Next Article in Journal
Cytoprotective Polyketides from Sponge-Derived Fungus Lopadostoma pouzarii
Next Article in Special Issue
New Azido Coumarins as Potential Agents for Fluorescent Labeling and Their “Click” Chemistry Reactions for the Conjugation with closo-Dodecaborate Anion
Previous Article in Journal
X-ray Absorption Spectroscopy Study of Iron Site Manganese Substituted Yttrium Orthoferrite
Previous Article in Special Issue
Janus-Type AIE Fluorophores: Synthesis and Properties of π-Extended Coumarin-Bearing Triskelions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Synthesis of Fluorescent Coumarins and Phosphorous-Containing Coumarin-Type Heterocycles via Palladium Catalyzed Cross-Coupling Reactions

1
Faculty of Chemistry and Pharmacy, Sofia University “St. Kliment Ohridski”, 1164 Sofia, Bulgaria
2
Department of Inorganic Solid-State Chemistry, Saarland University, 66123 Saarbrücken, Germany
3
Department of Medicinal Chemistry and Biochemistry, Medicinal University of Sofia, 1431 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(21), 7649; https://doi.org/10.3390/molecules27217649
Submission received: 10 October 2022 / Revised: 24 October 2022 / Accepted: 2 November 2022 / Published: 7 November 2022
(This article belongs to the Special Issue Coumarin and Its Derivatives II)

Abstract

:
Quantum-chemical calculations on the spectral properties of some aryl substituted 3-phosphonocoumarins were performed, and the effect of the substituents in the aryl moiety was evaluated. The structures possessing promising fluorescent properties were successfully synthesized via Suzuki and Sonogashira cross-coupling. The synthetic protocol was also applied for the phosphorous chemoisomer of 3-phosphonocoumarin, 1,2-benzoxaphosphorin, and their carboxylate analogues. The optical properties of the arylated and alkynylated products were experimentally determined. The obtained quantum-chemical and experimental results give the possibility for a fine tuning of the optical properties of phosphorous-containing coumarin systems by altering the substituent at its C-6 position.

Graphical Abstract

1. Introduction

The presence of the coumarin structure in natural products and biologically active molecules has promoted considerable interest toward their synthesis [1,2,3]. The coumarin framework is widely used as a building block for obtaining derivatives that exhibit a wide variety of biological properties [4], especially anti-HIV, antitumor, anticoagulant, and antibiotic activities [5]. Moreover, some representatives of this class of compounds were screened as new drug candidates showing promising inhibition activity against the main protease of SARS-CoV-2 (PDB ID: 5N5O) [6,7].
Furthermore, the photophysical properties of coumarin-based fluorescent dyes have shown the advantage of using the benzopyran moiety as a fluorogenic scaffold [8]. Various coumarin-based compounds were applied in fields as laser dyes [9], cell-imaging biomarkers [10], and optical brighteners [11].
The synthesis of substituted coumarins is still dominated by classical methods, such as the Knoevenagel, Perkin, and Pechmann reactions [12,13,14], which although powerful and proven, have limited scope in terms of functional group compatibility. Recent studies have centered on the use of palladium-catalyzed cross-coupling C–C bond formation leading to the 3-, 4- and 6-substituted coumarins [15,16,17,18,19]; thus the number of Pd-catalyzed approaches for obtaining coumarins is constantly growing [20,21,22]. However, most of these synthetic protocols are focused on monosubstituted coumarins. Only limited applications of metal-catalyzed reactions for the synthesis of benzene ring-functionalized 3-substituted coumarins, especially when electron withdrawing groups are present, have been reported [23,24,25]. Therefore, herein we present an efficient method for the synthesis of fluorescent coumarins and phosphorous-containing coumarin-type heterocycles via palladium cross-coupling reactions, whereby the visible absorption properties of the compounds can be controlled by introducing various C-3 and C-6 substituents directly to the heteroaromatic structure.

2. Results and Discussion

2.1. Theoretical Investigation of Substituted 3-Phosphonocoumarin Structures—Evaluation of the Fluorescent Properties

As a part of our systematic investigation on the chemical behavior of 3-phosphonocoumarin, 1,2-benzoxaphosphorins as well as other 3-substituted coumarins, we theoretically studied the possibility to obtain substituted fluorescent species, where a “push–pull” effect with the substituent in the 3-rd position could be observed.
To check the photophysical properties and the possibility of fine tuning the spectral properties of phosphorous-containing model coumarins, quantum-chemical calculations of a series of phosphonocoumarins in acetonitrile as a solvent media were performed. For this purpose, density functional theory (DFT) [26,27,28,29] and time-dependent (TD) DFT [30,31] with the Gaussian16 suite of programs [32] were used.
The absorption and emission spectra of several compounds (CM-1—CM-12, Table 1), varying the substituent in position C-6 of the coumarin ring were calculated. Based on previous investigations in our group, coumarins bearing substituents in position C-6 [33] and C-7 [34] exhibit good fluorescent properties. The model compounds phenyl and aryl groups were chosen to enlarge the conjugated system, connected to the coumarin fragment. In order to take the solvent effect into account, an implicit solvent model (PCM) was used. The long-range corrected hybrid exchange-correlation functional CAM-B3LYP [35] paired with 6-31++G** basis set was employed [36].
The calculated absorption and emission energies of the phosphorous-containing coumarins were compared—structure CM-1 and models having either EDG (electron-donating group) or EWG (electron withdrawing group) in 6-th position of the coumarin ring (CM-2 to CM-12). The calculations showed that the chosen structures bearing different substituents red-shifted the absorption by 4–32 nm compared to the unsubstituted one (CM-1). This shift to a longer wavelength indicates that the differently substituted aryl groups have a significant impact on the absorbance of the coumarin derivatives.
The absorption spectra of structure CM-5 with a p-methoxyphenyl group in position C-6 is more red-shifted compared to its ortho-isomer (CM-3), 320 nm versus 312 nm, respectively. For the model-bearing methoxy group in meta position, CM-10, the value for λabs is 309 nm. The absorption and the emission wavelengths of the model C-10 are very similar to the non-substituted phenyl-bearing coumarin model CM-2. As expected, due to the lack of conjugation between the substituent and the coumarin moiety, an MeO-group in meta position would not affect the photochemical properties of the phosphorous-containing coumarin. This shows that not only the nature of the substituent in the phenyl ring is important, but also its position. It could be assumed that better conjugation occurs when there is an electron-donating group in para-position in the aryl substituent; therefore, the push–pull effect tends to be stronger.
In order to check whether a phenyl group (structure CM-2) could influence the spectral properties of the compound compared to the unsubstituted 3-phosphonocoumarin (CM-1), calculations of the absorption spectra data were performed. The λabs value for CM-2 compound is 14 nm red-shifted compared to the one for CM-1; the same trend was observed for the emission of these molecules, Table 1. Therefore, a scrutinous selectivity of the substituents could efficiently be used to tune the fluorescent properties of the coumarins.
The effect of EWG in the aryl substituent was also considered. Two models having -CN or -F group were tested. The spectral properties of structure CM-4, containing CN-group in ortho-position, were investigated and the presence of this group lowers the λabs value, Table 1. When the CN-group is at para or meta position, CM-11 and CM-12, the absorption maximum for both structures is slightly bathochromically shifted by 5 nm compared to the value for CM-4 model, which is 299 nm. Comparing the calculated absorption of CM-1 with CM-6, the 15 nm red-shift indicates that a minor change in the UV-spectra could be made when a fluorine atom is introduced in the aryl substituted.
Moreover, the effect of the alkyl groups in the aryl substituent (mesityl group), CM-7, was evaluated. The absorption wavelength of the compound is slightly red-shifted (only 4 nm) compared to CM-1. The presence of mesityl group (CM-7) would suppress the fluorescent properties compared to the phenyl substituent (CM-2), but still the emission is slightly red-shifted, compared to CM-1.
The possibility of having a longer conjugated system was also considered. Two compounds having triple C≡C bond were investigated (CM-8 and CM-9). Comparing their emission to the one of CM-1 (384 nm for CM-8, 402 nm for CM-9, and 344 nm for CM-1) indicates that the presence of an elongated π-system is beneficial to the fluorescent properties of the coumarin derivatives. In addition, (in CM-9) the presence of a methoxy group at para-position leads to higher λem compared to CM-8. Interestingly the emission for the model compounds CM-2 and CM-8 (having phenyl groups, 378 nm and 384 nm, respectively) and CM-5 and CM-9 (404 nm and 402 nm, respectively) slightly differ; therefore the addition of a triple bond does not affect the fluorescence properties of the coumarin species. The models bearing CN-group (CM-4, CM-11, and CM-12) exhibit emission maxima at a shorter wavelength (around 360 nm) compared to the structure having unsubstituted phenyl ring CM-2.
The calculated Stokes shifts for all the compounds were compared to the unsubstituted phosphorous coumarin CM-1 (49 nm), Table 1. The absolute values of the calculated shifts for all the substituted compounds (CM-2CM-9) are higher (64–84 nm) than the Stokes shift for CM-1. Compound CM-4 bearing -CN group in the aryl moiety has the lowest calculated Stokes shift (58 nm) in comparison to all of the substituted coumarin derivatives. The highest one was calculated for structure CM-5 that might be due to the EDG group in para-position in the aryl group.
The obtained theoretical results imply that the optical properties of phosphorous-containing coumarin systems can be tuned by altering the substituent in its C-6 position. These findings motivate us to synthesize the model structures that possess promising fluorescent properties and to experimentally check the spectral properties of the substituted 3-phosphonocoumarins.

2.2. Synthesis of Fluorescent Coumarins and Phosphorous Containing Coumarin-Type Heterocycles

To obtain the calculated substituted fluorescent species, where a “push-pull” effect with the substituent in 3-rd position could be observed, a Pd-catalyzed cross-coupling reaction and in particulate the Suzuki–Miyaura coupling, was investigated.
The Knoevenagel reaction, as one of the classical cyclization methods, was the first tested approach to obtain the desired structures. For this purpose, the phenyl substituted salicylaldehyde 2 had to be synthesized. One of the possible reaction paths includes adding the desired phenyl fragment via the Suzuki coupling starting from 5-bromo-2-hydroxybenzaldehyde 1, Scheme 1. The reaction of compound 1 with phenylboronic acid was performed by using sodium carbonate decahydrate as a base in mixed solvent media (toluene:ethanol:water) and was catalyzed via bis(triphenylphosphine)palladium(II) dichloride (2 mol%) in inert atmosphere for 24 h at 80 °C. The yield of the arylated product 2 was moderate—69%. This compound was previously synthesized with slightly better yield using different base and solvent media, however, applying 10 mol% of the palladium catalyst [37].
The isolated 2-hydroxy-5-phenylbenzaldehyde 2 was used as a starting material in the Knoevenagel reaction with triethyl phosphonoacetate. It should be mentioned that compound 2 is not very stable—it slowly oxidized in air, both at room temperature or when stored at 4–10 °C. The condensation reaction was carried out following the procedure developed in our group [38] applying piperidine as a catalyst in dry toluene at reflux, Scheme 2.
The reaction was completed in 5 h. Even though the reaction went smoothly, there were some difficulties related to the evaporation of the solvent and the separation of the two chemoisomers. The first obstacle was that compounds 3a and 4a have similar Rf-values on both silica and alumina (TLC-monitoring); therefore, it was difficult to separate the compounds to sufficient purity by column chromatography. On the other hand, both compounds are slowly solidifying oils which made further purification by recrystallization impracticable.
It is interesting to note that the phenyl substituted benzopyrans 3a and 4a possess fluorescent properties as calculated. Therefore, a parallel approach for obtaining these structures was planned based on the preparation of 6-bromo-3-phosphonocoumarin 5 by Knoevenagel condensation of 5-bromo-2-hydroxybenzaldehyde 1 with triethyl phosphonoacetate [38], Scheme 2, followed by a subsequent derivatization of the products by Suzuki coupling.
The advantages of this synthetic pathway are the usage of a stable and cheaper aldehyde and mainly the easier separation and further purification by column chromatography of products 5 and 6.
Thereafter, the reaction of 6-bromo-3-phosphonocoumarin 5 with phenylboronic acid, Scheme 3, was carried out using a relatively high load of palladium catalyst (3 mol%) due to the electron-donating oxygen atom from the lactone, bonded to the benzene ring in the coumarin moiety.
During the preliminary studies, some unexpected problems were faced due to the similar chromatographical behavior of compound 5 and the desired product 3a. This complicated the TLC monitoring of the reaction and the purification of diethyl 6-phenylcoumarin-3-phosphonate 3a. Formation of the dehalogenated coumarin structure 7 was also observed as a side product. Therefore, better reaction conditions needed to be developed for full conversion of 3b where no side reaction could be observed.
The optimal conditions for the palladium-catalyzed coupling were determined through a series of experiments in which different combinations of catalyst (Figure 1) and solvent were applied. Two groups of palladium catalysts were employed, differing by the Pd oxidation state—Pd(0) and Pd(II), and by the coordinated ligand—phosphine and NHC-type.
Due to the drawbacks mentioned above, a series of small-scale experiments were performed, Scheme 4, Table 2. The ratio of compounds 5, 3a, and 7 was monitored by NMR spectroscopy of the crude reaction mixture. Based on the chemical shift of the H-4 proton in all coumarin species 5, 3a, and 7, the ratio of the compounds was determined.
For better comparison of the obtained results, the same reaction time was implied for all of the listed small-scale reactions (20 h). The reactions were performed using 1.2 equiv. of phenylboronic acid under an argon atmosphere at 80 °C. According to the literature data [39,40,41], addition of water as a co-solvent or other protic solvent to the mixture might improve the outcome of the reaction; thus, several combinations of water or ethanol with dry toluene or dioxane were also tested. Potassium carbonate (3 equiv.) was tested as a base for the reaction with the coumarin species due to the possibility of opening the lactone ring if stronger bases were used [42,43]. Not only did the replacement of the sodium carbonate with potassium carbonate improve the yields significantly but also K2CO3 is nontoxic, cheap, and strong enough to activate [44,45,46,47,48] all of the used Pd(II) catalysts. The results of the initial experiments are summarized in Table 2.
The catalysts were introduced as a solution in THF (0.5 mL) except for PdCl2(PPh3)2. Due to its insolubility, this catalyst was introduced to the reaction in its crystalline form. To properly evaluate the obtained results for PdCl2(PPh3)2, 0.5 mL THF was additionally added to the reaction.
As we mentioned above, the reaction was performed in the presence of different types of palladium catalyst according to the ligands coordinated to the metal atom. The catalytic activity of the phosphine type complexes PdCl2(PPh3)2 and Pd(PPh3)4 in the studied reaction differed significantly. The usage of the PdCl2(PPh3)2 in the arylation reaction led to almost full conversion of the starting 6-bromo-3-phosphonocoumarin 5. As it could be seen from the listed results, Table 2, when employing bis(triphenylphosphine)palladium(II) dichloride, the reaction outcome is almost solvent-media-independent. However, not only the starting material but also the dehalogenated coumarin 7 was observed in the crude reaction mixture. The catalytic activity of tetrakis(triphenylphosphine)palladium(0) with the presented coumarin 5 was not satisfying—Pd(PPh3)4 complex gave the worst results amongst all of the chosen catalytic systems. However, when a mixed solvent (toluene:water or dioxane:water; entry 7 and entry 9, Table 2) was used, the yield of the desired product was increased.
The next set of experiments was performed by palladium complexes coordinating NHC-ligands. Two catalysts having bulkier ligands were tested in the Suzuki coupling—IMesPd(dmba)Cl and PEPPSI-type. The main observation in the IMesPd(dmba)Cl catalyzed reactions was the conversion and the yield of the targeted product. The selectivity of the reaction increased when water was introduced in the solvent media, entry 12 and entry 14, Table 2.
The full conversion of the 6-bromo-3-phosphocoumarin 5 into 6-phenyl-3-phosphonocoumarin 3a was observed only when the reaction was carried out under the conditions listed as entry 17 and entry 19 (Table 2) in the presence of the PEPPSI-type catalyst.
It is interesting to note that when the reaction was performed in ethanol as a solvent media, a full conversion of the starting product was observed; however, the formation of 3a was not detected. This statement was based on the NMR spectra where the characteristic signals for the protons at position C-4 for neither the 5, 3a, or 7 were found. This might be due to opening of the lactone ring in the presence of ethanol/base as it was previously reported [42,43,49,50,51,52].
Further scrutiny of the solvent media gave additional clarity of the overall conditions for better coupling reactivity. This made us consider the combination of PEPPSI-type catalyst and dioxane/water or toluene/water as the most effective in accomplishing high cross-coupling yield with full conversion of the starting material. The combination of toluene/water was preferred for the next Pd-catalyzed coupling reactions because of its low toxicity and price.
The structure of the 6-phenyl-3-phosphonocoumarin 3a was determined by a single crystal X-ray analysis (CCDC 2209998, Figure 2) where only one substance was identified in the crystalline matter.
The cross-coupling reactivity of 6-bromo-3-phosphonocoumarin 5 was further tested under the above-optimized conditions, Scheme 5. Aiming to increase the conjugation of the push–pull system, a series of boronic acids bearing electron donating and electron withdrawing groups were chosen, Table 3. The reactions were performed in five-time larger scales compared with the preliminary studies illustrated in Table 2.
Interestingly, under the optimized reaction conditions and in the presence of the selected boronic acid, the cross-coupling reaction with 6-bromo-3-phosphonocoumarin 5 could be achieved, Table 3, Scheme 5. In cases of unsubstituted or para-substituted boronic acids, the yields of the desired products 3a, 3c, and 3d varied from good to almost quantitative (61–95%) after purification by column chromatography. In all the presented experiments, the time needed for the full conversion of the coumarin 5 when substituted boronic reagents were used had increased compared with the unsubstituted phenylboronic acid. This implies that the hindrance of the reaction center and the electronic effect of the groups play a major role in the outcome of the reactions.
As a proof of that hypothesis, we observed that when the reaction center of the boronic reagent is hindered, the yield of the products 3b and 3e were low (entry 2 and entry 6, Table 3) using PEPPSI-type catalyst. The more hindered the boronic adduct is, the less yield is obtained. This observation is not new when a Suzuki reaction is performed with ortho or di-ortho substituted boronic species [53,54]; therefore, we decided to test other not so voluminous palladium catalysts that have shown good catalytic activity in the initial reaction with 5. The arylation was performed with PdCl2(PPh3)2 (entry 3 and entry 7) thus resulting in significantly increasing of the yields for product 3b and 3e, from 41% to 63% and from 19% to 45%, respectfully.
Even when using less hindered catalyst, entry 7, Table 3, the reaction with mesityl boronic acid gave the lowest yields in comparison with the other arylated products. Thence, an experiment for obtaining 3e with better yields was performed where the reaction temperature was increased to 100 °C; however, this did not give the positive outcome that we expected. The catalyst slowly decomposed, and only partial conversion of the starting material was indicated.
Phosphorus-containing coumarins such as 7 [1] are of a great importance in the areas of LiveScience due to the similarity of phosphorous compounds to the naturally occurring carboxylic acid derivatives and their potential application in varies biological systems. In order to further explore the generality of our modified procedure and to obtain analogous structures, the chemical behavior of halogenated ethyl and methyl ester of coumarin-3-carboxylic acid 8a,b was studied in the Suzuki reaction, Scheme 6.
The same reaction conditions—PEPPSI-type catalyst, mixed solvent (toluene:water) and potassium carbonate as a base—were employed in for the reactions with ethyl and methyl ester of coumarin-3-carboxylic acid 8a,b. A striking observation was the extreme reactivity of the bromo carboxylates in the Suzuki reaction catalyzed by the PEPPSI complex. For example, the time needed for full conversion of both starting materials, in case of unsubstituted or substituted in para-position boronic acid, was determined to be in the range of 30 to 60 min.
Longer reaction time and lower yields were observed again when 2-metoxyphenyl boronic acid was used as an adduct in the Suzuki reaction with coumarin 8a, entry 3, Table 4. The replacement of the bulky PEPPSI-type catalyst with less hindered complex as PdCl2(PPh3)2 resulted in increasing the yield from 20% to 68% (entry 4, Table 4). The usage of bis(triphenylphosphine)palladium(II) dichloride gave us the opportunity to obtain coumarin derivatives with even more hindered boronic acids—as 2,4,6-trimethylphenyl boronic acid and 2,6-dimethoxyphenyl boronic acid (entry 10–13, Table 4).
A weak reactivity of the coumarin species in the studied conditions in cases listed as entry 10–13 was observed. The obtained low yields for products 9e,f and 10e,f might be due not only to the fact that there was not full conversion of the esters but also to difficulties in their purification.
Longer reaction time and lower yields were observed again when 2-metoxyphenyl boronic acid was used as an adduct in the Suzuki reaction with coumarin 8a, entry 3, Table 4. The replacement of the bulky PEPPSI-type catalyst with less hindered complex as PdCl2(PPh3)2 resulted in increasing the yield from 20% to 68% (entry 4, Table 4). The usage of bis(triphenylphosphine)palladium(II) dichloride gave us the opportunity to obtain coumarin derivatives with even more hindered boronic acids—as 2,4,6-trimethylphenyl boronic acid and 2,6-dimethoxyphenyl boronic acid (entry 10–13, Table 4).
The synthesis of compounds 9a,c,d via the Suzuki reaction was previously reported by two different research collectives [23,24]; however, different catalytic systems were used. The obtained yields for these compounds were lower under longer reaction times. For example, Gobec et al. [24] obtained the ethyl 6-phenylcoumarin-3-carboxylate 9a with only 33% yield, while the reaction time was not specified (overnight); Carbonaro et al. [23] reported the formation of product 9a in 55% yield for 12–16 h. In the procedure that we developed, the ethyl 6-phenylcoumarin-3-carboxylate 9a was isolated with 90% yield after only 30 min, entry 1 (Table 4). The yield of compound 9c was reported to be 67% (overnight) [24] and 66% (12–16 h) [23]. Using the optimized procedure based on the PEPPSI-type catalyst, we managed to isolate the ethyl 6-(4-methoxyphenyl)coumarin-3-carboxylate 9c in 92% for 45 min, entry 6 (Table 4). The yield of 9d obtained from a Suzuki reaction catalyzed by Pd(PPh3)4 was reported [24] to be 54% (overnight). As can be seen from Table 4, changing the catalytic system could significantly increase the reactivity of the coumarin 8a, thus resulting in the 80% yield for coumarin 9d and the conversion time of 45 min, entry 8.
The reactivity of 6-bromocoumarin-3-carboxylates 8a,b was tested with even bulkier adduct—2,6-dimetoxyboronic acid, entry 12 and entry 13, Table 4. Unfortunately, in both cases, no full conversion of the starting compounds was observed. Even though the reaction times were long, a complexed mixture was obtained, and none of the structures were isolated due to the absence of predominant product. Another drawback was the similar chromatographic properties of some components of the mixture on both silica and alumina.
Motivated by these results, we next sought to expand the scope of the optimized procedure to the arylation of a more challenging substrate, such as 1,2-benzoxaphosphorin 6, Scheme 7.
Under the optimized reaction conditions—PEPPSI-type catalyst, solvent mixture of toluene:water, potassium carbonate and in the presence of the selected boronic acid—the cross-coupling reaction with ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphorine-3-carboxylate 6 was achieved, Scheme 7, Table 5. The yields for the arylated products 11a–d are similar to the ones obtained for the corresponding chemoisomeric structures 3a–d. However, the time needed for the consumption of the starting material compared with the 3-phosphonocoumarin derivatives was shorter.
A different catalyst was used, Pd(PPh3)2Cl2 (entry 3, Table 5), only when 2-methoxyphenylboronic acid was employed. The bis(triphenylphosphine)palladium(II) dichloride complex gave better results with the studied carboxylates and with 6-bromo-3-phosphonocoumarin 5. However, in the case of oxaphosphorine-3-carboxylate 6, the yield of product 11b was moderate (45%).
In the studied reaction conditions, poor reactivity of the coumarin species was observed with respect to hindered boronic acids; therefore, no experiments were performed with 2,4,6-trimethylphenylboronic acid and 2,6-dimethoxyphenylboronic acid. On the other hand, oxaphosphorine-3-carboxylate 6 is a byproduct, and there is still no efficient method for its preparation on a large scale.
The Suzuki reaction of ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphorine-3-carboxylate 6 is applicable for boronic acids bearing electron-donating and electron-withdrawing groups. The only limitation of the procedure is when a hindered organoboron substrate is used.
To extend the π-system of the heterocyclic system, another Pd-catalyzed reaction was tested on the studied coumarins—the Sonogashira alkynylation, Scheme 8, Table 6.
Most reactions with the chosen substrates were performed under classical conditions for Sonogashira coupling [55,56]—Pd(PPh3)2Cl2/CuI as a catalytic system in the presence of Et3N in DMF at 80 °C under argon atmosphere. To compensate the electron-rich nature of the 4-ethynylanisole, entry 4, Table 6, PEPPSI-type complex was applied. However, these conditions resulted in the formation of product 12d in only a 56% yield. Reaction of the same alkynylating reagent and ethyl coumarin-3-carboxylate 8a was not carried out since the product is a known compound and was prepared under the same classical conditions [57]. The Sonogashira reaction in analogues conditions involving the methyl carboxylate 8b had also been reported [58].
Surprisingly, the reactions of ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphorine-3-carboxylate 6 under the described conditions led to the formation of complexed reaction mixtures, and none of the structures 12c,e were isolated due to the absence of predominant product.

2.3. Photophysical Properties

The photophysical properties of the synthesized compounds were experimentally determined by UV-VIS and fluorescent spectroscopy. The absorption and fluorescence spectra of the aryl substituted coumarin, 3a, were recorded in five different solvents—acetonitrile (MeCN), methanol (MeOH), ethyl acetate (EtOAc), dimethyl sulfoxide (DMSO), and dichloromethane (DCM), Figure 3, Table 7. The obtained results showed that the solvent slightly affects the position of the longest absorption wavelength band. For example, in the polar protic solvent (MeOH) the absorbance maximum was exhibited at 346 nm, and in the non-polar one—dichloromethane was hypsochromically shifted only by 2 nm (Table 7 and Figure 3). In the emission spectra obtained, MeOH and CH2Cl2 bands at 459 nm and at 452 nm, respectively, were observed. Interestingly, although DMSO is an aprotic solvent, the maximum in the fluorescence spectrum was bathochromically shifted only by 1 nm compared to MeOH. The biggest difference was noted in the case of EtOAc—the maximum was blue-shifted by 16 nm with respect to the corresponding in DMSO, which was the most red-shifted one.
The spectral data of the other three coumarins bearing phenyl substituents—9a, 10a and 11a dissolved in acetonitrile, methanol and dichloromethane demonstrated minor solvatochromic effect from 13 to 16 nm—in MeOH the bands were red-shifted, while in DCM they were blue-shifted, and the maximum in acetonitrile was always between these two maxima (Table 7, Figure 3). Due to the higher solubility of the coumarins in acetonitrile compared to methanol, MeCN was selected as a solvent for all other synthesized derivatives.
The UV-VIS and emission spectra of 3b–e, 12a, and 12d compounds in acetonitrile are shown in Figure 4. As predicted by the DFT calculations, the presence of phenyl substituent in 6-th position of the coumarin ring (3a) leads to a redshift both in the absorption and the fluorescence spectra by 19 and 42 nm, respectively, compared to the unsubstituted substance. The absorption maximum in the UV spectrum of the 3e bearing mesityl group, was at shorter wavelength like in the simulated one—333 nm in comparison with the corresponding value for 3a—343 nm. The compounds having methoxy group in ortho-position and F-atom in the aryl substituent exhibited maximum at 345 nm, while for Ph-C≡C λmax was slightly red-shifted by 6 nm. The compound containing methoxyphenyl group in ortho-position in the aryl substituent emitted at 493 nm.
The derivative 12a with triple bond and without OMe group in para position exhibits a maximum at a shorter wavelength 461 nm. The largest bathochromic shift in the absorption and emission spectra was observed for the coumarin derivates with longer conjugated π-system, compounds 3c and 12d. They exhibited the same absorption maximum at 356 nm, which by 13 nm red-shifted compared to the corresponding spectra of the phenyl derivate, 3a, while their emission is shifted to the longest wavelength by 82 and 90 nm for 3c and 12d, respectively.
The same trend was observed for the ethyl and methyl coumarin-3-carboxylates and benzoxaphosphorines—the absorption and emission bands’ shift was highest when there is para-methoxyphenyl substituent (Figure 5) as the fluorescence is very similar to that of the corresponding phosphonocoumarins 3c and 12d—546, 543, and 541 nm for 9c, 10c, and 11c, respectively, Figure 6. Thus, the substituent in position C-3 does not affect the fluorescence of the coumarin species. The λmax in the fluorescence spectra of phenyl derivates is red-shifted by nearly 50 nm with respect to the unsubstituted coumarins, which for each series are denoted with the number of the parent series followed by H (Table 7).
The observed Stokes shift for all compounds with C-6 substituents was above 100 nm, while the highest was determined for the para-methoxyphenyl derivatives, and it varied from 179 for 3c to 191 for 11c. For the unsubstituted coumarins, the Stokes shift is smaller and is around 85 nm.
The computed maxima of the absorption spectra, multiplied by the scaling factor, is in very good agreement with the experimental data (Table 8). The differences do not exceed 5 nm. The order of λmax in the calculated UV spectra is the same as in the experimentally obtained—in direction H-, mesityl-, Ph-, ortho-MeOC6H4-, F-, Ph-C≡C-, para-MeOC6H4-, and para-MeOC6H4-C≡C substituents at position C-6 is the observed bathochromic shift.
The variance for the fluorescence bands in some cases is much more pronounced by up to 62 nm for coumarin 12d. However, in the results for 3a, 3e, and 12a, the simulated data are red-shifted by only 1, 7, and 2 nm, respectively. The calculated Stokes shift differs by around 40–45 nm. As for the two structures with methoxy group in para position, 3c and 12d, the distinction is higher and is 95 and 112 nm, respectively.
To test the fluorescent properties of coumarins 3a and 3c in different pH, a series of experiments by varying the concentration of sulfuric acid and sodium hydroxide (Figure 7) were performed. Interestingly, the fluorescence of the para-methoxyphenyl derivative 3c was suppressed when lowering the pH of the solution by adding sulfuric acid, while the emission of the phenyl derivative 3a had increased. This different behavior of the coumarin species might be due to different protonation positions. In compound 3a, the protonation could occur at the lactone oxygen atom therefore the conjugation would be more effective. The quenching of the fluorescence of coumarin 3c might be due to the protonation of the MeO-group which decreases the donor ability of the oxygen atom. The addition of sodium hydroxide might lead to lactone ring opening, thus loss of conjugation (Figure 7c,d).

3. Materials and Methods

Melting points were determined on an SRS MPA120 EZ-Melt apparatus and are used without correction. The IR spectra were recorded with a Shimadzu FTIR-8400S spectrophotometer. 1H, 13C, and 31P NMR spectra were recorded on a Bruker AVNEO 400 spectrometer (at 400 MHz for 1H, 100.6 MHz for 13C, 376.46 MHz for 19F, and 161.98 MHz for 31P, respectively), Bruker Avance III 500 spectrometer (at 500 MHz for 1H, 125.7 MHz for 13C and 202.4 MHz for 31P, respectively). Chemical shifts are given in ppm from tetramethylsilane as internal standard with CDCl3 as solvent. Chemical shifts in 19F NMR referenced by IUPAC recommendations (2000) and in 31P NMR spectra were referred from external standard—85% H3PO4. The NMR-spectra of the newly synthesized compounds could be found in the Supplementary Materials. Liquid chromatography mass spectrometry analysis (LC-HRAM) was carried out on Q Exactive® hybrid quadrupole-Orbitrap® mass spectrometer (ThermoScientific Co, Waltham, MA, USA) equipped with a HESI® (heated electrospray ionization) module, TurboFlow® Ultra High Performance Liquid Chromatography (UHPLC) system (ThermoScientific Co, Waltham, MA, USA) and HTC PAL® autosampler (CTC Analytics, Zwingen, Switzerland). The chromatographic separations of the analyzed compounds were achieved on Nucleoshell C18 (100 × 2.1 mm, 2.7 µm) analytical column (Macherey-Nagel, Düren, Germany) using gradient elution at 300 µL/min flow rate. The used eluent systems were: A—0.1% formic acid in water; B—0.1% formic acid in CH3CN. Full-scan mass spectra over the m/z range 100–600 were acquired in positive ion mode at resolution settings of 140,000. The used mass spectrometer operating parameters were: spray voltage—4.0 kV; capillary temperature—320 °C; probe heater temperature—300 °C; sheath gas flow rate 40 units; auxiliary gas flow 12 units; sweep gas 2 units (units refer to arbitrary values set by the Q Exactive Tune software); and S-Lens RF level of 50.00. Nitrogen was used for sample nebulization and collision gas in the HCD cell. All derivatives were quantified using 5 ppm mass tolerance filters to their theoretical calculated m/z values. Data acquisition and processing were carried out with XCalibur® ver 2.4 software package (ThermoScientific Co, Waltham, MA, USA). UV-Vis spectra were carried out on a Shimadzu UV-1800. Fluoresce spectra were recorded at room temperature on a PerkinElmer LS45. Reactions were monitored by TLC on silica gel 60 F254. Column chromatography was carried out on silica gel (Merck 0.043–0.063 mm) using as eluent n-hexane/EtOAc mixture with increasing polarity. The X-ray data set was collected using a Bruker D8 Venture diffractometer with a microfocus sealed tube and a Photon II detector. Monochromated Mo radiation (λ = 0.71073 Å) was used. Data were collected at 133(2) K and corrected for absorption effects using the multi-scan method. The structure was solved by direct methods using SHELXT [59] and was refined by full matrix least squares calculations on F2 (SHELXL2018 [60]) in the graphical user interface Shelxle [61]. Refinement—all non H-atoms were located in the electron density maps and refined anisotropically. C-bound H atoms were placed in positions of optimized geometry and treated as riding atoms. Their isotropic displacement parameters were coupled to the corresponding carrier atoms by a factor of 1.2 (CH, CH2) or 1.5 (CH3). Disorder: The ethoxy-group on O4 was split over two positions. Its occupancy factors refined to 0.775 for the major compound.
Computational details—the quantum-chemical calculations of the series of phosphonocoumarins in acetonitrile were performed using density functional theory (DFT) [26,27,28,29], time-dependent (TD) DFT [30,31] with Gaussian16 suite of programs [32] and the long range corrected hybrid exchange-correlation functional CAM-B3LYP [35], paired with 6-31++G** basis set was employed and a polarizable continuum model (PCM) was used [36].
All chemical reagents were purchased from Merck and Sigma Aldrich. The starting 6-bromo-3-substitueted-2-oxo-2H-1-benzopyrans 5,6,8a,b were prepared according to reported procedures [38]. The used Pd-complexes (PdCl2(PPh3)2 [62], IMesPd(dmba)Cl [45], and PEPPSI-type catalyst [44]) were prepared according to known procedures.

3.1. Starting Materials

  • 4-Hydroxy-[1,1’-biphenyl]-3-carbaldehyde, 2
Compound 2 was prepared according to the reported procedure [37].
1H NMR (500 MHz, CDCl3) δ = 10.926 (s, 1H, OH), 9.889 (s, 1H, CHO), 7.672–7.781 (m, 2H, aromatic), 7.171–7.573 (m, 5H, aromatic), 7.000 (d, J = 8.4 Hz, 1H, aromatic).
  • Diethyl (6-bromo-2-oxo-2H-chromen-3-yl)phosphonate, 5
Compound 5 was prepared according to the reported procedure [38].
1H NMR (500 MHz, CDCl3) δ = 8.425 (d, J = 17.1 Hz, 1H, H-4), 7.714–7.728 (m, 2H, H-7, H-5), 7.26 (d, J = 9.4 Hz, 1H, H-8), 4.215–4.347 (m, 4H, two CH3CH2O), 1.387 (t, J = 7.1 Hz, 6H, two CH3CH2O).
  • Ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphinine-3-carboxylate 2-oxide, 6
Compound 6 was prepared according to the reported procedure [38].
1H NMR (500 MHz, CDCl3) δ = 8.132 (d, J = 36.9 Hz, 1H, H-4), 7.604 (dd, J = 8.6 Hz, 2.1 Hz, 1H, H-7), 7.576 (d, J = 2.1 Hz, 1H, H-5), 7.267 (d, J = 8.5 Hz, 1H, H-8), 4.323–4.476 (m, 4H, two CH2O), 1.412 (t, J = 7.1 Hz, 3H, CH3CH2OP), 1.391 (t, J = 7.1 Hz, 3H, CH3CH2O).

3.2. Procedure for Suzuki Reaction—Preliminary Study

3.2.1. Procedure for Suzuki Reaction with PEPPSI, Pd(PPh3)4, and IMesPd(dmba)Cl

In a Schlenk tube 0.108 g (0.3 mmol, 1 equiv) of diethyl (6-bromo-2-oxo-2H-chromen-3-yl)phosphonate 5, phenylboronic acid (0.044 g, 0.36 mmol, 1.2 equiv), finely powdered K2CO3 (0.124 g, 0.9 mmol, 3 equiv) were placed. A mixture of water:toluene (0.6 mL:2.4 mL), toluene (3 mL), water:dioxane (0.6 mL:2.4 mL), dioxane (3 mL), ethanol:water:toluene (0.5 mL:1 mL:1 mL), or ethanol:water (2 mL:1 mL) were added. The vessel was evacuated under vacuum and refilled with argon four times. After that, the respective catalyst (PEPPSI-type, Pd(PPh3)4, and IMesPd(dmba)Cl) (3 mol %) was introduced as 0.5 mL THF solution. The resulting suspension was heated at 80 °C for 20 h. After that, the reaction mixture was cooled to room temperature, diluted with CHCl3, dried with Na2SO4 if water was used, and filtered through a celite pad. Part of the solution (about 3 mL) was evaporated under reduced pressure; the residue was dissolved in CDCl3 and was analyzed by NMR spectroscopy.

3.2.2. Procedure for Suzuki Reaction with Pd(PPh3)2Cl2

In a Schlenk tube, 0.108 g (0.3 mmol, 1 equiv) of diethyl (6-bromo-2-oxo-2H-chromen-3-yl)phosphonate 5, phenylboronic acid (0.044 g, 0.36 mmol, 1.2 equiv), finely powdered K2CO3 (0.124 g, 0.9 mmol, 3 equiv) were placed. A mixture of water:toluene (0.6 mL:2.4 mL), toluene (3 mL), water:dioxane (0.6 mL:2.4 mL), dioxane (3 mL), ethanol:water:toluene (0.5 mL:1 mL:1 mL) or ethanol:water (2 mL:1 mL) were added. After that, the catalyst Pd(PPh3)2Cl2 was introduced as solid, and 0.5 mL of THF were additionally added. The vessel was evacuated under vacuum and refilled with argon four times. The resulting suspension was heated at 80 °C for 20 h. After that, the reaction mixture was cooled to room temperature, diluted with CHCl3, dried with Na2SO4 if water was used, and filtered through a celite pad. Part of the solution (about 3 mL) was evaporated under reduced pressure. The residue was dissolved in CDCl3 and was analyzed by NMR spectroscopy.

3.3. General Procedure for the Preparative Suzuki Reaction

In a 50 mL Schlenk flask, a mixture of the respective brominated heterocycle 1.5 mmol (1 equiv), boronic acid 1.8 mmol (1.2 equiv), potassium carbonate 0.622 g (4.5 mmol, 3 equiv), and palladium complex 45 µmol (3 mol%)–29 mg PEPPSI or 32 mg Pd(PPh3)2Cl2, was placed, and the respective solvent was added—3 mL:12 mL water:toluene when a PEPPSI-type catalyst was used, or 15 mL toluene for Pd(PPh3)2Cl2. The flask was evacuated and refilled with argon (four times). The reaction was heated at 80 oC until the coumarin derivative was consumed (TLC-monitoring). The needed time for the reactions is given in Table 2, Table 3 and Table 4. The reaction was quenched by adding water to the reaction, and the organic layer was extracted with chloroform (5 × 20 mL). The organic extracts were dried with anhydrous sodium sulfate. After the evaporation of the solvent, the residue was purified by column chromatography using n-hexane/EtOAc, n-hexane/CH2Cl2, or n-hexane/Et2O as gradient eluent system.
  • Diethyl (2-oxo-6-phenyl-2H-chromen-3-yl)phosphonate, 3a
The product was prepared from diethyl (6-bromo-2-oxo-2H-chromen-3-yl)phosphonate 5 (0.540 g), phenylboronic acid (0.220 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 17 h. The product was purified by column chromatography using n-hexane/Et2O, 0.482 g, 90%, white powder, m.p. = 146.3–149.3 °C. IR (nujol): ν = 1738, 1240, 1052, 1037 cm−1.
Molecules 27 07649 i048
1H NMR (500 MHz, CDCl3) δ = 8.509 (d, J = 17.1 Hz, 1H, H-4), 7.780 (dd, J = 8.6 Hz, 2.1 Hz, 1H, H-7), 7.690 (d, J = 2.1 Hz, 1H, H-5), 7.498 (m, 2H, H-4′, H-8), 7.413 (m as t, 2H, H-2′ and H-6′), 7.337 (m as q, 2H, H-3′and H-5′), 4.153–4.283 (m, 4H, two CH3CH2O), 1.330 (t, J = 7.1 Hz, 6H, two CH3CH2O);13C NMR (125.7 MHz, CDCl3) δ = 158.24 (d, JCP = 15.1 Hz, C-2), 154.59 (s, C-8a), 153.46 (d, JCP = 6.5 Hz, C-4), 138.91 (s, C-6), 138.38 (s, C-1′), 133.14 (s, C-7), 129.15 (s, C-2′ and C-6′), 128.08 (s, C-5), 127.33 (C-4′), 127.03 (s, C-3′ and C-5′), 118.27 (d, JCP = 196.2 Hz, C-3), 118.18 (d, JCP = 14.3 Hz, C-4a), 117.29 (s, C-8), 63.46 (d, JCP = 6.0 Hz, two CH3CH2O), 16.41 (d, JCP = 6.0 Hz, two CH3CH2O); 31P NMR (161.98 MHz, CDCl3) δ = 10.765.
HRMS (ESI) m/z calculated for C19H19O5P [M+H]+ 359.10429 found 359.10413 (ppm: 0.45).
  • Diethyl (6-(2-methoxyphenyl)-2-oxo-2H-chromen-3-yl)phosphonate, 3b
The product was prepared from diethyl (6-bromo-2-oxo-2H-chromen-3-yl)phosphonate 5 (0.540 g), 2-methoxyphenylboronic acid (0.273 g), K2CO3 (0.622 g) and Pd(PPh3)2Cl2 (32 mg) in 15 mL toluene for 48 h. The product was purified by column chromatography using n-hexane/Et2O, 0.366 g, 63%, yellow slowly solidifying oil. IR (nujol): ν = 1739, 1243, 1056, 1024 cm−1.
Molecules 27 07649 i049
1H NMR (400 MHz, CDCl3) δ = 8.556 (d, J = 17.2 Hz, 1H, H-4), 7.805 (dd, J = 8.6 Hz, 2.2 Hz, 1H, H-7), 7.760 (d, J = 2.0 Hz, 1H, H-5), 7.392 (d, J = 8.5 Hz, 1H, H-8), 7.372 (dd, J = 7.5 Hz, 1.1 Hz, 1H, H-3′), 7.308 (dd, J = 5.6 Hz, 1.7 Hz, 1H, H-5′), 7.065 (dd, J = 7.5 Hz, 1.1 Hz, 1H, H-4′), 7.020 (dd, J = 8.3 Hz, 0.7 Hz, 1H, H-6′), 4.203–4.358 (m, 4H, two CH3CH2O), 3.835 (s, 3H, CH3O), 1.391 (t, J = 7.1 Hz, 6H, two CH3CH2O);13C NMR (100 MHz, CDCl3) δ = 158.49 (d, JCP = 14.4 Hz, C-2), 156.35 (s, C-8a), 154.25 (s, C-1′), 153.99 (d, JCP = 6.7 Hz, C-4), 135.77 (s, C-7), 135.54 (s, C-2′), 130.57 (s, C-3′), 129.99 (s, C-5′), 129.55 (s, C-5), 128.14 (s, C-6), 121.10 (C-4′), 117.62 (d, JCP = 14.3 Hz, C-4a), 117.51 (d, JCP = 196.2 Hz, C-3), 116.43 (s, C-8), 111.34 (s, C-6′) 63.43 (d, JCP = 6.2 Hz, two CH3CH2O), 55.54 (s, CH3O), 16.39 (d, JCP = 6.0 Hz, two CH3CH2O); 31P NMR (161.98 MHz, CDCl3) δ = 11.141.
HRMS (ESI) m/z calculated for C20H21O6P [M+H]+ 389.11485 found 389.11511 (ppm: −0.67).
  • Diethyl (6-(4-methoxyphenyl)-2-oxo-2H-chromen-3-yl)phosphonate, 3c
The product was prepared from diethyl (6-bromo-2-oxo-2H-chromen-3-yl)phosphonate 5 (0.540 g), 4-methoxyphenylboronic acid (0.273 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 48 h. The product was purified by column chromatography using CH2Cl2/Et2O, 0.551 g, 95%, pale yellow powder, m.p. = 111.6–112.5 °C IR (nujol): ν = 1740, 1242, 1047, 1025 cm−1.
Molecules 27 07649 i050
1H NMR (400 MHz, CDCl3) δ = 8.494 (d, J = 17.2 Hz, 1H, H-4), 7.739 (dd, J = 8.6 Hz, 2.2 Hz, 1H, H-7), 7.634 (d, J = 2.2 Hz, 1H, H-5), 7.328 (d, J = 8.6 Hz, 1H, H-8), 7.429 (dd, J = 8.8 Hz, 3.1 Hz, 2H, H-3′ and H-5′), 6.937 (dd, J = 8.8 Hz, 3.7 Hz, 2H, H-2′ and H-6′), 4.137–4.290 (m, 4H, two CH3CH2O), 3.793 (s, 3H, CH3O), 1.321 (t, J = 7.1 Hz, 3H, CH3CH2O) 1.319 (t, J = 7.1 Hz, 3H, CH3CH2O);13C NMR (100 MHz, CDCl3) δ = 159.73 (C-4′), 158.31 (d, JCP = 14.4 Hz, C-2), 154.19 (s, C-8a), 153.54 (d, JCP = 6.5 Hz, C-4), 138.03 (s, C-6), 132.78 (s, C-7), 131.35 (s, C-1′), 128.11 (s, C-6′and C-2′), 126.71 (s, C-5), 118.16 (d, JCP = 14.3 Hz, C-4a), 118.10 (d, JCP = 196.2 Hz, C-3), 117.19 (s, C-8), 114.57 (s, C-3′ and C-5′), 63.44 (d, JCP = 6.0 Hz, two CH3CH2O), 55.42 (s, CH3O), 16.40 (d, JCP = 6.2 Hz, two CH3CH2O); 31P NMR (161.98 MHz, CDCl3) δ = 10.879.
HRMS (ESI) m/z calculated for C20H21O6P [M+H]+ 389.11485 found 389.11494 (ppm: −0.23).
  • Diethyl (6-(4-fluorophenyl)-2-oxo-2H-chromen-3-yl)phosphonate, 3d
The product was prepared from diethyl (6-bromo-2-oxo-2H-chromen-3-yl)phosphonate 5 (0.540 g), 4-fluorophenylboronic acid (0.251 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 48 h. The product was purified by column chromatography using CH2Cl2, 0.343 g, 61%, pale yellow powder, m.p. = 115.0–118.2 °C IR (nujol): ν = 1734, 1251, 1057, 1027 cm−1.
Molecules 27 07649 i051
1H NMR (400 MHz, CDCl3) δ = 8.495 (d, J = 17.2 Hz, 1H, H-4), 7.732 (dd, J = 8.6 Hz, 2.2 Hz, 1H, H-7), 7.641 (d, J = 2.2 Hz, 1H, H-5), 7.466 (ddq, JHF = 8.8 Hz, 5.0 Hz, JHP = 8.8 Hz, 3.1 Hz, 2H, H-3′ and H-5′), 7.356 (d, J = 8.6 Hz, 1H, H-8), 7.102 (ddq, JHF = 8.6 Hz, 3.1 Hz, JHP = 8.6 Hz, 3.1 Hz, 2H, H-2′ and H-6′), 4.143–4.296 (m, 4H, two CH3CH2O), 1.324 (t, J = 7.1 Hz, 3H, CH3CH2O) 1.323 (t, J = 7.1 Hz, 3H, CH3CH2O);13C NMR (100 MHz, CDCl3) δ = 162.84 (d, JCF = 248.0 Hz, C-4′), 158.16 (d, JCP = 14.4 Hz, C-2), 154.54 (s, C-8a), 153.26 (d, JCP = 6.5 Hz, C-4), 137.39 (s, C-6), 135.07 (d, JCP = 3.3 Hz, C-1′), 132.94 (s, C-7), 128.71 (d, JCP = 10.5 Hz, C-6′and C-2′), 127.17 (s, C-5), 118.50 (d, JCP = 196.7 Hz, C-3), 118.21 (d, JCP = 15.5 Hz, C-4a), 117.36 (s, C-8), 116.10 (d, JCP = 21.7 Hz, C-3′ and C-5′), 63.47 (d, JCP = 6.0 Hz, two CH3CH2O), 16.39 (d, JCP = 6.0 Hz, two CH3CH2O); 19F NMR (376.46 MHz, CDCl3) δ = -114.18; 31P NMR (161.98 MHz, CDCl3) δ = 10.637.
HRMS (ESI) m/z calculated for C19H18FO5P [M+H]+ 377.09486 found 377.09473 (ppm: 0.34).
  • Diethyl (6-mesityl-2-oxo-2H-chromen-3-yl)phosphonate, 3e
The product was prepared from diethyl (6-bromo-2-oxo-2H-chromen-3-yl)phosphonate 5 (0.540 g), 2,4,6-trimethylphenylboronic acid (0.294 g), K2CO3 (0.622 g) and Pd(PPh3)2Cl2 (32 mg) in 15 mL toluene for 48 h. The product was purified by column chromatography using n-hexane/Et2O, 0.270 g, 45%, white powder, m.p. = 130.0–132.6 °C. IR (nujol): ν = 1744, 1242, 1053, 1013 cm−1.
Molecules 27 07649 i052
1H NMR (400 MHz, CDCl3) δ = 8.535 (d, J = 17.2 Hz, 1H, H-4), 7.428(s, 1H, H-5), 7.425 (s, 1H, H-8), 7.376 (dd as t, J = 1.5 Hz, 1H, H-7), 6.972 (d, JHP = 0.4 Hz, 2H, H-3′ and H-5′), 4.215–4.367 (m, 4H, two CH3CH2O), 2.345 (s, 3H, CH3), 1.993 (s, 6H, two CH3), 1.397 (t, J = 7.1 Hz, 6H, two CH3CH2O);13C NMR (100 MHz, CDCl3) δ = 158.34 (d, JCP = 14.2 Hz, C-2), 154.12 (s, C-8a), 153.58 (d, JCP = 6.7 Hz, C-4), 138.10 (s, C-6), 137.55 (s, C-5), 136.49 (s, C-1′), 135.89 (s, C-6′and C-2′), 135.70 (d, JCF = 248.0 Hz, C-4′), 129.74 (s, C-7), 128.40 (d, JCP = 21.7 Hz, C-3′ and C-5′), 118.02 (d, JCP = 14.3 Hz, C-4a), 118.01 (d, JCP = 196.1 Hz, C-3),116.96 (s, C-8), 63.44 (d, JCP = 5.9 Hz, two CH3CH2O), 21.03 (s, CH3), 20.76 (s, two CH3), 16.40 (d, JCP = 6.3 Hz, two CH3CH2O); 31P NMR (161.98 MHz, CDCl3) δ = 10.984.
HRMS (ESI) m/z calculated for C22H25O5P [M+H]+ 401.15124 found 401.15143 (ppm: −0.47).
  • Ethyl 2-oxo-6-phenyl-2H-chromene-3-carboxylate, 9a
The product was prepared from ethyl 6-bromo-2-oxo-2H-chromene-3-carboxylate 8a (0.444 g), phenylboronic acid (0.219 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 30 min. The product was purified by column chromatography using n-hexane/Et2O and then CH2Cl2, 0.397 g, 90%, pale yellow powder, m.p. = 150.4–151.5 °C. IR (nujol): ν = 1758, 1694 cm−1.
Molecules 27 07649 i053
1H NMR (400 MHz, CDCl3) δ = 8.510 (s, 1H, H-4), 7.783 (dd, J = 10.8 Hz, 6.4 Hz, 1H, H-7), 7.709 (d, J = 2.2 Hz, 1H, H-5), 7.503 (dd as dt, J = 7 Hz, 1.5 Hz, 2H, H-2′ and H-6′), 7.411 (ddd as dt, J = 7.4 Hz, 1.5 Hz, 2H, H-3′and H-5′), 7.352 (d, J = 8.6 Hz, 1H, H-8), 7.326 (tt, J = 7.2 Hz, 1.3 Hz, 1H, H-4′), 4.356 (q, J = 7.1 Hz, 2H, CH3CH2O), 1.348 (t, J = 7.1 Hz, 3H, CH3CH2O); 13C NMR (100 MHz, CDCl3) δ = 163.06 (s, COOEt), 156.69 (s, C-2), 154.49 (s, C-8a), 148.65 (s, C-4), 138.92 (s, C-6), 138.29 (s, C-1′), 133.29 (s, C-7), 129.14 (s, C-3′ and C-5′), 128.07 (s, C-4′), 127.47 (s, C-5), 127.04 (s, C-2′ and C-6′), 118.66 (s, C-3), 118.12 (s, C-4a), 117.19 (s, C-8), 62.05 (s, CH3CH2O), 14.26 (s, CH3CH2O).
HRMS (ESI) m/z calculated for C18H14O4 [M+H]+ 295.09649 found 295.09616 (ppm: 1.12).
  • Ethyl 6-(2-methoxyphenyl)-2-oxo-2H-chromene-3-carboxylate, 9b
The product was prepared from ethyl 6-bromo-2-oxo-2H-chromene-3-carboxylate 8a (0.444 g), 2-methoxyphenylboronic acid (0.273 g), K2CO3 (0.622 g) and Pd(PPh3)2Cl2 (32 mg) in 15 mL toluene for 24 h. The product was purified by column chromatography using n-hexane/CH2Cl2 and then CH2Cl2, 0.331 g, 68%, yellow powder, m.p. = 120.0–124.2 °C. IR (nujol): ν = 1773, 1712 cm−1.
Molecules 27 07649 i054
1H NMR (400 MHz, CDCl3) δ = 8.566 (s, 1H, H-4), 7.811 (dd, J = 8.6 Hz, 2.2 Hz, 1H, H-7), 7.757 (d, J = 2.1 Hz, 1H, H-5), 7.387 (d, J = 8.6 Hz, 1H, H-8), 7.374 (td, J = 5.6 Hz, 1.7 Hz, 1H, H-5′), 7.313 (td, J = 7.5 Hz, 1.7 Hz, 1H, H-3′), 7.064 (td, J = 7.5 Hz, 1.1 Hz, 1H, H-4′), 7.020(dd, J = 8.3 Hz, 0.8 Hz, 1H, H-6′), 4.418 (q, J = 7.1 Hz, 2H, CH3CH2O), 3.834 (s, 3H, CH3O), 1.421 (t, J = 7.1 Hz, 3H, CH3CH2O); 13C NMR (100 MHz, CDCl3) δ = 163.20 (s, COOEt), 156.91 (s, C-2), 156.33 (s, C-1′), 154.15 (s, C-8a), 148.99 (s, C-4), 135.93 (s, C-7), 135.48 (s, C-2′), 130.59 (s, C-3′), 130.05 (s, C-5), 129.54 (s, C-5′), 128.20 (s, C-6), 121.10 (s, C-4′), 118.16 (s, C-3), 118.12 (s, C-4a), 117.58 (s, C-8), 111.34 (s, C-6′), 61.98 (s, CH3CH2O), 55.56 (s, CH3O), 14.26 (s, CH3CH2O).
HRMS (ESI) m/z calculated for C19H16O5 [M+H]+ 325.10705 found 325.1069 (ppm: 0.46).
  • Ethyl 6-(4-methoxyphenyl)-2-oxo-2H-chromene-3-carboxylate, 9c
The product was prepared from ethyl 6-bromo-2-oxo-2H-chromene-3-carboxylate 8a (0.444 g), 4-methoxyphenylboronic acid (0.273 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 45 min. The product was purified by column chromatography using n-hexane/CH2Cl2 and then CH2Cl2, 0.444 g, 92%, yellow powder, m.p. = 144.8–146.2 °C. IR (nujol): ν = 1761, 1689 cm−1.
Molecules 27 07649 i055
1H NMR (400 MHz, CDCl3) δ = 8.572 (s, 1H, H-4), 7.816 (dd, J = 8.6 Hz, 2.2 Hz, 1H, H-7), 7.726 (d, J = 2.2 Hz, 1H, H-5), 7.398 (d, J = 9.3 Hz, 1H, H-8), 7.507 (dq, J = 8.8 Hz, 3.1 Hz, 2H, H-3′ and H-5′), 7.009 (dq, J = 8.8 Hz, 3.1 Hz, 2H, H-4′ and H-6′), 4.419 (q, J = 7.1 Hz, 2H, CH3CH2O), 3.867 (s, 3H, CH3O), 1.422 (t, J = 7.2 Hz, 3H, CH3CH2O); 13C NMR (100 MHz, CDCl3) δ = 163.11 (s, COOEt), 159.72 (s, C-4′), 156.77 (s, C-2), 154.10 (s, C-8a), 148.73 (s, C-4), 137.95 (s, C-6), 132.93 (s, C-7), 131.36 (s, C-1′), 128.10 (s, C-3′ and C-5′), 126.83 (s, C-5), 118.54 (s, C-3), 118.10 (s, C-4a), 117.09 (s, C-8), 114.56 (s, C-2′ and C-6′), 62.02 (s, CH3CH2O), 55.42 (s, CH3O), 14.25 (s, CH3CH2O).
HRMS (ESI) m/z calculated for C19H16O5 [M+H]+ 325.10705 found 325.10675 (ppm: 0.92).
  • Ethyl 6-(4-fluorophenyl)-2-oxo-2H-chromene-3-carboxylate, 9d
The product was prepared from ethyl 6-bromo-2-oxo-2H-chromene-3-carboxylate 8a (0.444 g), 4-fluorophenylboronic acid (0.251 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 45 min. The product was purified by column chromatography using CH2Cl2, 0.374 g, 80%, pale yellow powder, m.p. = 180.3–184.7 °C. IR (nujol): ν = 1742, 1711 cm−1.
Molecules 27 07649 i056
1H NMR (400 MHz, CDCl3) δ = 8.577 (s, 1H, H-4), 7.808 (dd, J = 8.6 Hz, 2.2 Hz, 1H, H-7), 7.738 (d, J = 2.2 Hz, 1H, H-5), 7.543 (ddq, JHF = 8.9 Hz, 5.2 Hz; J = 8.9 Hz, 3.2 Hz, 2H, H-3′ and H-5′), 7.423 (d, J = 8.6 Hz, 1H, H-8), 7.185 (ddq as tq, JHF = 8.6 Hz, 3.1 Hz; J = 8.6 Hz, 3.1 Hz, 2H, H-2′ and H-6′), 4.432 (q, J = 7.1 Hz, 2H, CH3CH2O), 1.423 (t, J = 7.1 Hz, 3H, CH3CH2O); 13C NMR (100 MHz, CDCl3) δ = 163.02 (s, COOEt), 162.84 (d, J = 248.0 Hz, C-4′), 156.60 (s, C-2), 154.45 (s, C-8a), 148.50 (s, C-4), 137.32 (s, C-6), 135.09 (d, J = 3.2 Hz, C-1′), 133.10 (s, C-7), 128.71 (d, J = 8.2 Hz, C-2′ and C-6′), 127.32 (s, C-5), 118.81 (s, C-3), 118.14 (s, C-4a), 117.27 (s, C-8), 116.20 (d, J = 21.7 Hz, C-3′ and C-5′), 62.09 (s, CH3CH2O), 14.24 (s, CH3CH2O); 19F NMR (376.46 MHz, CDCl3) δ = -114.196.
HRMS (ESI) m/z calculated for C18H13FO4 [M+H]+ 313.08706 found 313.08682 (ppm: 0.77).
  • Ethyl 6-mesityl-2-oxo-2H-chromene-3-carboxylate, 9e
The product was prepared from ethyl 6-bromo-2-oxo-2H-chromene-3-carboxylate 8a (0.444 g), 2,4,6-trimethylphenylboronic acid (0.294 g), K2CO3 (0.622 g) and Pd(PPh3)2Cl2 (32 mg) in 15 mL toluene for 72 h. The product was purified by column chromatography using n-hexane/Et2O, 0.147 g, 29%, pale yellow powder, m.p. = 142.8–146.2 °C. IR (nujol): ν = 1758, 1691 cm−1.
Molecules 27 07649 i057
1H NMR (400 MHz, CDCl3) δ = 8.518 (s, 1H, H-4), 7.427 (s, 1H, H-8), 7.424 (s, 1H, H-5), 7.386 (dd as t, J = 1.2 Hz, 1H, H-7), 6.968 (d, J = 0.4 Hz, 1H, 2H, H-3′ and H-5′), 4.420 (q, J = 7.1 Hz, 2H, CH3CH2O), 2.343 (s, 3H, CH3Ar), 1.993 (s, 6H, CH3Ar), 1.416 (t, J = 7.1 Hz, 3H, CH3CH2O); 13C NMR (100 MHz, CDCl3) 163.16 (s, COOEt), 156.84 (s, C-2), 154.01 (s, C-8a), 148.63 (s, C-4), 138.04 (s, C-6), 137.54 (s, C-5), 136.51 (s, C-1′), 135.91 (s, C-2′ and C-6′), 135.79 (s, C-4′), 129.83 (s, C-7), 128.38 (s, C-3′ and C-5′),118.51 (s, C-3), 117.99 (s, C-4a), 116.95 (s, C-8), 62.02 (s, CH3CH2O), 21.04 (s, CH3Ar), 20.76 (s, two CH3Ar), 14.26 (s, CH3CH2O).
HRMS (ESI) m/z calculated for C21H20O4 [M+H]+ 337.14344 found 337.14352 (ppm: −0.24).
  • Methyl 2-oxo-6-phenyl-2H-chromene-3-carboxylate, 10a
The product was prepared from methyl 6-bromo-2-oxo-2H-chromene-3-carboxylate 8b (0.425 g), phenylboronic acid (0.219 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 60 min. The product was purified by column chromatography using n-hexane/CH2Cl2 and then CH2Cl2, 0.409 g, 97%, pale yellow powder, m.p. = 160.7–164.5 °C. IR (nujol): ν = 1755, 1694 cm−1.
Molecules 27 07649 i058
1H NMR (400 MHz, CDCl3) δ = 8.627 (s, 1H, H-4), 7.871 (dd, J = 8.6 Hz, 2.2 Hz, 1H, H-7), 7.785 (d, J = 2.2 Hz, 1H, H-5), 7.565–7.589 (m, 2H, H-2′ and H-6′), 7.468–7.487 (m, J = 7.4 Hz, 2H, H-3′and H-5′), 7.390–7.432 (m, 1H, H-4′), 7.434 (d, J = 8.8 Hz, 1H, H-8), 3.972 (s, J = 3H, CH3OOC); 13C NMR (100 MHz, CDCl3) 163.72 (s, COOCH3), 156.69 (s, C-2), 154.54 (s, C-8a), 149.22 (s, C-4), 138.87 (s, C-6), 138.37 (s, C-1′), 133.45 (s, C-7), 129.15 (s, C-3′ and C-5′), 128.09 (s, C-4′), 127.52 (s, C-5), 127.04 (s, C-2′ and C-6′), 118.26 (s, C-3), 118.09 (s, C-4a), 117.20 (s, C-8), 52.98 (s, CH3O).
HRMS (ESI) m/z calculated for C17H12O4 [M+H]+ 281.08084 found 281.08112 (ppm: −1.00).
  • Methyl 6-(2-methoxyphenyl)-2-oxo-2H-chromene-3-carboxylate, 10b
The product was prepared from methyl 6-bromo-2-oxo-2H-chromene-3-carboxylate 8b (0.425 g), 2-methoxyphenylboronic acid (0.273 g), K2CO3 (0.622 g) and Pd(PPh3)2Cl2 (32 mg) in 15 mL toluene for 21 h. The product was purified by column chromatography using n-hexane/CH2Cl2 and then CH2Cl2, 0.385 g, 83%, yellow powder, m.p. = 130.9–135.5 °C. IR (nujol): ν = 1760, 1759 cm−1.
Molecules 27 07649 i059
1H NMR (400 MHz, CDCl3) δ = 8.611 (s, 1H, H-4), 7.819 (dd, J = 8.6 Hz, 2.1 Hz, 1H, H-7), 7.763 (d, J = 2.1 Hz, 1H, H-5), 7.393 (d, J = 8.8 Hz, 1H, H-8), 7.374 (td, J = 5.6 Hz, 1.7 Hz 1H, H-5′), 7.311 (dd, J = 7.5 Hz, 1.8 Hz, 1H, H-3′), 7.066 (td, J = 7.5 Hz, 1.0 Hz, H-4′), 7.022 (d, J = 8.3 Hz, H-6′), 3.967 (s, 3H, CH3OOC), 3.834 (s, 3H, CH3O); 13C NMR (100 MHz, CDCl3) δ = 163.86 (s, COOCH3), 156.91 (s, C-2), 154.21 (s, C-8a), 156.32 (s, C-1′), 149.56 (s, C-4), 136.10 (s, C-7), 135.56 (s, C-2′), 130.59 (s, C-5′), 130.11 (s, C-5),129.57 (s, C-3′), 128.15 (s, C-6), 121.11 (s, C-4′), 117.77 (s, C-3), 117.56 (s, C-4a), 116.33 (s, C-8), 111.34 (s, C-6′), 52.93 (s, CH3O).
HRMS (ESI) m/z calculated for C18H14O5 [M+H]+ 311.0914 found 311.09083 (ppm: −1.00).
  • Methyl 6-(4-methoxyphenyl)-2-oxo-2H-chromene-3-carboxylate, 10c
The product was prepared from methyl 6-bromo-2-oxo-2H-chromene-3-carboxylate 8b (0.425 g), 4-methoxyphenylboronic acid (0.273 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 45 min. The product was purified by column chromatography using n-hexane/CH2Cl2 and then CH2Cl2, 0.463 g, 99%, lemon yellow powder, m.p. = 158.7–161.4 °C. IR (nujol): ν = 1756, 1700 cm−1.
Molecules 27 07649 i060
1H NMR (400 MHz, CDCl3) δ = 8.613 (s, 1H, H-4), 7.826 (dd, J = 8.6 Hz, 2.6 Hz, 1H, H-7), 7.726 (d, J = 2.2 Hz, 1H, H-5), 7.506 (dq, J = 8.6 Hz, 3.0 Hz, 2H, H-3′ and H-5′), 7.403 (d, J = 8.6 Hz, 1H, H-8), 7.011 (dq, J = 8.8 Hz, 3.0 Hz, 2H, H-2′ and H-6′), 3.969 (s, J = 3H, CH3OOC), 3.868 (s, J = 3H, CH3O); 13C NMR (100 MHz, CDCl3) δ = 163.77 (s, COOCH3), 159.75 (s, C-4′), 156.77 (s, C-2), 154.16 (s, C-8a), 149.22 (s, C-4), 138.03 (s, C-6), 133.10 (s, C-7), 131.31 (s, C-1′), 129.15 (s, C-5′ and C-3′), 126.89 (s, C-5), 118.07 (s, C-3), 118.15 (s, C-8), 117.12 (s, C-4a), 114.55 (s, C-2′ and C-6′), 52.96 (s, CH3O).
HRMS (ESI) m/z calculated for C18H14O5 [M+H]+ 311.0914 found 311.09106 (ppm: 1.09).
  • Methyl 6-(4-fluorophenyl)-2-oxo-2H-chromene-3-carboxylate, 10d
The product was prepared from methyl 6-bromo-2-oxo-2H-chromene-3-carboxylate 8a (0.425 g), 4-fluorophenylboronic acid (0.252 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 60 min. The product was purified by column chromatography using CH2Cl2, 0.443 g, 99%, pale yellow powder, m.p. = 170.8–174.5.7 °C. IR (nujol): ν = 1743, 1712 cm−1.
Molecules 27 07649 i061
1H NMR (400 MHz, CDCl3) δ = 8.618 (s, 1H, H-4), 7.819 (dd, J = 8.6 Hz, 2.2 Hz, 1H, H-7), 7.734 (d, J = 2.2 Hz, 1H, H-5), 7.550 (ddq, JHF = 8.8 Hz, 5.2 Hz; J = 8.8 Hz, 3.1 Hz, 2H, H-3′ and H-5′), 7.430 (d, J = 8.6 Hz, 1H, H-8), 7.202 (ddq as tq, JHF = 8.6 Hz, 3.6 Hz; J = 8.6 Hz, 3.2 Hz, 2H, H-2′ and H-6′), 3.973 (s, 3H, CH3OOC); 13C NMR (100 MHz, CDCl3) δ = 163.68 (s, COOEt), 162.86 (d, J = 248.0 Hz, C-4′), 156.60 (s, C-2), 154.51 (s, C-8a), 149.07 (s, C-4), 137.40 (s, C-6), 135.05 (d, J = 3.3 Hz, C-1′), 133.26 (s, C-7), 128.71 (d, J = 10.5 Hz, C-2′ and C-6′), 127.38 (s, C-5), 118.11 (s, C-3), 117.29 (s, C-4a), 118.42 (s, C-8), 116.12 (d, J = 21.7 Hz, C-3′ and C-5′), 53.00 (s, CH3O); 19F NMR (376.46 MHz, CDCl3) δ = −114.135.
HRMS (ESI) m/z calculated for C17H11FO4 [M+H]+ 299.07141 found 299.07090 (ppm: 1.71).
  • Methyl 6-mesityl-2-oxo-2H-chromene-3-carboxylate, 10e
The product was prepared from methyl 6-bromo-2-oxo-2H-chromene-3-carboxylate 8a (0.425 g), 2,4,6-trimethylphenylboronic acid (0.294 g), K2CO3 (0.622 g) and Pd(PPh3)2Cl2 (32 mg) in 15 mL toluene for 90 h. The product was purified by column chromatography using n-hexane/CH2Cl2, 0.085 g, 18%, white powder, m.p. = 139.9–143.5 °C. IR (nujol): ν = 1745, 1706 cm−1.
Molecules 27 07649 i062
1H NMR (400 MHz, CDCl3) δ = 8.564 (s, 1H, H-4), 7.438 (s, 1H, H-8), 7.434 (s, 1H, H-5), 7.392 (dd as t, J = 1.0 Hz, 1H, H-7), 6.970 (d, J = 0.4 Hz, 2H, H-3′ and H-5′), 3.968 (s, 3H, CH3OOC), 2.343 (s, 3H, CH3Ar), 1.994 (s, 6H, CH3Ar); 13C NMR (100 MHz, CDCl3) 163.80 (s, COOEt), 156.84 (s, C-2), 154.07 (s, C-8a), 149.21 (s, C-4), 138.11 (s, C-6), 137.57 (s, C-5), 136.46 (s, C-1′), 135.97 (s, C-2′ and C-6′), 137.10 (s, C-4′), 129.90 (s, C-7), 128.39 (s, C-3′ and C-5′), 118.57 (s, C-3), 117.96 (s, C-4a), 116.97 (s, C-8), 52.96 (s, CH3O), 21.04 (s, CH3Ar), 20.77 (s, two CH3Ar).
HRMS (ESI) m/z calculated for C20H18O5 [M+H]+ 323.12779 found 323.12753 (ppm: −0.80).
  • Ethyl 2-ethoxy-6-phenylbenzo[e][1,2]oxaphosphinine-3-carboxylate-2-oxide, 11a
The product was prepared from ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphinine-3-carboxylate-2-oxide 6 (0.540 g), phenylboronic acid (0.220 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 16 h. The product was purified by column chromatography using n-hexane/Et2O, 0.476 g, 89%, pale yellow powder, m.p. = 118.7–120.3 °C. IR (nujol): ν = 1715, 1196, 1076, 1028 cm−1.
Molecules 27 07649 i063
1H NMR (400 MHz, CDCl3) δ = 8.307 (d, J = 36.9 Hz, 1H, H-4), 7.702 (dd, J = 8.6 Hz, 2.0 Hz, 1H, H-7), 7.657 (d, J = 2.1 Hz, 1H, H-5), 7.545 (dq, J = 7.6 Hz, 2.0 Hz, 1.5 Hz, 2H, H-2′ and H-6′), 7.459 (tt, J = 2.0 Hz, 1.5 Hz, 2H, H-3′and H-5′), 7.382 (tt, J = 7.3, 2.2, 1.3 Hz, 1H, H-4′), 7.266 (d, J = 8.5 Hz, 1H, H-8), 4.413–4.493 (m, 2H, CH3CH2OP), 4.323–4.403 (m, 2H, CH3CH2OOC), 1.438 (t, J = 7.0 Hz, 3H, CH3CH2OP), 1.403 (t, J = 7.1 Hz, 3H, CH3CH2OOC);13C NMR (100 MHz, CDCl3) δ = 163.75 (d, JCP = 14.7 Hz, COOEt), 152.01 (d, JCP = 8.7 Hz, C-8a), 150.46 (d, JCP = 3.5 Hz, C-4), 139.06 (s, C-1′), 137.59 (s, C-6), 132.34 (s, C-7), 129.87 (d, JCP = 1.5 Hz, C-5), 127.85 (C-4′), 129.04 (s, C-3′ and C-5′), 126.93 (s, C-2′ and C-6′), 119.68 (s, C-4a), 119.19 (d, JCP = 7.5 Hz, C-8), 118.88 (d, JCP = 193.3 Hz, C-3), 64.93 (d, JCP = 6.3 Hz, CH3CH2OP), 62.04 (s, CH3CH2O), 16.47 (d, JCP = 6.5 Hz, CH3CH2OP), 14.24 (s, CH3CH2O); 31P NMR (161.98 MHz, CDCl3) δ = 5.321.
HRMS (ESI) m/z calculated for C19H19O5P [M+H]+ 359.10429 found 359.10446 (ppm: −0.47).
  • Ethyl 2-ethoxy-6-(2-methoxyphenyl)benzo[e][1,2]oxaphosphinine-3-carboxylate-2-oxide, 11b
The product was prepared from ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphinine-3-carboxylate-2-oxide 6 (0.540 g), 2-methoxyphenylboronic acid (0.273 g), K2CO3 (0.622 g) and Pd(PPh3)2Cl2 (32 mg) in 15 mL toluene for 42 h. The product was purified by column chromatography using n-hexane/EtOAc, 0.366 g, 45%, yellow oil. IR (nujol): ν = 1700, 1190, 1074, 1035 cm−1.
Molecules 27 07649 i064
1H NMR (400 MHz, CDCl3) δ = 8.284 (d, J = 37.0 Hz, 1H, H-4), 7.651 (dd, J = 8.6 Hz, 2.5 Hz, 1H, H-7), 7.636 (s, 1H, H-5), 7.355 (td, J = 8.0 Hz, 2.2 Hz, 1H, H-3′), 7.283 (dd, J = 7.5, 1.7 Hz, 1H, H-5′), 7.223 (d, J = 8.2 Hz, 1H, H-8), 7.023 (td, J = 7.4, 0.9 Hz, 1H, H-6′), 6.982 (td, J = 7.5 Hz, 1.1 Hz, 1H, H-4′), 4.315–4.487 (m, 4H, CH3CH2OP and CH3CH2OOC), 3.820 (s, 3H, OCH3), 1.420 (t, J = 7.1 Hz, 3H, CH3CH2OP), 1.406 (t, J = 7.1 Hz, 3H, CH3CH2OOC); 13C NMR (100 MHz, CDCl3) δ = 163.89 (d, JCP = 13 Hz, COOEt), 156.32 (s, C-1′), 151.63 (d, JCP = 8.8 Hz, C-8a), 150.89 (d, JCP = 3.6 Hz, C-4), 134.98 (s, C-7), 134.74 (s, C-5), 130.54 (s, C-5′), 129.32 (s, C-3′), 128.36 (s, C-6), 121.04 (s, C-6′), 121.03 (C-4′), 119.18 (d, JCP = 15.8 Hz, C-4a), 118.36 (d, JCP = 7.4 Hz, C-8), 118.35 (d, JCP = 177.7 Hz, C-3), 111.31 (s, C-2′), 64.78 (d, JCP = 6.4 Hz, CH3CH2OP), 61.98 (s, CH3CH2O), 55.56 (s, OCH3), 16.46 (d, JCP = 6.5 Hz, CH3CH2OP), 14.24 (s, CH3CH2O); 31P NMR (161.98 MHz, CDCl3) δ = 5.537.
HRMS (ESI) m/z calculated for C20H21O6P [M+H]+ 389.11485 found 389.1149 (ppm: 0.13).
  • Ethyl 2-ethoxy-6-(4-methoxyphenyl)benzo[e][1,2]oxaphosphinine-3-carboxylate-2-oxide, 11c
The product was prepared from ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphinine-3-carboxylate-2-oxide 6 (0.540 g), 4-methoxyphenylboronic acid (0.273 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 3 h. The product was purified by column chromatography using n-hexane/EtOAc, 0.506 g, 87%, pale yellow powder, m.p. = 132.7–135.3 °C IR (nujol): ν = 1720, 1195, 1043, 1026 cm−1.
Molecules 27 07649 i065
1H NMR (400 MHz, CDCl3) δ = 8.296 (d, J = 36.9 Hz, 1H, H-4), 7.662 (dd, J = 8.6 Hz, 2.0 Hz, 1H, H-7), 7.603 (d, J = 2.3 Hz, 1H, H-5), 7.475 (dq, J = 8.8 Hz, 3.0 Hz, 2H, H-2′ and H-6′), 6.989 (dq, J = 8.8 Hz, 3.0 Hz, 2H, H-3′and H-5′), 7.237 (d, J = 8.5 Hz, 1H, H-8), 4.319–4.490 (m, 4H, CH3CH2OP and CH3CH2OOC), 1.423 (t, J = 7.0 Hz, 3H, CH3CH2OP), 1.409 (t, J = 7.1 Hz, 3H, CH3CH2OOC);13C NMR (100 MHz, CDCl3) δ = 163.78 (d, JCP = 13.0 Hz, COOEt), 151.58 (d, JCP = 8.6 Hz, C-8a), 150.59 (d, JCP = 3.5 Hz, C-4), 131.55 (s, C-1′), 137.26 (s, C-6), 131.95 (s, C-7), 129.35 (d, JCP = 1.2 Hz, C-5), 159.55 (C-4′), 114.47 (s, C-3′ and C-5′), 127.99 (s, C-2′ and C-6′), 119.64 (s, C-4a), 119.11 (d, JCP = 7.4 Hz, C-8), 118.66 (d, JCP = 177.3 Hz, C-3), 64.89 (d, JCP = 6.3 Hz, CH3CH2OP), 62.02 (s, CH3CH2O), 16.47 (d, JCP = 6.4 Hz, CH3CH2OP), 14.24 (s, CH3CH2O); 31P NMR (161.98 MHz, CDCl3) δ = 5.408.
HRMS (ESI) m/z calculated for C20H21O6P [M+H]+ 389.11485 found 389.11517 (ppm: −0.82).
  • Ethyl 2-ethoxy-6-(4-fluorophenyl)benzo[e][1,2]oxaphosphinine-3-carboxylate-2-oxide, 11d
The product was prepared from ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphinine-3-carboxylate-2-oxide 6 (0.540 g), 4-fluorophenylboronic acid (0.251 g), K2CO3 (0.622 g) and PEPPSI (29 mg) in 3 mL:12 mL water:toluene for 2.5 h. The product was purified by column chromatography using n-hexane/EtOAc, 0.500 g, 89%, pale yellow powder, m.p. = 115.0–118.2 °C IR (nujol): ν = 1718, 1198, 1078, 1028 cm−1.
Molecules 27 07649 i066
1H NMR (400 MHz, CDCl3) δ = 8.294 (d, J = 36.5 Hz, 1H, H-4), 7.647 (dd, J = 8.5 Hz, 2.1 Hz, 1H, H-7), 7.606 (d, J = 2.2 Hz, 1H, H-5), 7.502 (ddq, J = 8.8 Hz, 5,2 Hz, 3.4 Hz, 2H, H-2′ and H-6′), 7.149 (tq, J = 8.6 Hz, 3.4 Hz, 2H, H-3′and H-5′), 7.261 (d, J = 8.4 Hz, 1H, H-8), 4.323–4.498 (m, 4H, CH3CH2OP and CH3CH2OOC), 1.429 (t, J = 7.1 Hz, 3H, CH3CH2OP), 1.411 (t, J = 7.1 Hz, 3H, CH3CH2OOC);13C NMR (100 MHz, CDCl3) δ = 163.68 (d, JCP = 13.0 Hz, COOEt), 151.98 (d, JCP = 8.6 Hz, C-8a), 150.27 (d, JCP = 3.5 Hz, C-4), 135.23 (d, J = 3.1 Hz C-1′), 136.62 (s, C-6), 132.16 (s, C-7), 129.71 (s, C-5), 162.72 (d, JCP = 247.7 Hz, C-4′), 128.58 (d, JCP = 8.1 Hz, C-3′ and C-5′), 115.98 (d, JCP = 21.6 Hz, C-2′ and C-6′), 119.88 (s, C-4a), 119.27 (d, JCP = 7.4 Hz, C-8), 118.92 (d, JCP = 162.2 Hz, C-3), 64.98 (d, JCP = 6.3 Hz, CH3CH2OP), 62.07 (s, CH3CH2O), 16.47 (d, JCP = 6.4 Hz, CH3CH2OP), 14.24 (s, CH3CH2O); 31P NMR (161.98 MHz, CDCl3) δ = 5.232; 19F NMR (376.46 MHz, CDCl3) δ = −114.590.
HRMS (ESI) m/z calculated for C19H18FO5P [M+H]+ 377.09486 found 377.09515 (ppm: −0.77).

3.4. General Procedure for the Sonogashira Reaction

A mixture of the corresponding coumarin derivative (1.4 mmol), aryl ethyne (1.8 mmol, 1.2 equiv.), PdCl2(PPh3)2/PEPPSI (2 mol%), CuI (4 mol%), Et3N (3 equiv), and DMF (15 mL) was heated at 80 °C under an Ar atmosphere until the coumarin derivative was consumed (TLC-monitoring). The needed time for the reactions is given in Table 5. Then, it was concentrated under reduced pressure. The crude product was dissolved in EtOAc, watched several times with brine (5 × 5 mL), and once with water (1 × 10 mL). The organic layer was dried with anhydrous sodium. After the evaporation of the solvent, the residue was purified by column chromatography.
  • Diethyl (2-oxo-6-(phenylethynyl)-2H-chromen-3-yl)phosphonate, 12a
The product was prepared from diethyl (6-bromo-2-oxo-2H-chromen-3-yl)phosphonate 5 (0.500 g), ethynylbenzene (0.170 g), PdCl2(PPh3)2 (19.4 mg), CuI (10.5 mg), Et3N (0.58 mL, 0.42 g) and DMF (15 mL) for 20 h. The product was purified by column chromatography using n-hexane/EtOAc, 0.437 g, 82%, reddish colored powder, m.p. = 94.3–99.6 °C. IR (nujol): ν = 1745, 1239, 1046, 1018 cm−1.
Molecules 27 07649 i067
1H NMR (500 MHz, CDCl3) δ = 8.413 (d, J = 17.2 Hz, 1H, H-4), 7.691 (s, 1H, H-5), 7.674 (s, 1H, H-7), 7.465 (s, 1H, H-8), 7.297 (s, 2H, H-2′ and H-6′), 7.271 (s, 2H, H-3′ and H-5′), 7.254 (s, 1H, H-4′), 4.146–4.261 (m, 4H, two CH3CH2O), 1.316 (t, J = 7.1 Hz, 6H, two CH3CH2O);13C NMR (125.7 MHz, CDCl3) δ = 157.81 (d, JCP = 14.4 Hz, C-2), 154.62 (s, C-8a), 152.64 (d, JCP = 6.6 Hz, C-4), 137.06 (s, C-7), 132.07 (s, C-5), 131.67 (s, C-6′and C-2′), 128.83 (s, C-4′), 128.48 (s, C-3′ and C-5′), 122.44 (s, C-6), 120.49 (s, C-1′), 118.75 (d, JCP = 196.4 Hz, C-3), 117.99 (d, JCP = 14.5 Hz, C-4a), 117.93 (s, C-8), 90.67 (s, C≡C), 87.02 (s, C≡C), 63.57 (d, JCP = 6.9 Hz, two CH3CH2O), 16.38 (d, JCP = 6.3 Hz, two CH3CH2O); 31P NMR (161.98 MHz, CDCl3) δ = 10.343.
HRMS (ESI) m/z calculated for C21H19O5P [M+H]+ 383.10429 found 383.10468 (ppm: −1.02).
  • Ethyl 2-oxo-6-(phenylethynyl)-2H-chromene-3-carboxylate, 12b
The product was prepared from ethyl 6-bromo-2-oxo-2H-chromene-3-carboxylate 8a (0.444 g), ethynylbenzene (0.184 g), PdCl2(PPh3)2 (21 mg), CuI (11.4 mg), Et3N (0.63 mL, 0.46 g) and DMF (15 mL) for 22 h. The product was purified by column chromatography using n-hexane/EtOAc, 0.247 g, 52%, reddish colored powder, m.p. = 57.1–61.8 °C. IR (nujol): ν = 1744, 1711 cm−1.
Molecules 27 07649 i068
1H NMR (400 MHz, CDCl3) δ = 8.485 (s, 1H, H-4), 7.774 (s, 1H, H-5), 7.755 (dd, J = 7.3 Hz, 2.0 Hz, 1H, H-7), 7.530–7.555 (m, 2H, H-2′ and H-6′), 7.341 (dd, J = 9.2 Hz, 0.4 Hz, 1H, H-8), 7.363–7.388 (m, 2H, H-3′ and H-5′), 4.421 (q, J = 7.1 Hz, 2H, CH3CH2O), 1.420 (t, J = 7.1 Hz, 3H, CH3CH2O); 13C NMR (100 MHz, CDCl3) δ = 162.87 (s, COOEt), 156.23 (s, C-2), 154.54 (s, C-8a), 147.82 (s, C-4), 122.46 (s, C-6), 120.43 (s, C-1′), 137.19 (s, C-7), 128.49 (s, C-3′ and C-5′), 128.83 (s, C-4′), 132.19 (s, C-5), 131.66 (s, C-2′ and C-6′), 117.94 (s, C-3), 119.11 (s, C-4a), 117.10 (s, C-8), 62.16 (s, CH3CH2O), 14.22 (s, CH3CH2O).
HRMS (ESI) m/z calculated for C20H14O4 [M+H]+ 319.09649 found 319.09641 (ppm: 0.25).
  • Diethyl (6-((4-methoxyphenyl)ethynyl)-2-oxo-2H-chromen-3-yl)phosphonate, 12d
The product was prepared from diethyl (6-bromo-2-oxo-2H-chromen-3-yl)phosphonate 5 (0.540 g), 1-ethynyl-4-methoxybenzene (0.238 g), PEPPSI (19.4 mg), CuI (11.4 mg), Et3N (0.62 mL, 0.55 g) and DMF (15 mL) for 30 h. The product was purified by column chromatography using n-hexane/EtOAc, 0.349 g, 56%, reddish colored powder, m.p. = 131.7–136.0 °C. IR (nujol): ν = 2211, 1744, 1243, 1060, 1027 cm−1.
Molecules 27 07649 i069
1H NMR (400 MHz, CDCl3) δ = 8.478 (d, J = 17.2 Hz, 1H, H-4), 7.735 (dd, J = 8.5 Hz, 2.0 Hz, 1H, H-7), 7.716 (d, J = 2.0 Hz, 1H, H-5), 7.327 (d, J = 8.5 Hz, 1H, H-8), 7.485 (dq, J = 8.8 Hz, 2.8 Hz, 2H, H-2′ and H-6′), 6.901 (dq, J = 8.8 Hz, 2.8 Hz, 2H, H-3′ and H-5′), 4.226–4.343 (m, 4H, two CH3CH2O), 3.844 (s, 3H, CH3O), 1.392 (t, J = 7.1 Hz, 6H, two CH3CH2O); 13C NMR (100 MHz, CDCl3) δ = 160.06 (s, C-4′), 157.82 (s, C-2), 154.43 (s, C-8a), 152.69 (d, JCP = 6.6 Hz, C-4), 136.95 (s, C-7), 131.79 (s, C-5), 133.19 (s, C-6′and C-2′), 114.49 (s, C-3′ and C-5′), 120.90 (s, C-6), 119.69 (s, C-1′), 118.72 (d, JCP = 196.1 Hz, C-3), 117.98 (d, JCP = 14.5 Hz, C-4a), 117.13 (s, C-8), 90.79 (s, C≡C), 85.83 (s, C≡C), 55.35 (s, CH3O), 63.54 (d, JCP = 6.0 Hz, two CH3CH2O), 16.39 (d, JCP = 6.3 Hz, two CH3CH2O); 31P NMR (161.98 MHz, CDCl3) δ = 10.390.
HRMS (ESI) m/z calculated for C22H21O6P [M+H]+ 413.11485 found 413.11524 (ppm: −0.94).

4. Conclusions

A synthetic protocol for the efficient preparation of new substituted 3-phosphonocoumarins, 1,2-benzoxaphosphorines, methyl and ethyl coumarin-3-carboxylates via palladium catalyzed cross-coupling reactions was developed. The Suzuki and Sonogashira reaction was applied successfully to obtain the structures in good to quantitative yields.
Quantum-chemical calculations on the spectral properties of aryl and alkynyl 3-phosphonocoumarins were performed as a continuation of our systematic studies on phosphorous-containing coumarin derivatives. It was found that by altering the substituent in C-6, fine tuning of the fluorescent properties could be achieved. Introducing a CH3O-group in para-position in the aryl moiety has the most noticeable effect leading to bathochromic shift in the absorption and fluorescence spectra. This tendency was observed both in the calculated and the experimental data. The photophysical properties of the newly obtained compounds were studied in different solvents. It was found that the type of media has little or no effect on both absorption and fluorescence maxima. The experimental data are in good agreement with the theoretically predicted photophysical properties of the compounds.

Supplementary Materials

Crystallographic data for the structures in this paper have been deposited in the Cambridge Crystallographic Data Center as a supplementary publication (3a, CCDC 2209998). Copies of the data can be obtained, free of charge, on application to CCDC, 12 Union Road, Cambridge CB12 1EZ, U.K [Fax: +44 1223 336033 or e-mail: [email protected]. The Supplemental Material file of CCDC includes the CIF file of 3a. The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27217649/s1, NMR spectroscopy data and X-ray analysis.

Author Contributions

Conceptualization, R.L., A.I.K. and I.Z.K.; methodology, R.L., I.Z.K. and K.B.S.; spectroscopic analysis, N.P.-Y., B.M., I.Z.K. and V.L.; investigation, R.L., A.I.K., I.Z.K., K.S., I.M. and K.B.S.; data curation, R.L., A.I.K., I.Z.K., K.S., I.M., N.P.-Y. and K.B.S.; writing—original draft preparation, R.L., A.I.K., I.Z.K. and K.B.S.; writing—review and editing, N.P.-Y., R.D.N. and P.Y.P.; visualization, R.L., A.I.K. and I.Z.K.; supervision, R.L. and A.I.K.; project administration, A.I.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science Fund project, grant number KP-06-M-59/6 from 19.11.2021.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Instrumentation and technical assistance for this work were provided by the Service Center X-ray Diffraction, with financial support from Saarland University and German Science Foundation (project number INST 256/506-1). R.L., A.I.K., I.Z.K., K.B.S., N.P. and R.D.N. are grateful to the Extreme Project within the program “EUROPEAN SCIENTIFIC NETWORKS”, funded by the Bulgarian Ministry of Education and Science. K.B.S. acknowledges the European Regional Development Fund within the Operational Programme “Science and Education for Smart Growth 2014–2020” under the Project CoE “National center of mechatronics and clean technologies” BG05M2OP001-1.001-0008 for the provided computational resources.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Koleva, A.I.; Petkova-Yankova, N.I.; Nikolova, R.D. Synthesis and Chemical Properties of 3-Phosphonocoumarins and 1,2-Benzoxaphosphorins as Precursors for Bioactive Compounds. Molecules 2019, 24, 2030. [Google Scholar] [CrossRef] [Green Version]
  2. Bhatia, R.; Pathania, S.; Singh, V.; Rawalet, R.K. Metal-catalyzed synthetic strategies toward coumarin derivatives. Chem. Heterocycl. Comp. 2018, 54, 280–291. [Google Scholar] [CrossRef]
  3. Szwaczko, K. Coumarins Synthesis and Transformation via C–H Bond Activation—A Review. Inorganics 2022, 10, 23. [Google Scholar] [CrossRef]
  4. Annunziata, F.; Pinna, C.; Dallavalle, S.; Tamborini, L.; Pinto, A. An Overview of Coumarin as a Versatile and Readily Accessible Scaffold with Broad-Ranging Biological Activities. Int. J. Mol. Sci. 2020, 21, 4618. [Google Scholar] [CrossRef] [PubMed]
  5. Cheke, R.S.; Patel, H.M.; Patil, V.M.; Ansari, I.A.; Ambhore, J.P.; Shinde, S.D.; Kadri, A.; Snoussi, M.; Adnan, M.; Kharkar, P.S.; et al. Molecular Insights into Coumarin Analogues as Antimicrobial Agents: Recent Developments in Drug Discovery. Antibiotics 2022, 11, 566. [Google Scholar] [CrossRef]
  6. Yañez, O.; Osorio, M.I.; Uriarte, E.; Areche, C.; Tiznado, W.; Pérez-Donoso, J.M.; García-Beltrán, O.; González-Nilo, F. In Silico Study of Coumarins and Quinolines Derivatives as Potent Inhibitors of SARS-CoV-2 Main Protease. Front. Chem. 2021, 8, 595097. [Google Scholar] [CrossRef]
  7. Abdelmohsen, A.R.; Albohy, A.; Abdulrazik, B.S.; Bayoumi, S.A.L.; Malak, L.G.; Khallaf, I.S.A.; Bringmann, G.; Salwa, F.; Farag, S.F. Natural coumarins as potential anti-SARS-CoV-2 agents supported by docking analysis. RSC Adv. 2021, 11, 16970. [Google Scholar] [CrossRef]
  8. Sinkeldam, R.W.; Greco, N.J.; Tor, Y. Fluorescent Analogs of Biomolecular Building Blocks: Design, Properties, and Applications. Chem. Rev. 2010, 110, 2579. [Google Scholar] [CrossRef] [Green Version]
  9. Traven, V.F.; Cheptsov, D.A.; Solovjova, N.P.; Chibisova, T.A.; Voronov, I.I.; Dolotov, S.M.; Ivanov, I.V. Photoinduced Formation of the Laser Dye Coumarin 6 from Its Dihydro Derivatives. Dyes Pigments 2017, 146, 159. [Google Scholar] [CrossRef]
  10. Signore, G.; Nifosì, R.; Albertazzi, L.; Storti, B.; Bizzarri, R. Polarity-Sensitive Coumarins Tailored to Live Cell Imaging. J. Am. Chem. Soc. 2010, 132, 1276. [Google Scholar] [CrossRef]
  11. Bayrakçeken, F.; Yaman, A.; Hayvalı, M. Photophysical and Photochemical Study of Laser-Dye Coumarin-481 and Its Photoproduct in Solution. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2005, 61, 983. [Google Scholar] [CrossRef] [PubMed]
  12. Rajesh, H.; Vekariya, R.H.; Patel, H.D. Recent Advances in the Synthesis of Coumarin Derivatives via Knoevenagel Condensation: A Review. Synth. Commun. 2014, 44, 2756. [Google Scholar] [CrossRef]
  13. Ansary, I.; Taher, A. One-Pot Synthesis of Coumarin Derivatives. In Phytochemicals in Human Health; Intech Open: London, UK, 2019. [Google Scholar]
  14. Gulati, S.; Singh, R.; Sangwan, S. A review on convenient synthesis of substituted coumarins using reuseable solid acid catalysts. RSC Adv. 2021, 11, 29130. [Google Scholar] [CrossRef] [PubMed]
  15. Kanchana, U.S.; Diana, E.J.; Mathew, T.V.; Anilkumar, G. Palladium-catalyzed cross-coupling reactions of coumarin derivatives: An overview. Appl Organomet. Chem. 2020, 34, e5983. [Google Scholar] [CrossRef]
  16. Wu, J.; Zhanga, L.; Xia, H. Palladium-catalyzed Suzuki–Miyaura couplings of potassium aryl trifluoroborates with 4-tosyloxycoumarins or 4-tosyloxyquinolin-2(1H)-one. Tetrahedron Letters 2006, 47, 1525. [Google Scholar] [CrossRef]
  17. Li, K.; Zeng, Y.; Neuenswander, B.; Tunge, J.A. Sequential Pd(II)-Pd(0) Catalysis for the Rapid Synthesis of Coumarins. J. Org. Chem. 2005, 70, 6515. [Google Scholar] [CrossRef]
  18. Yin, Q.; Yan, H.; Zhang, Y.; Wang, Y.; Zhang, G.; He, Y.; Zhang, W. Palladium-catalyzed synthesis of 8-allyl or 8-prenylcoumarins by using organotin reagents as multicoupling nucleophiles. Appl. Organometal. Chem. 2013, 27, 85. [Google Scholar] [CrossRef]
  19. Mao, W.; Wang, T.; Zeng, H.; Wang, Z.; Chen, J.; Shen, J. Synthesis and evaluation of novel substituted 5-hydroxycoumarin and pyranocoumarin derivatives exhibiting significant antiproliferative activity against breast cancer cell lines. Bioorg. Med. Chem. Lett. 2009, 19, 4570. [Google Scholar] [CrossRef]
  20. Bhaumick, P.; Jana, A.; Choudhury, L.H. Synthesis of novel coumarin containing conjugated fluorescent polymers by Suzuki cross-coupling reactions and their chemosensing studies for iron and mercury ions. Polymer 2021, 218, 123415. [Google Scholar] [CrossRef]
  21. Reddy, K.H.V.; Brion, J.; Messaoudi, S.; Alami, M. Synthesis of Biheterocycles Based on Quinolinone, Chromone, and Coumarin Scaffolds by Palladium-Catalyzed Decarboxylative Couplings. J. Org. Chem. 2016, 81, 424. [Google Scholar] [CrossRef]
  22. Balalas, T.; Abdul-Sada, A.; Hadjipavlou-Litina, D.J.; Litinas, K.E. Pd-Catalyzed Efficient Synthesis of Azacoumestans Via Intramolecular Cross Coupling of 4-(Arylamino)coumarins in the Presence of Copper Acetate under Microwaves. Synthesis 2017, 49, 2575. [Google Scholar] [CrossRef]
  23. Cocco, A.; Caria, P.; Sanna, G.; Stagi, L.; Cadoni, E.; Corpino, R.; Ricci, R.C.; Carbonaro, C.M.; Secci, F. Synthesis and Photophysical Properties of Fluorescent 6-Aryl-D-π-A Coumarin Derivatives. ACS Omega 2021, 6, 33708. [Google Scholar] [CrossRef] [PubMed]
  24. Starčević, Š.; Brožič, P.; Turk, S.; Cesar, J.; Rižner, T.L.; Stanislav Gobec, S. Synthesis and Biological Evaluation of (6- and 7-Phenyl) Coumarin Derivatives as Selective Nonsteroidal Inhibitors of 17β-Hydroxysteroid Dehydrogenase Type 1. J. Med. Chem. 2011, 54, 248. [Google Scholar] [CrossRef] [PubMed]
  25. Thanzeel, F.Y.; Wolf, C. Chemoselective bioconjugation based on modular click chemistry with 4-halocoumarins and arylsulfonates. RSC Adv. 2021, 11, 18960. [Google Scholar] [CrossRef]
  26. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef] [Green Version]
  27. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef] [Green Version]
  28. Parr, R.G.; Yang, W. Density-Functional Theory of Atoms and Molecules; Oxford University Press: Oxford, UK, 1989; ISBN 9780195092769. [Google Scholar]
  29. Salahub, D.R.; Zerner, M.C. The Challenge of d and f Electrons. In Chemical Congress of North America; American Chemical Society: Washington, DC, USA, 1989; ISBN 9780841216280. [Google Scholar] [CrossRef]
  30. Adamo, C.; Jacquemin, D. The calculations of excited-state properties with Time-Dependent Density Functional Theory. Chem. Soc. Rev. 2013, 42, 845. [Google Scholar] [CrossRef]
  31. Laurent, A.D.; Adamo, C.; Jacquemin, D. Dye chemistry with time-dependent density functional theory. Phys. Chem. Chem. Phys. 2014, 16, 14334. [Google Scholar] [CrossRef]
  32. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Montgomery, J.A., Jr.; Vreven, T.; Kudin, K.N.; Burant, J.C.; et al. Revision, B; Gaussian. Inc.: Pittsburgh, PA, USA, 2016. [Google Scholar]
  33. Oberhofer, K.E.; Musheghyan, M.; Wegscheider, S.; Wörle, M.; Iglev, E.D.; Nikolova, R.D.; Reinhard Kienberger, R.; Pekov, P.S.; Iglev, H. Individual control of singlet lifetime and triplet yield in halogen-substituted coumarin derivatives. RSC Adv. 2020, 10, 27096. [Google Scholar] [CrossRef]
  34. Wagner, M.S.; Ilieva, E.D.; Petkov, P.S.; Nikolova, R.D.; Kienbergera, R.; Iglev, H. Ultrafast hydrogen bond dynamics and partial electron transfer after photoexcitation of diethyl ester of 7-(diethylamino)-coumarin-3-phosphonic acid and its benzoxaphosphorin analog. Phys. Chem. Chem. Phys. 2015, 17, 9919. [Google Scholar] [CrossRef] [Green Version]
  35. Yanai, T.; Tew, D.; Handy, N. A new hybrid exchange-correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51. [Google Scholar] [CrossRef] [Green Version]
  36. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum mechanical continuum solvation models. Chem. Rev. 2005, 105, 2999. [Google Scholar] [CrossRef] [PubMed]
  37. Raju, B.C.; Tiwari, A.K.; Kumar, J.A.; Ali, A.Z.; Agawane, S.B.; Saidachary, G.; Madhusudana, K. a-Glucosidase inhibitory antihyperglycemic activity of substituted chromenone derivatives. Bioorganic Med. Chem. 2010, 18, 358. [Google Scholar] [CrossRef] [PubMed]
  38. Bojilova, A.; Nikolova, R.; Ivanov, C.; Rodios, N.; Terzis, A.; Raptopoulou, C. A comparative study of the interaction of salicylaldehydes with phosphonoacetates under Knoevenagel reaction conditions. Synthesis of 1,2-benzoxaphosphorines and their dimers. Tetrahedron 1996, 52, 12597–12612. [Google Scholar] [CrossRef]
  39. Reeves, E.K.; Olivia, R.; Bauman, O.R.; Mitchem, G.B.; Neufeldt, S.R. Solvent Effects on the Selectivity of Palladium-Catalyzed Suzuki-Miyaura Couplings. Isr. J. Chem. 2019, 59, 1. [Google Scholar] [CrossRef]
  40. Carrow, B.P.; Hartwig, J.F. Distinguishing Between Pathways for Transmetalation in Suzuki-Miyaura Reactions. J. Am. Chem. Soc. 2011, 133, 2116. [Google Scholar] [CrossRef] [Green Version]
  41. Amatore, C.; Jutand, A.; Le Duc, G. Kinetic Data for the Transmetalation/Reductive Elimination in Palladium-Catalyzed Suzuki–Miyaura Reactions: Unexpected Triple Role of Hydroxide Ions Used as Base. Chem. Eur. J. 2011, 17, 2492. [Google Scholar] [CrossRef]
  42. Bojilova, A.; Kostadinova, T.; Ivanov, C. Interaction of Nitromethane with Esters and Amides of 2-Oxo-2H-1-benzopyrane-3-carboxylic Acid. Synth. Commun. 1989, 19, 2963. [Google Scholar] [CrossRef]
  43. Grant, J.A.; Bonnick, T.; Gossell-Williams, M.; Clayton, T.; Cook, J.M.; Jackson, Y.A. Synthesis, pharmacological studies and molecular modeling of some tetracyclic 1,3-diazepinium chlorides. Bioorganic Med. Chem. 2010, 18, 909. [Google Scholar] [CrossRef]
  44. O’Brien, C.J.; Kantchev, E.A.B.; Valente, C.; Hadei, N.; Chass, G.A.; Lough, A.; Hopkinson, A.C.; Michael, G.; Organ, M.G. Easily Prepared Air- and Moisture-Stable Pd–NHC (NHC=N-Heterocyclic Carbene) Complexes: A Reliable, User-Friendly, Highly Active Palladium Precatalyst for the Suzuki–Miyaura Reaction. Chem. A Eur. J. 2006, 12, 4743. [Google Scholar] [CrossRef]
  45. Kantchev, E.A.B.; Peh, G.; Zhang, C.; Ying, J.Y. Practical Heck−Mizoroki Coupling Protocol for Challenging Substrates Mediated by an N-Heterocyclic Carbene-Ligated Palladacycle. Org. Lett. 2008, 10, 3949. [Google Scholar] [CrossRef] [PubMed]
  46. Clique, B.; Fabritius, C.; Couturier, C.; Monteiro, N.; Balme, G. Unexpected isolation, and structural characterization, of a β-hydrogen-containing σ-alkylpalladium halide complex in the course of an intramolecular Heck reaction. Synthesis of polycyclic isoquinoline derivatives. Chem. Commun. 2003, 272–273. [Google Scholar] [CrossRef] [PubMed]
  47. Lázaro-Milla, C.; Quirós, M.T.; Cárdenas, D.J.; Almendros, P. Triflyl-assisted reductive Pd-catalyzed Tsuji–Trost type reaction. Chem. Commun. 2020, 56, 6070. [Google Scholar] [CrossRef] [PubMed]
  48. Kagalwala, H.N.; Tong, L.; Zong, R.; Kohler, L.; Ahlquist, M.S.G.; Fan, T.; Gagnon, K.J.; Thummel, R.P. Evidence for Oxidative Decay of a Ru-Bound Ligand during Catalyzed Water Oxidation. ACS Catal. 2017, 7, 2607. [Google Scholar] [CrossRef]
  49. Bojilova, A.; Kostadinova, T.; Ivanov, C. Allylation-Assisted Addition of Nitromethane to 2H-1-Benzopyran-2-ones Substituted in Position 3. Liebigs Ann. Chem. 1989, 1041–1043. [Google Scholar] [CrossRef]
  50. Ilieva, E.D.; Petkova, N.I.; Nikolova, R.D. Ring Opening Reactions of 3-Phosphonocoumarin Under Michael Reaction Conditions. Phosphorus Sulfur Silicon 2012, 187, 39, published online December 2011. [Google Scholar] [CrossRef]
  51. Jackson, Y.A.; Williams, M.F. Heterocycles from 2-Aminopyridine and Derivatives of 3-Methylbenzofuran-2-carboxylic Acid. Heterocycles 1997, 45, 787. [Google Scholar] [CrossRef]
  52. Skarga, V.V.; Matrosov, A.A.; Nichugovskiy, A.I.; Negrebetsky, V.V.; Maslov, M.A.; Boldyrev, I.A.; Malakhov, M.V. pH-Dependent Photoinduced Interconversion of Furocoumaric and Furocoumarinic Acids. Molecules 2021, 26, 2800. [Google Scholar] [CrossRef]
  53. Khoobi, M.; Molaverdi, F.; Alipour, M.; Jafarpour, F.; Foroumadi, A.; Shafiee, A. Palladium-catalyzed domino protodecarboxylation/oxidative Heck reaction: Regioselective arylation of coumarin-3-carboxylic acids. Tetrahedron 2013, 69, 11164. [Google Scholar] [CrossRef]
  54. Zhang, L.; Meng, T.; Fan, R.; Wu, J. General and Efficient Route for the Synthesis of 3,4-Disubstituted Coumarins via Pd-Catalyzed Site-Selective Cross-Coupling Reactions. J. Org. Chem. 2007, 72, 7279. [Google Scholar] [CrossRef]
  55. Majumdar, K.C.; Mondal, S. A new strategy for the synthesis of coumarin- and quinolone-annulated pyrroles via Pd(0) mediated cross-coupling followed by Cu(I) catalyzed heteroannulation. Tetrahedron Lett. 2008, 49, 2418. [Google Scholar] [CrossRef]
  56. Elenkova, D.; Lyapchev, R.; Romanova, J.; Morgenstern, B.; Dimitrova, Y.; Dimov, D.; Tsvetkov, M.; Zaharieva, J. Luminescent Complexes of Europium (III) with 2-(Phenylethynyl)-1,10-phenanthroline: The Role of the Counterions. Molecules 2021, 26, 7272. [Google Scholar] [CrossRef] [PubMed]
  57. Yin, H.; Zhang, B.; Haizhu, Y.H.; Zhu, L.; Feng, Y.; Zhu, M.; Guo, Q.; Meng, X. Two-Photon Fluorescent Probes for Biological Mg2+ Detection Based on 7-Substituted Coumarin. J. Org. Chem. 2015, 80, 4306. [Google Scholar] [CrossRef]
  58. Arsenyan, P.; Vasiljeva, J.; Domracheva, I.; Kanepe-Lapsa, I.; Gulbe, A. Selenopheno[2,3-f]coumarins: Novel scaffolds with antimetastatic activity against melanoma and breast cancer. N. J. Chem. 2019, 43, 11851. [Google Scholar] [CrossRef]
  59. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. A 2017, 71, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. C 2015, 71, 3. [Google Scholar] [CrossRef] [Green Version]
  61. Hübschle, C.B.; Sheldrick, G.M.; Dittrich, B. ShelXle: A Qt graphical user interface for SHELXL. J. Appl. Crystallogr. 2011, 44, 1281. [Google Scholar] [CrossRef] [Green Version]
  62. Miyaura, N.; Suzuki, A. Palladium-catalyzed reaction of 1-alkenylboronates with vinylic halides: (1Z,3E)-1-phenyl-1,3-octadiene. Org. Synth. 1990, 68, 130. [Google Scholar] [CrossRef]
Scheme 1. Suzuki reaction of 5-bromo-2-hydroxybenzaldehyde 1 with phenylboronic acid.
Scheme 1. Suzuki reaction of 5-bromo-2-hydroxybenzaldehyde 1 with phenylboronic acid.
Molecules 27 07649 sch001
Scheme 2. Knoevenagel condensation.
Scheme 2. Knoevenagel condensation.
Molecules 27 07649 sch002
Scheme 3. Suzuki coupling.
Scheme 3. Suzuki coupling.
Molecules 27 07649 sch003
Figure 1. Pd-complexes used as catalyst in the studied reaction.
Figure 1. Pd-complexes used as catalyst in the studied reaction.
Molecules 27 07649 g001
Scheme 4. Suzuki coupling—optimizing the conditions.
Scheme 4. Suzuki coupling—optimizing the conditions.
Molecules 27 07649 sch004
Figure 2. ORTEP representation with thermal ellipsoids 40%—Crystal Data for C19H19O5P (M = 358.333 g/mol): monoclinic, space group P21/n (no. 14), a = 13.6869(9) Å, b = 8.5123(5) Å, c = 14.9033(9) Å, α = 90°, β = 90.726(2)°, γ = 90°, V = 1736.20(19) Å3, Z = 4, T = 133.00 K, μ(MoKα) = 0.185 mm−1, Dcalc = 1.371 g/cm3, 26,401 reflections measured (4.06° ≤ 2Θ ≤ 55.84°), 4155 unique (Rint = 0.0591, Rsigma = 0.0405) which were used in all calculations. The final R1 was 0.0390 (I > 2σ(I)) and wR2 was 0.1092 (all data). The crystallographic data could be found in the Supplementary Materials.
Figure 2. ORTEP representation with thermal ellipsoids 40%—Crystal Data for C19H19O5P (M = 358.333 g/mol): monoclinic, space group P21/n (no. 14), a = 13.6869(9) Å, b = 8.5123(5) Å, c = 14.9033(9) Å, α = 90°, β = 90.726(2)°, γ = 90°, V = 1736.20(19) Å3, Z = 4, T = 133.00 K, μ(MoKα) = 0.185 mm−1, Dcalc = 1.371 g/cm3, 26,401 reflections measured (4.06° ≤ 2Θ ≤ 55.84°), 4155 unique (Rint = 0.0591, Rsigma = 0.0405) which were used in all calculations. The final R1 was 0.0390 (I > 2σ(I)) and wR2 was 0.1092 (all data). The crystallographic data could be found in the Supplementary Materials.
Molecules 27 07649 g002
Scheme 5. Suzuki coupling of 6-bromo-3-phosphonocoumarin 5.
Scheme 5. Suzuki coupling of 6-bromo-3-phosphonocoumarin 5.
Molecules 27 07649 sch005
Scheme 6. Palladium-catalyzed arylation of 6-bromocoumarin-3-carboxylates 8a,b.
Scheme 6. Palladium-catalyzed arylation of 6-bromocoumarin-3-carboxylates 8a,b.
Molecules 27 07649 sch006
Scheme 7. Palladium-catalyzed arylation of ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphorine-3-carboxylate 6.
Scheme 7. Palladium-catalyzed arylation of ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphorine-3-carboxylate 6.
Molecules 27 07649 sch007
Scheme 8. Palladium-catalyzed alkynylation of 3-substituted coumarin species—Sonogashira coupling.
Scheme 8. Palladium-catalyzed alkynylation of 3-substituted coumarin species—Sonogashira coupling.
Molecules 27 07649 sch008
Figure 3. Normalized fluorescent spectra of the phenyl substituted derivatives.
Figure 3. Normalized fluorescent spectra of the phenyl substituted derivatives.
Molecules 27 07649 g003
Figure 4. UV-VIS and normalized fluorescence spectra of 3a-b, 12a, and 12d phosphonocoumarins in acetonitrile.
Figure 4. UV-VIS and normalized fluorescence spectra of 3a-b, 12a, and 12d phosphonocoumarins in acetonitrile.
Molecules 27 07649 g004
Figure 5. UV-VIS and normalized fluorescence spectra in acetonitrile: (a) series of ethyl coumarin-3-carboxylates compounds 9a–e, 12b; (b) series of methyl coumarin-3-carboxylates compounds 10a–e; (c) series of ethyl 2-ethoxybenzo[e][1,2]oxaphosphorine-3-carboxylate 11a–d.
Figure 5. UV-VIS and normalized fluorescence spectra in acetonitrile: (a) series of ethyl coumarin-3-carboxylates compounds 9a–e, 12b; (b) series of methyl coumarin-3-carboxylates compounds 10a–e; (c) series of ethyl 2-ethoxybenzo[e][1,2]oxaphosphorine-3-carboxylate 11a–d.
Molecules 27 07649 g005aMolecules 27 07649 g005b
Figure 6. Normalized fluorescence spectra in acetonitrile for derivatives bearing para-methoxyphenyl groups.
Figure 6. Normalized fluorescence spectra in acetonitrile for derivatives bearing para-methoxyphenyl groups.
Molecules 27 07649 g006
Figure 7. Fluorescence intensity in acetonitrile according to the pH: (a) coumarin 3a (c = 10−5 M) behavior in sulfuric acid, 1N; (b) coumarin 3c (c = 10−5 M) behavior in sulfuric acid, 1N; (c) coumarin 3a (c = 10−5 M) behavior in NaOH, 1N; (d) coumarin 3c (c = 10−5 M) behavior in NaOH, 1N. The green arrow is showing the changing of the fluorescence intensity depending on the H+ concentration.
Figure 7. Fluorescence intensity in acetonitrile according to the pH: (a) coumarin 3a (c = 10−5 M) behavior in sulfuric acid, 1N; (b) coumarin 3c (c = 10−5 M) behavior in sulfuric acid, 1N; (c) coumarin 3a (c = 10−5 M) behavior in NaOH, 1N; (d) coumarin 3c (c = 10−5 M) behavior in NaOH, 1N. The green arrow is showing the changing of the fluorescence intensity depending on the H+ concentration.
Molecules 27 07649 g007aMolecules 27 07649 g007b
Table 1. Calculated excitation—λabs and emission wavelengths—λem, f—oscillator strengths, Stokes shifts.
Table 1. Calculated excitation—λabs and emission wavelengths—λem, f—oscillator strengths, Stokes shifts.
Calculated Models (CM)Structureλabs,
(nm)
fabsλem,
(nm)
femStokes Shift,
(nm)
CM-1Molecules 27 07649 i0012950.28003440.498949
CM-2Molecules 27 07649 i0023090.33053780.241969
CM-3Molecules 27 07649 i0033120.25743860.180274
CM-4Molecules 27 07649 i0042990.46653570.382158
CM-5Molecules 27 07649 i0053200.27034040.248084
CM-6Molecules 27 07649 i0063100.34463780.251368
CM-7Molecules 27 07649 i0072990.52673720.261173
CM-8Molecules 27 07649 i0083190.34043840.311465
CM-9Molecules 27 07649 i0093270.27034020.248075
CM-10Molecules 27 07649 i0103090.35693740.282666
CM-11Molecules 27 07649 i0113040.46673630.383259
CM-12Molecules 27 07649 i0123040.41953640.323061
Table 2. Ratio of the formed products in the crude reaction mixture *.
Table 2. Ratio of the formed products in the crude reaction mixture *.
EntrySolvent53a7Catalyst
1PhCH3traces>99tracesPdCl2(PPh3)2
2PhCH3:H2Otraces>99tracesPdCl2(PPh3)2
3Dioxane~14~86tracesPdCl2(PPh3)2
4Dioxane:H2Otraces>99tracesPdCl2(PPh3)2
5EtOH:H2O:PhCH3traces>99tracesPdCl2(PPh3)2
6PhCH3~54~45tracesPd(PPh3)4
7PhCH3:H2O18747Pd(PPh3)4
8Dioxane~54~45tracesPd(PPh3)4
9Dioxane:H2O25686Pd(PPh3)4
10EtOH:H2O:PhCH3~34~65tracesPd(PPh3)4
11PhCH3~57~42tracesIMesPd(dmba)Cl
12PhCH3:H2Otraces>99-IMesPd(dmba)Cl
13Dioxane~56~43tracesIMesPd(dmba)Cl
14Dioxane:H2Otraces>99tracesIMesPd(dmba)Cl
15EtOH:H2O:PhCH3traces>99tracesIMesPd(dmba)Cl
16PhCH33070-PEPPSI-type
17PhCH3:H2O-100-PEPPSI-type
18Dioxane2476-PEPPSI-type
19Dioxane:H2O-100-PEPPSI-type
20EtOH???PEPPSI-type
21EtOH:H2Otraces~77~23PEPPSI-type
22EtOH:H2O:PhCH3traces945PEPPSI-type
* The ratio is determined by NMR-spectroscopy.
Table 3. Parameters, reagents, and products of palladium-catalyzed arylation of 6-bromo-3-phosphonocoumarin 5.
Table 3. Parameters, reagents, and products of palladium-catalyzed arylation of 6-bromo-3-phosphonocoumarin 5.
EntryBoronic AcidProduct[Pd]Time, [h]Yield *, [%]
1Molecules 27 07649 i013Molecules 27 07649 i014PEPPSI-type17 h90%
2Molecules 27 07649 i015Molecules 27 07649 i016PEPPSI-type120 h41%
3Pd(PPh3)2Cl248 h63%
4Molecules 27 07649 i017Molecules 27 07649 i018PEPPSI-type48 h95%
5Molecules 27 07649 i019Molecules 27 07649 i020PEPPSI-type48 h61%
6Molecules 27 07649 i021Molecules 27 07649 i022PEPPSI-type48 h19%
7Pd(PPh3)2Cl248 h45%
* Isolated yields after column chromatography.
Table 4. Parameters, reagents, and products of palladium-catalyzed arylation of 6-bromocoumarin-3-carboxylates 8a,b.
Table 4. Parameters, reagents, and products of palladium-catalyzed arylation of 6-bromocoumarin-3-carboxylates 8a,b.
EntryBoronic AcidProduct[Pd]TimeYield *, [%]
1Molecules 27 07649 i023Molecules 27 07649 i024PEPPSI-type30 min90%
2PEPPSI-type60 min97%
3Molecules 27 07649 i025Molecules 27 07649 i026PEPPSI-type48 h20%
4Pd(PPh3)2Cl224 h68%
5Pd(PPh3)2Cl221 h83%
6Molecules 27 07649 i027Molecules 27 07649 i028PEPPSI-type45 min92%
7Molecules 27 07649 i029PEPPSI-type45 min99%
8Molecules 27 07649 i030Molecules 27 07649 i031PEPPSI-type45 min80%
9PEPPSI-type60 min99%
10Molecules 27 07649 i032Molecules 27 07649 i033Pd(PPh3)2Cl272 h29%
11Pd(PPh3)2Cl290 h18%
12Molecules 27 07649 i034Molecules 27 07649 i035Pd(PPh3)2Cl272 hcomplex mixture
13Pd(PPh3)2Cl272 hcomplex mixture
* Isolated yields after column chromatography.
Table 5. Parameters, reagents, and products of palladium-catalyzed arylation of ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphorine-3-carboxylate 6.
Table 5. Parameters, reagents, and products of palladium-catalyzed arylation of ethyl 6-bromo-2-ethoxybenzo[e][1,2]oxaphosphorine-3-carboxylate 6.
EntryBoronic AcidProduct[Pd]Time, [h]Yield *, [%]
1Molecules 27 07649 i036Molecules 27 07649 i037PEPPSI-type16 h89%
2Molecules 27 07649 i038Molecules 27 07649 i039Pd(PPh3)2Cl242 h45%
3Molecules 27 07649 i040Molecules 27 07649 i041PEPPSI-type3 h87%
4Molecules 27 07649 i042Molecules 27 07649 i043PEPPSI-type2.5 h89%
* Isolated yields after column chromatography.
Table 6. Parameters, reagents, and products of palladium-catalyzed alkynylation of 3-substituted coumarin species—Sonogashira reaction.
Table 6. Parameters, reagents, and products of palladium-catalyzed alkynylation of 3-substituted coumarin species—Sonogashira reaction.
EntryAlkyneProduct[Pd]Time, [h]Yield *, [%]
1Molecules 27 07649 i044Molecules 27 07649 i045Pd(PPh3)2Cl220 h82%
2Pd(PPh3)2Cl222 h52%
3Pd(PPh3)2Cl248 hcomplex mixture
4Molecules 27 07649 i046Molecules 27 07649 i047PEPPSI-type30 h56%
5Pd(PPh3)2Cl248 hcomplex mixture
* Isolated yields after column chromatography.
Table 7. Photophysical data of the coumarin derivatives.
Table 7. Photophysical data of the coumarin derivatives.
CompoundsSolventλabs, nmλem, nmStokes Shift, nm
7MeCN32441187
3aDMSO345460115
MeOH346459113
MeCN343453110
DCM344452108
EtOAc343447104
3bMeCN345493148
3cMeCN356535179
3dMeCN345456148
3eMeCN333451118
12aMeCN351461110
12dMeCN356543187
9-H *MeCN32641286
9aMeOH355470115
MeCN355463108
DCM36045797
9bMeCN353503150
9cMeCN366546180
9dMeCN36645892
9eMeCN343458115
12bMeCN357472115
10-H *MeCN32641185
10aMeOH357472115
MeCN353466113
DCM35845799
10bMeCN351504153
10cMeCN363543180
10dMeCN354466112
10eMeCN340460120
11-H *MeCN32441086
11aMeOH343468125
MeCN342459117
DCM342452110
11bMeCN340500160
11cMeCN350541191
11dMeCN341460119
* unsubstituted at C-6 in the benzene moiety compounds.
Table 8. Comparison between the experimental and calculated spectral characteristics. Scaling factor for the absorbance—1.098 and 1.195—for the emission.
Table 8. Comparison between the experimental and calculated spectral characteristics. Scaling factor for the absorbance—1.098 and 1.195—for the emission.
ExpCM λ a b s s c a l e d Δ λ a b s s c a l e d λ a b s e x p λ e m s c a l e d Δ λ e m s c a l e d λ e m e x p Δ S t o k e s   s h i f t ( c a l c e x p )
7CM-132404110−38
3aCM-2339−44521−41
3bCM-3343−246132−74
3cCM-5351−548352−95
3dCM-6340−545241−43
3eCM-7328−54447−45
12aCM-8350−14592−45
12dCM-9359348062−112
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lyapchev, R.; Koleva, A.I.; Koleva, I.Z.; Subev, K.; Madzharova, I.; Simeonova, K.B.; Petkova-Yankova, N.; Morgenstern, B.; Lozanova, V.; Petrov, P.Y.; et al. Efficient Synthesis of Fluorescent Coumarins and Phosphorous-Containing Coumarin-Type Heterocycles via Palladium Catalyzed Cross-Coupling Reactions. Molecules 2022, 27, 7649. https://doi.org/10.3390/molecules27217649

AMA Style

Lyapchev R, Koleva AI, Koleva IZ, Subev K, Madzharova I, Simeonova KB, Petkova-Yankova N, Morgenstern B, Lozanova V, Petrov PY, et al. Efficient Synthesis of Fluorescent Coumarins and Phosphorous-Containing Coumarin-Type Heterocycles via Palladium Catalyzed Cross-Coupling Reactions. Molecules. 2022; 27(21):7649. https://doi.org/10.3390/molecules27217649

Chicago/Turabian Style

Lyapchev, Rumen, Ana I. Koleva, Iskra Z. Koleva, Kristian Subev, Ivelina Madzharova, Kristina B. Simeonova, Nevena Petkova-Yankova, Bernd Morgenstern, Vesela Lozanova, Petar Y. Petrov, and et al. 2022. "Efficient Synthesis of Fluorescent Coumarins and Phosphorous-Containing Coumarin-Type Heterocycles via Palladium Catalyzed Cross-Coupling Reactions" Molecules 27, no. 21: 7649. https://doi.org/10.3390/molecules27217649

Article Metrics

Back to TopTop