 (5300) 371/2000, 543/10[thin space (1/6-em)]500
6a H Me CH2Cl2 233 (49[thin space (1/6-em)]200), 239 (50[thin space (1/6-em)]000), 262 (18[thin space (1/6-em)]600), 276 (sh, 9000), 312 (sh, 7000), 325 (9700), 339 (7600) 367/2300
6b F Me CH2Cl2 236 (39[thin space (1/6-em)]500), 253 (13[thin space (1/6-em)]100), 262 (15[thin space (1/6-em)]700), 273 (8900), 317 (sh, 6300), 331 (9100), 346 (7700) 373/2100
6c H Phenethyl CH2Cl2 234 (33[thin space (1/6-em)]400), 240 (34[thin space (1/6-em)]400), 252 (sh, 12[thin space (1/6-em)]300), 262 (13[thin space (1/6-em)]200), 275 (sh, 7100), 310 (sh, 5700), 325 (6400), 339 (5300) 367/2300
7a H Me CH2Cl2 247 (45[thin space (1/6-em)]000), 269 (17[thin space (1/6-em)]300), 285 (sh, 10[thin space (1/6-em)]500), 297 (sh, 9500), 313 (sh, 12[thin space (1/6-em)]800), 319 (11[thin space (1/6-em)]900), 324 (sh, 11[thin space (1/6-em)]100), 340 (8700) 364/2000
7b H Phenethyl CH2Cl2 249 (35[thin space (1/6-em)]700), 270 (15[thin space (1/6-em)]900), 286 (sh, 8400), 297 (sh, 7500), 313 (sh, 8700), 318 (8700), 325 (sh, 8100), 340 (5400) 364/2000


The N-aryl/N-alkylphenanthridinones 4 and 6 exhibit an intense absorption maximum with a shoulder at high energy in the UV. In contrast, the N-aryl/N-alkylcrinasiadines 5 and 7 display only one characteristic maximum without a shoulder in this region, which, unlike compounds 4 and 6, is slightly bathochromically shifted. The subsequent longer wavelength maxima of compounds 5 and 7 are also slightly bathochromically shifted in comparison to 4 and 6. At lowest energy, three maxima appear for each compound, which are attributed to vibronic fine structure of the same electronic transition. The longest wavelength absorption maxima of all compounds 4–7 occurring at a similar wavelength to that of the unsubstituted 6(5H)-phenanthridinone 8 (R[double bond, length as m-dash]H, isolated intermediate) (Fig. 2). Therefore, the longest wavelength absorption maximum of all compounds probably corresponds to the HOMO → LUMO transition, in accordance with literature for compound 8.35,51


image file: d5ra06934c-f2.tif
Fig. 2 Comparison of the normalized UV/vis absorption and emission spectra of phenanthridinones 4a, 5a, 6a, 7a and 6(5H)-phenanthridinone 8 (absorption spectra recorded in CH2Cl2, T = 293 K, c = 10−5 M (bold lines) and emission spectra recorded in CH2Cl2, T = 293 K, c = 10−7 M (dashed lines)).

Furthermore, the absorption maxima at lower energy of the N-aryl/N-alkylphenanthridinones 4b–c and 6b, as well as the N-arylcrinasiadines 5b–c, are very similar and slightly bathochromically shifted compared to compounds 4a, 5a, 6a, 6c, and 7a–b devoid of fluorine substituent at position R1. Consequently, the substituent at position R2, unlike that at position R1, has only a minor influence on the spectra for the present derivatives. Therefore, it can be assumed that the longest wavelength absorption band is primarily associated with the phenanthridinone moiety.

The polarity of the solvents influences the position of the absorption maxima, as illustrated by the bathochromic shift observed in acetonitrile (Table 1). This suggests that the absorption bands cannot be attributed solely to a locally excited state (LE band), but rather also possess partial charge transfer (CT) character. The shapes of the absorption bands remain largely unchanged, which indicates a structurally similar electronic transition in both solvents (Fig. 3).


image file: d5ra06934c-f3.tif
Fig. 3 Comparison of the normalized absorption spectra of N-arylphenanthridinone derivatives 4b (A) and 5b (B) in acetonitrile (black lines) and dichloromethane (gray lines) (T = 293 K, c = 10−5 M).

The emission spectra of the N-arylsubstituted compounds 4 and 5 show two emission maxima, in contrast to those of the N-alkylsubstituted compounds 6 and 7, which indicate dual fluorescence (Fig. 4). The shorter wavelength emission band only displays a small Stokes shift Δ[small nu, Greek, tilde]s, indicating emission from an LE band, as only minor geometric changes are likely to occur between the ground and excited states. Moreover, the shape and position of the LE band closely resemble the emission maximum of 6(5H)-phenanthridinone 8 (Fig. 2).35,51 The emission is therefore likely to originate from the phenanthridinone core.


image file: d5ra06934c-f4.tif
Fig. 4 Normalized emission spectra of N-arylphenanthridinone derivatives 4–5 in CH2Cl2 (A) and comparison of emission spectra of 4b and 5b in CH2Cl2 and of 4b and 5b in MeCN (B) (emission spectra recorded in CH2Cl2 or MeCN, T = 293 K, c = 10−7 M).

The second emission maximum is, unlike the LE band, strongly bathochromically shifted. This maximum is only weakly significant for compounds 4c and 5c, which bear a fluorine substituent at R1 and no substituent at R2. Investigations of the emission of 4b and 5b in dichloromethane and acetonitrile reveal that the intensity of the longer wavelength emission band increases relative to the LE band with increasing solvent polarity (Fig. 4). Moreover, since in acetonitrile the LE transition hardly shifts, the insensitivity to solvent polarity (Table 1) can be attributed to a π–π* transition. In contrast, the longer wavelength emission maxima exhibit a pronounced bathochromic shift with increasing polarity form dichloromethane to acetonitrile (positive solvatochromism). This emission maximum is therefore likely a CT band, as polar solvents stabilize charge separation more effectively than nonpolar ones.35 However, the LE emission remains more dominant in both solvents. Independent of the excitation wavelength the emission maxima appear at the same wavelength. The degassed and non-degassed emission spectra of compounds 4b and 5b in acetonitrile differ only marginally (see SI), which likely rules out the assignment of the longest wavelength emission band to be phosphorescence. Quantum chemical calculations might unravel the origin of the bands, either from a higher excited state (anti-Kasha behavior)34 or from a conformational change35 in the first excited state (vide infra).

Overall, a comparison of the emission spectra of the N-alkyl substituted compounds 6 and 7, as well as 6(5H)-phenanthridinone 8 (see SI), indicates that the presence of an aryl group at the nitrogen is essential for the occurrence of dual emission, as both an LE and a CT band were only observed for compounds 4 and 5. Furthermore, the nature of the substituents at positions R1 and R2 influences the intensity ratio of the LE to the CT band. The CT emission increases relative to the LE band for R1 bearing an electron-withdrawing substituent and for R2 bearing an electron-donating substituent, which is consistent with observations of Demeter et al.35 This additionally supports the hypothesis of a population of the phenanthridinone core in the excited state, which might be scrutinized through theoretical analyses (vide infra).

The absolute fluorescence quantum yield Φem could only be determined for compounds 4b and 5b, with values of 1% and 2%, respectively, whereas the other compounds exhibit emission that is too weak to be recorded (Table 1). At room temperature, the dominant deactivation process of the excited S1 state in phenanthridinones is intersystem crossing (ISC) to the triplet state, which accounts for the very low fluorescence quantum yield Φem.35

Calculated electronic structure

A deeper insight into the electronic structure of N-arylphenanthridinones 4 and N-arylcrinasiadines 5 can be obtained by (TD)DFT calculations exemplarily for compounds 4b and 5b with the program package Gaussian 16.52 In analogy to the literature, the B3LYP53,54 functional and the Pople basis set 6-31G* (ref. 55) were employed for the quantum chemical analysis of the phenanthridinones.34 All minima were confirmed unambiguously by analytical frequency analysis (NIMAG = 0). Based on the absorption and emission properties in solution, the intrinsic polarizable continuum model (PCM) with dichloromethane as the dielectric medium was applied in the quantum chemical calculations.56 The calculated absorption and emission energies agree quantitatively well with the experimental data in dichloromethane for the selected examples, with only minor over- or underestimation of the energy levels (Table 2).
Table 2 (TD)DFT calculations on the UV/vis absorption and emission maxima of N-arylphenanthridinone 4b and N-arylcrinasiadine 5b (Gaussian 16, B3LYP/6-31G*, PCM CH2Cl2, for details see SI)

image file: d5ra06934c-u2.tif

Com-pound R1 R2 λmax,abs(exp)a [nm] (ε [M−1 cm−1]) λmax,abs(calc) [nm] (oscillator strength f), most dominant contribution λmax,em(exp)b [nm] λmax,em(calc) [nm] (oscillator strength f), most dominant contribution
a Recorded in CH2Cl2, T = 293[thin space (1/6-em)]K, c = 10−5 M.b Recorded in CH2Cl2, T = 293 K, c = 10−7 M.
4b F tBu 237 (34[thin space (1/6-em)]700) 229 (0.0247) HOMO-3 → LUMO+1 (70%) 382, 530 405 (0.3277) HOMO → LUMO (95%), 579 (0.1270) HOMO → LUMO (99%)
231 (0.3859) HOMO → LUMO+4 (49%)
235 (0.2417) HOMO-4 → LUMO (41%)
248 (0.0256) HOMO-3 → LUMO (88%)
261 (13[thin space (1/6-em)]000) 256 (0.1342) HOMO-1 → LUMO+1 (47%)
269 (0.0506) HOMO-1 → LUMO (71%)
273 (sh, 8100) 292 (0.0311) HOMO → LUMO+1 (90%)
319 (5400) 316 (0.2320) HOMO → LUMO (93%)
331 (7200)
345 (sh, 6100)
5b F tBu 249 (43[thin space (1/6-em)]400) 241 (0.1814) HOMO → LUMO+4 (53%) 372, 550 386 (0.3533) HOMO → LUMO (96%), 563 (0.1417) HOMO → LUMO (99%)
274 (13[thin space (1/6-em)]100) 247 (0.5248) HOMO-3 → LUMO (44%)
287 (sh, 10[thin space (1/6-em)]500) 251 (0.0154) HOMO → LUMO+3 (84%)
268 (0.1603) HOMO → LUMO (77%)
286 (0.1849) HOMO-1 → LUMO (49%)
298 (sh, 8200) 301 (0.0114) HOMO → LUMO+1 (51%)
317 (sh, 9700) 319 (0.2547) HOMO → LUMO (89%)
329 (10[thin space (1/6-em)]300)
346 (8000)


Since the absorption spectra of structures 4b and 5b in dichloromethane display very similar profiles, only the calculated absorption spectrum of 4b will be discussed in detail (Fig. 5). The vertical bars represent the calculated transitions and illustrate that the longest wavelength absorption band is primarily dominated by the HOMO → LUMO transition (Fig. 5A). In the calculated spectrum, the transitions combine into broad bands at both lower and higher energy. In the experimental spectrum, these appear as weakly pronounced shoulders, indicating the superposition of multiple transitions. Frontier molecular orbitals (FMO) are expectedly involved in the dominant electronic transitions (Fig. 5B). The S0–S1 and S0–S13 transitions mainly occur between delocalized π-orbitals of the same conjugated system, indicating an LE character of these states. Only the S0–S12 transition, in which the electron density is transferred from the more localized HOMO−4 to the delocalized LUMO, suggests a slight CT character. TDDFT calculations showed that S1 predominantly results from the HOMO → LUMO transition. In contrast, the higher excited states are composed of multiple orbital pairs and can thus be interpreted as mixtures of various electronic transitions. The energies of the FMO of compounds 4b and 5b are visualized in dichloromethane (Fig. 6). A comparison of N-arylphenanthridinone 4b with the structurally related crinasiadine 5b reveals that the additional oxygen atoms in crinasiadine exert electron-donating effects, leading to a slight increase of both the HOMO and LUMO energy levels.57 Since the HOMO is raised slightly higher in comparison to the LUMO, the energy gap ΔE(EHOMOELUMO) of crinasiadine 5b is slightly smaller than that of compound 4b.


image file: d5ra06934c-f5.tif
Fig. 5 Comparison of the calculated (Gaussian 16, B3LYP/6-31G*, PCM CH2Cl2 (dashed lines)) and experimentally determined (recorded in CH2Cl2, T = 293 K, c = 10−5 M (bold lines)) UV/vis spectra of N-arylphenanthridinone 4b with the calculated transitions shown as gray bars (A) and calculated molecular orbitals of 4b for the dominant electronic transitions (B) (Gaussian 16, B3LYP/6-31G*, PCM CH2Cl2, isosurface value 0.025 a.u.).

image file: d5ra06934c-f6.tif
Fig. 6 Selected Kohn–Sham FMO of N-arylphenanthridinone 4b and N-arylcrinasiadine 5b employing the PCM with dichloromethane as solvent (Gaussian 16, B3LYP/6-31G*, PCM CH2Cl2, isosurface value at 0.025 a.u.).

However, this difference is not reflected in the experimental absorption spectra, as the position of the longest-wavelength absorption band is nearly identical for both compounds. This can be explained by the fact that the experimental band results from the superposition of several electronic transitions and therefore cannot be directly equated with the calculated HOMO → LUMO transition. For both compounds 4b and 5b, the coefficient densities of the HOMO and LUMO are predominantly localized on the phenanthridinone moiety, confirming the LE character of the lowest electronic transitions.

Demeter et al. investigated various dual fluorescent N-arylphenanthridinones, including N-arylphenanthridinone 4a. They recorded fluorescence decay curves in both the LE and CT regions of the spectrum and observed biexponential decays. From this, they concluded that the CT state is not directly excited from the ground state but originates instead from the initially excited LE state.35 They further postulated that the CT state must be significantly more planar than the LE state, as no CT fluorescence is observed in twisted N-arylphenanthridinones. In these cases, complete planarity is sterically hindered by ortho-methyl substituents on the aryl group.35 Based on these results, a potential energy scan was carried out in both the ground (S0 and S0*) and excited states (S1) to investigate the geometry dependence of the electronic states and possible relaxation pathways between the LE and CT states using quantum chemical methods. The CN-bond rotation was modeled by a torsional scan by 360° rotation of the N-aryl substituent with the torsion angle θ (Fig. 7). The initial geometry is defined as a rotational angle θ of 0°. The energy curves of compounds 4b and 5b are very analogous in the vibrationally relaxed (S0) and vibrationally excited (S0*) ground states (see SI), with the perpendicular geometry representing the global minimum in each case (NIMAG = 0), i.e. thermodynamically preferred conformations. In contrast, the coplanar orientation of the phenanthridinone unit and the aryl substituent occurs in a transition state with an imaginary frequency (NIMAG = 1). This rotational barrier ΔE amounts to 0.72 eV (17.0 kcal mol−1) for compound 4b and 0.74 eV (17.1 kcal mol−1) for compound 5b, which can be thermally surmounted at room temperature, yet significantly slowed down kinetically.58 The rate constants k (k(4b) = 4.14 s−1 and k(5b) = 1.90 s−1) of the rotation around the C–N-bond can be estimated using the Eyring equation59 to proceed quite slowly.


image file: d5ra06934c-f7.tif
Fig. 7 Illustration of the minimum geometry of 4b as the starting geometry for the torsion scan with the torsion angle θ, which describes the rotation around the N-aryl bond (A), with x1 as the N-aryl bond length and x2 as the N–CO amide bond length (B).

The energy curves of compounds 4b and 5b in the excited state (S1) also exhibit very similar profiles (see SI). In contrast to the ground states, however, the coplanar geometry (NIMAG = 0) now represents the global minimum, while a conformer slightly twisted from a perpendicular geometry constitutes a second local minimum (NIMAG = 0). The coplanar geometry reveals a shorter N-aryl bond length x1, which therefore indicates partial double bond character and suggests a more efficient π-conjugation compared to the perpendicular geometry. Furthermore, the N–CO amide bond length x2 is elongated in the coplanar geometry. Both geometries resemble the N-intra and N-extra conformations of phenothiazine.60 In the N-intra conformation, the nitrogen lone pair of phenothiazine overlaps with the π-system of the anellated benzene rings due to the quasi-equatorial orientation of the substituent. In the N-extra conformation, the substituent adopts a quasi-axial position. Unlike in phenothiazine, however, the nitrogen atom of the amide group is not sp3-hybridized and pyramidal, but sp2-hybridized and arranged in a trigonal planar geometry.35,61,62 Therefore, it is more appropriate to refer to a perpendicular pseudo-N-intra conformation and a coplanar pseudo-N-extra conformation. These two conformations are illustrated exemplarily for compound 4b (Fig. 8).


image file: d5ra06934c-f8.tif
Fig. 8 Comparison of the excited state geometries of pseudo-N-intra (A) and pseudo-N-extra (B) from different perspectives for compound 4b (top, Gaussian 16, B3LYP/6-31G*, PCM CH2Cl2), along with a schematic representation of the N-aryl group (Ar) in the quasi-equatorial intra-conformation and quasi-axial extra-conformation (bottom, R = anellated benzene rings).

The energy barrier ΔE(S1(LE)-S1(CT)) between the two conformers S1(LE) and S1(CT) in the excited state is 0.33 eV (7.61 kcal mol−1) for compound 4b and 0.30 eV (7.07 kcal mol−1) for compound 5b, and is easily overcome at room temperature,58 which may explain the occurrence of dual emission at this temperature. The calculated rate constants k in the excited state (k(4b) = 1.63 × 107 s−1 and k(5b) = 5.24 × 107 s−1) support this hypothesis, as the conformational interconversion proceeds with high efficiency. Assuming a Boltzmann distribution, the theoretical Gibbs free energy ΔG of both conformers (ΔG(4b) = 0.30 eV (6.89 kcal mol−1) and ΔG(5b) = 0.23 eV (5.22 kcal mol−1)) yields the equilibrium constants K for dyes 4b (7.29 × 10−6) and 5b (1.27 × 10−4) corresponding to a population ratio of 99.987 to 0.013 in favor of the coplanar pseudo-N-extra conformation S1(CT). The dual emission is therefore a consequence of the kinetic competition between emission from the pseudo-N-intra conformation S1(LE) and conformational interconversion to conformer S1(CT), leading to partial emission from both conformers.35

For a quantitative assessment of the dual emission behavior, the ratio of the integral emission intensities of both bands on a Jacobian energy-corrected scale can be considered (see SI, chpt. 9.5) and compared with the theoretically calculated oscillator strengths.63 For dye 4b, the integrated emission spectra give a ratio ALE/ACT of 2.14, while the quantum-chemically calculated oscillator strengths give a ratio fLE/fCT of 2.58. Thus, fLE/fCT moderately overestimates ALE/ACT but remains consistent in trend with the experimental result, thereby supporting the assignment of the two emission channels. For dye 5b, however, fLE/fCT (= 2.49) does not adequately capture ALE/ACT (= 4.25). The deviations can be attributed to differences in baseline choice and contributions from nonradiative processes.64,65

Jabłoński diagrams illustrate the absorption and emission of compounds 4b (Fig. 9 (A)) and 5b (Fig. 9(B)). In the pseudo-N-intra conformation, the vibrationally excited S1* state lies only slightly above the vibrationally relaxed S1 LE-state in energy (ΔE(4b) = 0.32 eV and ΔE(5b) = 0.24 eV), which represent small Stokes shifts (Δ[small nu, Greek, tilde]s(exp) (4b) = 2800 cm−1 vs. Δ[small nu, Greek, tilde]s(calc) (4b) = 7000 cm−1 and Δ[small nu, Greek, tilde]s(exp (5b) = 2000 cm−1 vs. Δ[small nu, Greek, tilde]s(calc) (5b) = 5400 cm−1). As a result, only minimal structural changes occur and the aryl substituent only slightly twists. This twisting causes the aryl unit to bear less electron density compared to the Franck–Condon geometry. Nevertheless, the electron density remains predominantly localized on the phenanthridinone moiety. Even after emission from the S1 state, the coefficient density remains localized on the phenanthridinone core and the resulting transition can be assigned to the short wavelength LE band in the experimental emission spectrum. However, due to coefficient density on the aryl substituent in the vibrationally excited S0* state, the transition cannot be regarded as a pure LE transition.


image file: d5ra06934c-f9.tif
Fig. 9 Schematic Jabłoński diagrams and Kohn–Sham-FMOs of the calculated S1 and T1 states, including the S0–S1* transition (absorption at the longest wavelength), the S1–S0* transition of the pseudo-N-intra (LE fluorescence) and pseudo-N-extra (CT fluorescence), the S1–T1* transition (ISC), as well as the T1–S0* transition of pseudo-N-intra (phosphorescence) for N-arylphenanthridinone 4b (A) and N-arylcrinasiadine 5b (B) (Gaussian 16, B3LYP/6-31G*, PCM CH2Cl2, isosurface value 0.025 a.u.).

The conformational change to the pseudo-N-extra conformer lowers of the excited-state energy level. Planarization enables more efficient delocalization of the electron density, which is also reflected in the shorter N-aryl bond length x1 of the conformer. In the pseudo-N-extra conformation, the vibrationally excited S1* state lies significantly above the vibrationally relaxed S1 CT-state in energy (ΔE (4b) = 0.58 eV and ΔE (5b) = 0.49 eV). As a result of significant structural changes, which cause large Stokes shifts (Δ[small nu, Greek, tilde]s(exp) (4b) = 10[thin space (1/6-em)]100 cm−1 vs. Δ[small nu, Greek, tilde]s(calc) (4b) = 12[thin space (1/6-em)]800 cm−1 and Δ[small nu, Greek, tilde]s(exp) (5b) = 10[thin space (1/6-em)]700 cm−1 vs. Δ[small nu, Greek, tilde]s(calc) (5b) = 13[thin space (1/6-em)]200 cm−1) and a smaller energy gap ΔE, the redshifted emission is observed. Therefore, S1–S0* CT-transition from the phenanthridinone to the aryl unit occurs. The coefficient distribution of the T1 state of the pseudo-N-intra conformation is for both compounds 4b and 5b very similar to the S1 state but lies energetically well below the S1 state of the pseudo-N-extra conformation (see SI).

Nevertheless, according to El-Sayed's rule ISC of similar transitions can be considered to occur less probable.66 The quantum chemically calculated phosphorescence of compounds 4b and 5b corresponds well with the experimentally observed long-wave emission bands. But the nearly identical experimental emission spectra in degassed and non-degassed solutions (see SI) indicate that no significant oxygen-induced triplet quenching occurs. Furthermore, the relatively large energy gap ΔEST between T1 and S1 in the pseudo-N-intra renders an ISC unlikely. Alternatively, a conformational change to the pseudo-N-extra conformation might occur prior to ISC, from which a transition to the T1 state of this geometry could take place, or an ISC to higher triplet states might occur, which could account for the low fluorescence quantum yield Φem.35

The localization of the LUMO on the phenanthridinone moiety in both conformers further explains the experimentally observed substituent effects on the LE to CT band ratio (Fig. 6). The weak fluorine acceptor at position R1 stabilizes the electron density in the excited state and thus favors both bands, while electron donors would probably cause destabilization. Donors at R2 facilitate mesomeric delocalization in the LUMO especially in a planar geometry thus enhancing CT emission. In contrast, acceptors at R2 withdraw electron density and destabilize the planar conformation and shift the emission ratio in favor of the LE band, which is consistent with experimental observations.

Conclusions

The consecutive sequentially palladium catalyzed three-component synthesis provides a concise approach to functionalized phenanthridinones and crinasiadines in a one-pot fashion. While N-aryl derivatives are obtained by this sequence, N-alkyl derivatives form by Suzuki arylation–lactamization followed by N-alkylation, also in a one-pot fashion. Particularly noteworthy is the concise synthesis of two natural products, N-methylcrinasiadine and N-phenethylcrinasiadine, whose cytotoxic effects against leukemia, lung, and colon cancer cells underscore the biological relevance of the synthetic derivatives.

Photophysical investigations reveal dual fluorescence of these compounds which can be attributed to involvement of LE (locally excited) and CT (charge transfer) states. The substitution pattern clearly affects the emission behavior, allowing fine-tuning of electronic properties to achieve desired material characteristics. Quantum chemical calculations provide initial rationalization of the origin of the LE and CT bands and suggest that the LE band emerges from the pseudo-N-intra conformation, while the redshifted CT band arises from the pseudo-N-extra conformation.

The results pave the way for the development of novel, accurately tunable phenanthridinone and crinasiadine emitters. Further studies on the syntheses of alkaloid analogues with increased emission quantum yield are currently underway.

Author contributions

The work consists of parts of the planned PhD thesis of R. K., which is supervised by T. J. J. M. Methodologic and synthetic studies, analytical characterization of the synthetic samples, photophysics (steady-state absorption and emission spectroscopy), and quantum chemical calculation by R. K., who compiled the obtained data. The conceptualization was outlined by T. J. J. M. The writing of the first draft was completed by R. K., and the review and editing was completed by R. K. and T. J. J. M. Project administration and funding acquisition by T. J. J. M. Both authors have read and agreed to the published version of the manuscript.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the supplementary information (SI). Supplementary information: synthetic details, 1H and 13C NMR spectra, HPLC traces, absorption and emission spectra as well as details on the (TD)DFT calculations. See DOI: https://doi.org/10.1039/d5ra06934c.

Acknowledgements

The authors are grateful to the Deutsche Forschungsgemeinschaft DFG (Mu 1088/9-1) and the Fonds der Chemischen Industrie for financial support. We also thank the CeMSA@HHU (Center for Molecular and Structural Analytics @ Heinrich Heine University) for recording the mass spectrometric and the NMR spectroscopic data.

Notes and references

  1. M. Feng, B. Tang, H.-X. Xu and X. Jiang, Org. Lett., 2016, 18, 4352–4355 CrossRef CAS PubMed.
  2. D. Liang, Z. Hu, J. Peng, J. Huang and Q. Zhu, Chem. Commun., 2013, 49, 173–175 RSC.
  3. R. R. Aleti, A. A. Festa, L. G. Voskressensky and E. V. Van der Eycken, Molecules, 2021, 26, 5560 CrossRef CAS PubMed.
  4. R. S. Theobald and K. Schofield, Chem. Rev., 1950, 46, 170–189 CrossRef.
  5. H. Gilman and J. Eisch, J. Am. Chem. Soc., 1957, 79, 5479–5483 CrossRef CAS.
  6. E. C. Horning, V. L. Stromberg and H. A. Lloyd, J. Am. Chem. Soc., 1952, 74, 5153–5155 CrossRef CAS.
  7. C. C. Woodroofe, B. Zhong, X. Lu and R. B. Silverman, J. Chem. Soc., Perkin Trans. 2, 2000, 55–59 RSC.
  8. K. Tanimoto, N. Nakagawa, K. Takeda, M. Kirihata and S. Tanimori, Tetrahedron Lett., 2013, 54, 3712–3714 CrossRef CAS.
  9. G. Wurz, O. Hofer and H. Greger, Nat. Prod. Lett., 1993, 3, 177–182 CrossRef CAS.
  10. H. R. Arthur, W. H. Hui and Y. L. Ng, J. Am. Chem. Soc., 1959, 4007–4009 RSC.
  11. H. R. Arthur, W. H. Hui and Y. L. Ng, J. Am. Chem. Soc., 1959, 1840–1845 RSC.
  12. T. Kametani, K. Kigasawa, M. Hiiragi and O. Kusama, J. Heterocycl. Chem., 1973, 10, 31–33 CrossRef CAS.
  13. X. Li, Q. Wang, L. Liu, Y. Shi, Y. Hong, W. Xu, H. Xu, J. Feng, M. Xie and Y. Li, Pharm, 2024, 17, 524 CAS.
  14. S. F. Martin, The Amaryllidaceae Alkaloids, Elsevier, San Diego, 1987 Search PubMed.
  15. G. Van Goietsenoven, A. Andolfi, B. Lallemand, A. Cimmino, D. Lamoral-Theys, T. Gras, A. Abou-Donia, J. Dubois, F. Lefranc and V. Mathieu, J. Nat. Prod., 2010, 73, 1223–1227 CrossRef CAS PubMed.
  16. G. Van Goietsenoven, V. Mathieu, F. Lefranc, A. Kornienko, A. Evidente and R. Kiss, Med. Res. Rev., 2013, 33, 439–455 CrossRef CAS PubMed.
  17. A. Evidente, A. S. Kireev, A. R. Jenkins, A. E. Romero, W. F. A. Steelant, S. Van Slambrouck and A. Kornienko, Planta Med., 2009, 75, 501–507 CrossRef CAS PubMed.
  18. M. Ghavre, J. Froese, M. Pour and T. Hudlicky, Angew. Chem., Int. Ed., 2016, 55, 5642–5691 CrossRef CAS PubMed.
  19. Z. Luo, F. Wang, J. Zhang, X. Li, M. Zhang, X. Hao, Y. Xue, Y. Li, F. D. Horgen and G. Yao, J. Nat. Prod., 2012, 75, 2113–2120 CrossRef CAS PubMed.
  20. M. de Fatima Pereira, C. Rochais and P. Dallemagne, Curr. Med. Chem.:Anti-Cancer Agents, 2015, 15, 1080–1091 CrossRef PubMed.
  21. J.-H. Li, L. Serdyuk, D. V. Ferraris, G. Xiao, K. L. Tays, P. W. Kletzly, W. Li, S. Lautar, J. Zhang and V. J. Kalish, Bioorg. Med. Chem. Lett., 2001, 11, 1687–1690 CrossRef CAS PubMed.
  22. M. Mann, S. Kumar, A. Sharma, S. S. Chauhan, N. Bhatla, S. Kumar, S. Bakhshi, R. Gupta and L. Kumar, Oncotarget, 2019, 10, 4262 CrossRef PubMed.
  23. P. Jagtap, F. G. Soriano, L. Virág, L. Liaudet, J. Mabley, É. Szabó, G. Haskó, A. Marton, C. B. Lorigados and F. Gallyas Jr, Crit. Care Med., 2002, 30, 1071–1082 CrossRef CAS PubMed.
  24. B. Zingarelli, S. Cuzzocrea, Z. Zsengellér, A. L. Salzman and C. Szabó, Cardiovasc. Res., 1997, 36, 205–215 CrossRef CAS PubMed.
  25. J. DeRuiter, B. E. Swearingen, V. Wandrekar and C. A. Mayfield, J. Med. Chem., 1989, 32, 1033–1038 CrossRef CAS PubMed.
  26. H. J. Crowley, B. Yaremko, W. M. Selig, D. R. Janero, C. Burghardt, A. F. Welton and M. O'Donnell, J. Pharmacol. Exp. Ther., 1991, 259, 78–85 CrossRef CAS PubMed.
  27. S. Pegoraro, M. Lang, T. Dreker, J. Kraus, S. Hamm, C. Meere, J. Feurle, S. Tasler, S. Prütting and Z. Kuras, Bioorg. Med. Chem. Lett., 2009, 19, 2299–2304 CrossRef CAS PubMed.
  28. E. Sperotto, G. P. M. van Klink, G. van Koten and J. G. de Vries, Dalton Trans., 2010, 39, 10338–10351 RSC.
  29. F. Meng, X.-T. Nguyen, X. Cai, J. Duan, M. Matteucci and C. P. Hart, Anticancer Drugs, 2007, 18, 435–445 CrossRef CAS PubMed.
  30. T. A. Grese, M. D. Adrian, D. L. Phillips, P. K. Shetler, L. L. Short, A. L. Glasebrook and H. U. Bryant, J. Med. Chem., 2001, 44, 2857–2860 CrossRef CAS PubMed.
  31. A. H. Parham, C. Pflumm, A. Jatsch, T. Eberle and P. Stoessel, Organic Electroluminescent Device, US Pat., 9337430B2, 2016 Search PubMed.
  32. R. L. d. Costa, D. A. F. da Silva, N. C. d. Lucas and S. J. Garden, Quim. Nova, 2016, 39, 310–319 CAS.
  33. M. Guérette, A. Najari, J. Maltais, J. R. Pouliot, S. Dufresne, M. Simoneau, S. Besner, P. Charest and M. Leclerc, Adv. Energy Mater., 2016, 6, 1502094 CrossRef.
  34. S. K. Ganegamage, Y. Zou and M. D. Heagy, J. Org. Chem., 2023, 88, 11424–11433 CrossRef CAS PubMed.
  35. A. Demeter, T. Bérces and K. A. Zachariasse, J. Phys. Chem. A, 2001, 105, 4611–4621 CrossRef CAS.
  36. D. M. D'Souza and T. J. J. Müller, Chem. Soc. Rev., 2007, 36, 1095–1108 RSC.
  37. M. D. Burke and S. L. Schreiber, Angew. Chem., Int. Ed., 2004, 43, 46–58 CrossRef PubMed.
  38. L. Brandner and T. J. J. Müller, Front. Chem., 2023, 11, 1124209 CrossRef CAS PubMed.
  39. G. H. Posner, Chem. Rev., 1986, 86, 831–844 CrossRef CAS.
  40. L. Levi and T. J. J. Müller, Chem. Soc. Rev., 2016, 45, 2825–2846 RSC.
  41. T. J. J. Müller, Science of Synthesis Series, Georg Thieme Verlag KG, Stuttgart, 2014 Search PubMed.
  42. L. Mayer, R. Kohlbecher and T. J. J. Müller, Chem.–Eur. J., 2020, 26, 15130–15134 CrossRef CAS PubMed.
  43. R. Kohlbecher and T. J. J. Müller, Chem.–Eur. J., 2024, 30, e202304119 CrossRef CAS PubMed.
  44. H. Doucet and J. C. Hierso, Angew. Chem., Int. Ed., 2007, 46, 834–871 CrossRef CAS PubMed.
  45. Y. Sunesson, E. Lime, S. O. Nilsson Lill, R. E. Meadows and P.-O. Norrby, J. Org. Chem., 2014, 79, 11961–11969 CrossRef CAS PubMed.
  46. L. Liang, J. Li, B. Shen, Y. Zhang, J. Liu, J. Chen and D. Liu, Org. Biomol. Chem., 2021, 19, 2767–2772 RSC.
  47. C. A. Moore, B. F. Ohman, M. J. Garman, M. E. Liquori, D. M. Degan, K. B. Voellinger, M. J. DePersis and E. T. Pelkey, Arkivoc, 2018, 50–69 Search PubMed.
  48. U. V. Mentzel, D. Tanner and J. E. Tønder, J. Org. Chem., 2006, 71, 5807–5810 CrossRef CAS PubMed.
  49. Y. Kuwata, M. Sonoda and S. Tanimori, J. Heterocycl. Chem., 2017, 54, 1645–1651 CrossRef CAS.
  50. W. H. Brown, T. Poon, Einführung in die Organische Chemie, Wiley-VCH, Weinheim, 2020 Search PubMed.
  51. B. E. Zaitsev, G. I. Migachev, O. V. Koval'chukova, G. V. Sheban and V. V. Matyushenko, Chem. Heterocycl. Compd., 1992, 28, 1159–1165 CrossRef.
  52. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. J. A. Montgomery, J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman, D. J. Fox, Gaussian 16, Revision A.03, Gaussian, Inc., Wallingford CT, 2016 Search PubMed.
  53. C. Lee, W. Yang and R. G. Parr, Phys. Rev. B:Condens. Matter Mater. Phys., 1988, 37, 785 CrossRef CAS PubMed.
  54. A. D. Becke, J. Chem. Phys., 1993, 98, 1372–1377 CrossRef CAS.
  55. R. Krishnan, J. S. Binkley, R. Seeger and J. A. Pople, J. Chem. Phys., 1980, 72, 650–654 CrossRef CAS.
  56. G. Scalmani and M. J. Frisch, J. Chem. Phys., 2010, 132, 114110 CrossRef PubMed.
  57. I. V. Alabugin, Stereoelectronic Effects: A Bridge between Structure and Reactivity, John Wiley & Sons Verlag, Chichester, 2016 Search PubMed.
  58. S. Vyazovkin, Thermochim. Acta, 2025, 743, 179911 CrossRef CAS.
  59. H. Eyring, J. Chem. Phys., 1935, 3, 107–115 CrossRef CAS.
  60. L. Mayer and T. J. J. Müller, Eur. J. Org Chem., 2021, 2021, 3516–3527 CrossRef CAS.
  61. D. Pan and D. L. Phillips, J. Phys. Chem. A, 1999, 103, 4737–4743 CrossRef CAS.
  62. K. P. C. Vollhardt, N. E. Schore, Organische Chemie, John Wiley & Sons, Hannover, 2011 Search PubMed.
  63. J. Mooney and P. Kambhampati, J. Phys. Chem. Lett., 2013, 4, 3316–3318 CrossRef CAS.
  64. C. Würth, M. Grabolle, J. Pauli, M. Spieles and U. Resch-Genger, Nat. Protoc., 2013, 8, 1535–1550 CrossRef PubMed.
  65. S. J. Strickler and R. A. Berg, J. Chem. Phys., 1962, 37, 814–822 CrossRef CAS.
  66. M. A. El-Sayed, J. Chem. Phys., 1963, 38, 2834–2838 CrossRef CAS.

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