Next Article in Journal
The Fabrication of Cu2O-u/g-C3N4 Heterojunction and Its Application in CO2 Photoreduction
Previous Article in Journal
Morphological Regulation of Bi5O7I for Enhanced Efficiency of Rhodamine B Degradation Under Visible-Light
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis of 4-(Methoxycarbonyl)phenyl 5-Arylfuran-2-Carboxylates via Readily Available Pd Catalyst–Their Thermodynamic, Spectroscopic Features and Nonlinear Optical Behavior

1
Department of Chemistry, Government College University Faisalabad, Faisalabad 38000, Pakistan
2
School of Chemistry and Chemical Engineering, Nanjing University of Science and Technology, Nanjing 210094, China
3
Department of Chemistry, Faculty of Science, Imam Mohammad Ibn Saud Islamic University, Riyad 11623, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(8), 713; https://doi.org/10.3390/catal15080713
Submission received: 13 May 2025 / Revised: 20 July 2025 / Accepted: 24 July 2025 / Published: 26 July 2025
(This article belongs to the Special Issue Transition-Metal-Catalyzed Organic Synthesis)

Abstract

In this work, we described the synthesis of 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate by reacting 5-bromofuroic acid with methylparaben in the incorporation of DCC/DMAP (Steglich esterification) as coupling agents. Later on, we subsequently synthesized a series of 4-(methoxycarbonyl)phenyl 5-aryl furan-2-carboxylates (5a5e) through Suzuki coupling catalyzed by palladium (0) between 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3) with several substituted arylated and heteroaryl boronic acids (4). DFT calculations were computed to elucidate electronic structural features of synthesized molecules (5a5e) and to validate these findings by correlating with theoretical and experimental spectroscopic analysis. Furthermore, geometrical optimization, thermodynamic features, as FMO orbitals, MESP maps, NLO behavior and reactivity descriptors, were also determined from the PBE0 D3BJ/def2-TZVP/SMD1,4-dioxane theory level to confirm the structural features of synthesized molecules.

1. Introduction

The development of high-performance NLO materials is crucial for advancing next-generation optical communication and laser systems, making the exploration of molecular frameworks with superior NLO properties a key area of research in material science and computational chemistry [1,2,3]. Nonlinear optical (NLO) materials are important in electro-optical switching, optical confinements, commercial lasers, logic devices, telecommunications, harmonic generation, environmental monitoring, materials processing, and biomedical imaging [4,5,6,7,8,9,10,11]. These NLO materials have numerous applications in optical and electro-optical systems including optical data storage, optical communication, optical signal processing, optical data processing devices [12,13,14], format switching [15], all-optical data conversion [16], Biophotonic Applications, and wavelength/sign conversion [17]. Most NLO materials that are sold commercially are inorganic solids. Organic materials, on the other hand, may have numerous benefits regarding processing capability, tenability, and response time [18]. Organic NLO materials have recently piqued the curiosity of researchers. Researchers have been interested in the study of NLO response (hyperpolarizabilities) of organic compounds for the past ten years [19,20]. The key parameter governing a molecule’s NLO behavior is its hyperpolarizability (β), optimal HOMO-LUMO energy gaps, and extended π-conjugation, which quantifies its ability to interact with and modulate light.
Furan-based molecules are key intermediates in the organic NLO materials synthesis due to their extended conjugation, lower band gaps, higher delocalization, and polarizability [21,22,23,24,25]. To date, for preparing esters, several methodologies have been reported. Mujahid et al. reported in 2022, a series of thiophene ester derivatives with Steglich esterification, as well as thiophene carboxylic acid and different halogenated alcohols [26]. In 2020, Rasool and coauthors described the synthesis of halogenated esters using N, N-dicyclohexylcarbodiimide (DCC), and 4-dimethylaminopyridine (DMAP) [27]. Furthermore, although beyond the scope of our research, Suzuki-Miyaura cross-coupling reactions (SMCs) are frequently utilized to synthesize biaryl moieties with a wide range of applications in materials science [28,29]. The arylation of organic compounds by coupling also plays a vital role in synthesis. Transition metal catalysis is well known for the C–C coupling for the arylation in asymmetric synthesis. Suzuki coupling (SMC) via Pd catalyst with aryl halides and arylboronic acids gives a highly effective and efficient method for designing new compounds. Boronic acids are nontoxic, chemically inert to moisture/air, thermal stability, and they are easier to handle than other commonly utilized coupling reagents [30,31,32].
This research study emphasizes the synthesis of 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3) and its arylated derivatives (5a5e) through SMC reaction. Computational studies were applied to calculate the chemical and thermodynamic features, reactivity parameters, and nonlinear optical (NLO) behavior. Further, the 1H NMR chemical shifts were compared with theoretical values.

2. Results and Discussion

In the current study, 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3) was successfully synthesized through the reaction of 5-bromofuroic acid with methylparaben in the presence of Steglich esterification, resulting in good yields of 88% (Scheme 1). Herein, DCC (N, N′-Dicyclohexylcarbodiimide) served as the coupling partner, and DMAP (dimethylaminopyridine) acted as the catalyst in the process mentioned above using acetonitrile as a solvent stirred at 80 °C for 2 h. Noteworthy, DCC/DMAP gives excellent yields with minimum side products through easy removal of side products. Nevertheless, solvent effects are not negligible. It was also noted that the maximum compound yield (3) was obtained with acetonitrile (Table 1). Additionally, 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3) was subjected to Suzuki–coupling with several aryl/heteroaryl boronic acids in Pd(PPh3)4 catalyst and K3PO4 base, yielding high yields of the desired compounds (5a5e). To the best of our knowledge, these synthesized molecules are new. Bulkier tetrakis(triphenylphosphine) ligand insertion leads to coupling to the palladium center-promoted heterocoupling between the reactive species over homocoupling, increasing the reaction’s feasibility. The coupling reaction was carried out in a dioxane-water solvent mixture because polar solvents coordination with the palladium complexes during the transition state, often leads to better reaction yields.
Aryl boronic acids comprising electron-withdrawing groups exhibited reduced nucleophilicity, resulting in slower transmetalation than their neutral counterparts. This decreased reactivity makes them more susceptible to side reactions like homocoupling. Similarly, sterically hindered boronic acids possess lower yields due to a reduced availability at the reactive site. It was observed that the electron-rich ring activating group (5a) gives a good yield, but the electron-deficient ring deactivating groups give moderate product yields (5c5e). The low yield (5c) may be associated with challenging purification by column chromatography (Scheme 2).
Reagents and conditions: 1 (0.51 g, 0.4 eq. 4.18 mmol), 2 (1.60 g, 1 eq. 10.47 mmol), DMAP (0.51 g, 0.4 eq. 4.18 mmol), DCC (2.38 g, 1.1 eq. 11.52 mmol), stirred for 2 h at 80 °C.
Reagents and conditions: 3 (0.2 g, 0.61 mmol, 1 eq), Pd(PPh3)4 (0.08 g, 7 mol%), K3PO4 (0.26 g, 1.22 mmol, 2 eq.), aryl boronic acids (1.1 eq., 0.67 mmol), 1,4-dioxane (4 mL), H2O (0.5 mL), reflux, 24 h.

2.1. DFT Studies

The thermodynamic and chemical characteristics were determined using computational studies utilizing the DFT approach. The three-dimensional models of all the furan carboxylates 5a5e were created via Gauss view [33] and then optimized at Gaussian 09W [34] at the PBE0-D3BJ/def2-TZVP/SMD1,4-dioxane level of theory [35]. Geometry optimization following the calculation of vibrational frequencies using the previously described theory level. Unconstrained geometry optimization was performed on all the resultant conformers and then NMR data and Boltzmann averages were calculated for all conformers [36]. To explore the NLO behaviors of all compounds, hyperpolarizability has also been calculated according to the formulae given by using the above-mentioned level of theory. The optimized geometries of all the compounds are given in Figure 1. Furthermore, conformational analysis was carried out by using five distinct dihedral angles, as illustrated in Figure S1, with each dihedral angle (atoms mentioned in Figure S1) in steps of 30° over one complete rotation (0–360°) of compound 5a. The energy value of all conformers of 5a are given in Table S1 in supporting information.

2.2. Analysis of FMO (Frontier Molecular Orbital) and Hyperpolarizability:

FMO studies can derive valuable statistics about a compound’s chemical reactivity and other features from the HOMO and LUMO energies [37]. Filled HOMOs and empty LUMOs energy gaps are important quantum chemical parameters that influence a molecule’s chemical stability and reactivity. It is estimated that compounds with the largest HOMO and LUMO energy gap are least reactive and the most stable, whereas molecules with the minimum energy gaps are unstable and the most reactive compounds. The evaluation of FMO surface plots to experimental data offers vital information about the computational method’s capacity to pronounce the molecule’s chemical reactivity [38,39]. The FMOs surface plots for molecules 5a5e are shown in Figure 2. Table 2 shows HOMO, LUMO energies and their energy gaps, as well as their hyperpolarizability values. In this series, compound 5b and 5c consists of furoic ester and chloro substitution at meta-para position shows the least energy gap of 4.52 eV which leads to showing as the most reactive compound. The compounds 5d and 5e consist of furan group linked with electron withdrawing biaryl substitution shows the highest energy gap of 4.60 eV leading to the most stable compound.
Nowadays nonlinear optics has become an interesting field because of its numerous uses in the area of data storage, optical communication, and laser technology [22,25,35]. Molecules exhibit high hyperpolarizability, resulting in a strong NLO behavior and vice versa [21,23,40]. These synthesized arylated furan carboxylates exhibited extended π-conjugation and various electron-donating or electron-withdrawing moieties on the aryl boronic acid, which influence their frontier molecular orbital (FMO) energies and β values. In this series, molecule 5b contains furoic ester and biphenyl scaffolds, which is subsequently associated with the chloro substitution at the para position and exhibits highest hyperpolarizability of 4396.08 Hartree, exhibiting a high NLO behavior in contrast to the other molecules in this series (5a5e). However, 5e shows the highest energy gap with the lowest hyperpolarizability value of 3859.80 Hartree, showing the least NLO responses (Table 2). The calculated first hyperpolarizability for compound 5b (β_total = 4396.08 a.u.) is significantly high compared to previously documented furan esters, which typically exhibit β values below approximately 3200 a.u. Furthermore, this value is comparable to efficient donor-π-acceptor organic chromophores used in nonlinear optics that exhibit β values in the range of 2000–10,000 a.u. [25,41,42,43,44]. In contrast, molecular building blocks of common inorganic NLO crystals often have much smaller β values (100–500 a.u.) due to their rigid frameworks and limited conjugation [45,46,47]. Such a high hyperpolarizability confirms the strong intramolecular charge transfer characteristics of 5b, validating its potential as a high-performance organic NLO material.

2.3. Molecular Electrostatic Potential (MEP) Analysis

A three-dimensional tool used to assess the charge distribution, size, and overall shape of a molecule. It shows graphical representation of molecule relative polarities. MEP can be employed to determine nucleophilic reactions, electrophilic attacks, and hydrogen bonding interactions. In the MEP maps, the electrophilic reactivity (electron-rich) sites are indicated with red color, while the nucleophilic reactivity (electron-deficient) sites with blue color [48,49]. MEP is an invaluable tool for gaining insights into molecular interactions. Noteworthy, under study molecules (5a5e), negative charge is particularly present at the O-atom of the ester as the reactive site for electrophilic attack. In contrast, the positive region is particularly present at the H atoms of the biphenyls, acts as reactive sites for nucleophilic attack. MEP maps of all of the synthesized molecules (5a5e) are shown in Figure 3 [50].

2.4. Reactivity Parameters

Investigating reactivity parameters such as ionization potential (I), electron affinity (A), chemical hardness (η), chemical potential (μ), and electrophilicity index (ω) is important for predicting the chemical behavior, electron-accepting ability, and optical characteristics of molecular systems. These descriptors specifically aid in describing intrinsic electronic reactivity [51], interactions with electrophiles and nucleophiles, and overall stability in polar regions—elements crucial for designing molecules for nonlinear optics (NLO) [52,53]. Furthermore, chemical reactivity is a crucial idea since it is intimately tied to reaction mechanisms, making it possible to comprehend chemical reactions and enhance synthetic processes to produce new materials. To compute the electron affinity (A = −ELUMO) and ionization potential (I = −EHOMO), Koopman’s theorem was used. The reactivity parameters, including chemical softness (σ), chemical hardness (η), electronic chemical potential (μ), and electrophilicity index (ω) are determined from Equations (1)–(4) [48,53].
η = (EHOMOELUMO)/2
μ = − (EHOMO + ELUMO)/2
ω = μ2/2 η
σ = 1/η
The fundamental ideas known as global reactivity parameters are chemical hardness and softness, which are computed by the use of DFT. The energy difference can predict the softness and hardness of furan carboxylates (5a5e). A big energy gap results in a hard molecule, while a small energy gap results in a soft molecule [54]. High electrophilicity (ω) and negative chemical potential (μ) indicate a greater ability to accept electrons, making it favorable in NLO applications. In this series, compound 5b exhibits the lowest hardness (η = 2.25 eV) and a relatively high electrophilicity (ω = 3.84 eV), indicating its least stability and highest reactivity in the series. Compound 5c shows the highest ionization potential (I = 6.54 eV) and electrophilicity index (ω = 4.05 eV), coupled with the most negative chemical potential (μ = –4.28 eV), suggesting a high tendency to accept electrons and a strong electrophilic character. Compound 5e combines high ionization potential (6.52 eV) and hardness (2.30 eV) with a significant electrophilicity index (3.87 eV), making it electronically stable yet fairly reactive toward nucleophiles. Compound 5a exhibits moderate reactivity across all parameters, while 5d and 5e, with maximum chemical hardness (η = 2.30 eV) and lowest softness (σ = 0.43 eV−1), are identified as the least reactive and electronically stable molecules (Table 3). These findings align with the energy gap trends and give the following reactivity order:
5b > 5c > 5e > 5a > 5d.
All synthesized compounds—including chemical hardness (η = 2.25–2.30 eV), electrophilicity index (ω = 3.54–4.05 eV), and softness (σ = 0.43–0.44 eV−1)—fit well within the ranges reported for analogous D–π–A and furan-based NLO systems (η: 2.1–2.6 eV, ω: 3.0–4.5 eV, σ: 0.40–0.45 eV−1) [23,55,56,57]. Ionization potentials and electron affinities of all compounds (I = 6.36–6.54 eV, A = 1.76–2.08 eV) also align closely with values reported in the literature (I: 6.0–7.0 eV, A: 1.0–2.0 eV) [58,59]. This comparison confirms the suitability of synthesized derivatives for NLO and electronic applications, as their reactivity profiles reflect the optimal balance of stability and electron-accepting ability observed in nonlinear optics.

3. Materials and Methods

3.1. General Remarks

Commercial-grade solvents were used, and all starting materials were bought from Alfa Aesar (Ward Hill, MA, USA) and Sigma Aldrich (St. Louis, MO, USA). For melting point, Buchi equipment was used (New Castle, B-540, Büchi Labortechnik AG, Flawil, Switzerland). To get NMR spectra, deuterated CDCl3, solvent was utilized on a NMR spectrophotometer (Billercia, Bruker, MA, USA). Standard techniques were used to prevent sensitive reagents. An inert surrounding was provided for the conduction of moisture sensitive reactions, especially in the presence of argon as an inert medium. The chemicals underwent column chromatography and silica gel (70–230 mesh) for purification.60 PF254 silica gel cards from Merck were used as TLC to monitor the progress of reactions. The result was seen on TLC using a UV lamp (254–365 nm) from Scientific Store (Lahore, Pakistan).

3.2. Experimental Protocol for Synthesis of 4-(Methoxycarbonyl)phenyl 5-Bromofuran-2-Carboxylate (3)

A calculated amount of 5-bromofuroic acid (1) (2 g, 1 eq., 10.47 mmol) and DMAP (0.51 g, 0.4 eq., 4.18 mmol) was taken in a round-bottom flask with a magnetic stirrer. After stirring for half an hour, methylparaben (2) (1.60 g, 1 eq., 10.47 mmol) was added to the reaction mixture. The mixture was stirred further after the addition of DCC (2.38 g, 1.1 equiv., 11.52 mmol). The reaction was allowed to proceed with continuous stirring for 24 h at room temperature [1,60] Completion of the reaction was monitored by thin-layer chromatography (TLC). The desired product, 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3), was purified by column chromatography using an n-hexane/ethyl acetate (10:90) ratio. The compound was subsequently characterized by 1H-NMR and 13C-NMR.

3.3. General Protocol for Synthesis of 4-(Methoxycarbonyl)phenyl 5-Bromofuran-2-Carboxylate Derivatives (5a5e)

In the Schlenk tube containing 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3) (0.2 g, 0.61 mmol, 1 eq), palladium catalyst Pd(PPh3)4 (0.08 g, 7 mol%), and 4 mL of 1,4-dioxane were added. The mixture was continuously stirred at room temperature for half an hour under inert environmental conditions. Then, K3PO4 base (0.26g, 1.22 mmol, 2 eq.), and aryl/heteroaryl boronic acids (1.1 eq., 0.67 mmol), and H2O (0.5 mL) were incorporated in followed reaction with the provision of an inert atmosphere to the mixture and stirred the reaction at 85–90 °C [32,61]. After 24 h of consecutive reaction, the reaction material was analyzed by TLC, and the mixture was subjected to filtration and subsequent washing with distilled water. Ethyl acetate was added to dilute the reaction mixture, and the organic layer was separated. The sample solvent was then dried and purified through column chromatography. Finally, the sample was dried and recrystallized. The sample was further characterized using 1H-NMR and 13C-NMR spectroscopic techniques.

3.4. Characterization

3.4.1. 4-(Methoxycarbonyl)phenyl 5-Bromofuran-2-Carboxylate (3)

Yield 88%, white crystals; M.P. 99–101 °C. 1H NMR (400 MHz, CDCl3) δ 8.09 (d, J = 8.8 Hz, 2H), 7.33 (d, J = 3.5 Hz, 1H), 7.28 (d, J = 8.7 Hz, 2H), 6.54 (d, J = 3.6 Hz, 1H), 3.91 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 166.2, 155.1, 153.6, 145.1, 131.3, 129.1, 128.0, 122.0, 121.5, 114.4, 52.2. Elemental Anal. Calcd. for C13H9BrO5 (325.11): C, 48.03; H, 2.79% Found: C, 48.09; H, 2.77%.

3.4.2. 4-(Methoxycarbonyl)phenyl 5-(3,5-Dimethylphenyl)furan-2-Carboxylate (5a)

Yield 81%, light yellow crystals; M.P. 109–110 °C. 1H NMR (400 MHz, CDCl3) δ 8.10 (d, J = 8.7 Hz, 2H), 7.48–7.40 (m, 3H), 7.31 (d, J = 8.7 Hz, 2H), 7.01 (s, 1H), 6.78 (d, J = 3.7 Hz, 1H), 3.91 (s, 3H), 2.35 (s, 6H). 13C NMR (101 MHz, CDCl3) δ 166.4, 159.3, 156.4, 154.0, 142.1, 138.5, 131.7, 131.2, 128.9, 127.7, 122.8, 122.2, 121.7, 115.3, 107.1, 52.2, 21.3. Elemental Anal. Calcd. for C21H18O5 (350.37): C, 71.99; H, 5.18% Found: C, 71.93; H, 5.17%.

3.4.3. 4-(Methoxycarbonyl)phenyl 5-(4-Chlorophenyl)furan-2-Carboxylate (5b)

Yield 72%, Reddish brown crystals; M.P. 115–116 °C. 1H NMR (400 MHz, CDCl3) δ 8.11 (d, J = 8.3 Hz, 2H), 7.75 (d, J = 8.2 Hz, 2H), 7.45 (d, J = 3.6 Hz, 1H), 7.41 (d, J = 8.3 Hz, 2H), 7.31 (d, J = 8.3 Hz, 2H), 6.81 (d, J = 3.7 Hz, 1H), 3.92 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 166.3, 157.6, 156.2, 153.9, 142.7, 135.3, 131.2, 129.2, 127.9, 127.7, 126.2, 122.0, 121.6, 107.6, 52.2. Elemental Anal. Calcd. for C19H13ClO5 (356.76): C, 63.97; H, 3.67% Found: C, 63.96; H, 3.59%.

3.4.4. 4-(Methoxycarbonyl)phenyl 5-(3,4-Dichlorophenyl)furan-2-Carboxylate (5c)

Yield 69%, Greyish crystals; M.P. 119–121 °C. 1H NMR (400 MHz, CDCl3) δ 8.11 (d, J = 8.7 Hz, 1H), 7.69 (dd, J = 5.7, 3.4 Hz, 2H), 7.64–7.59 (m, 1H), 7.54–7.48 (m, 2H), 7.45 (d, J = 3.6 Hz, 1H), 7.31 (d, J = 8.7 Hz, 1H), 6.88–6.82 (m, 1H), 3.92 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 167.8, 156.2, 153.8, 143.1, 132.4, 131.7, 131.3, 130.9, 130.6, 128.8, 127.9, 126.7, 124.1, 121.6, 117.5, 115.4, 108.4, 52.2. Elemental Anal. Calcd. for C19H12Cl2O5 (391.20): C, 58.34; H, 3.09% Found: C, 58.39; H, 3.01%.

3.4.5. 4-(Methoxycarbonyl)phenyl 5-(Thiophen-3-yl)furan-2-Carboxylate (5d)

Yield 76%, Pale yellow solids; M.P. 108–109 °C. 1H NMR (400 MHz, CDCl3) δ 8.10 (d, J = 8.6 Hz, 2H), 7.75 (dd, J = 2.9, 1.2 Hz, 1H), 7.51 (dd, J = 5.7, 3.3 Hz, 1H), 7.42 (t, J = 3.5 Hz, 1H), 7.37 (dd, J = 5.1, 2.9 Hz, 1H), 7.30 (d, J = 8.4 Hz, 2H), 6.64 (d, J = 3.6 Hz, 1H), 3.91 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 167.7, 166.3, 156.3, 155.5, 153.9, 141.8, 131.2, 128.8, 126.9, 124.9, 122.7, 122.1, 121.6, 107.1, 52.2. Elemental Anal. Calcd. for C17H12O5S (328.34): C, 62.19; H, 3.68% Found: C, 62.21; H, 3.69%.

3.4.6. 4-(Methoxycarbonyl)phenyl 5-(3-Chloro-4-Fluorophenyl)furan-2-Carboxylate (5e)

Yield 73%, Greenish solids; M.P. 117–118 °C. 1H NMR (400 MHz, CDCl3) δ 7.90 (d, J = 8.8 Hz, 2H), 7.69 (dd, J = 5.7, 3.3 Hz, 1H), 7.67–7.61 (m, 1H), 7.54 (dtd, J = 14.2, 6.5, 2.4 Hz, 1H), 7.46 (td, J = 7.6, 2.9 Hz, 1H), 7.23–7.07 (m, 1H), 6.88 (d, J = 8.8 Hz, 2H), 3.86 (s, 3H). 13C NMR (101 MHz, CDCl3) δ 168.1, 167.5, 160.9, 132.5, 132.3, 132.1, 132.0, 131.8, 131.6, 131.3, 131.0, 130.5, 128.8, 128.7, 121.7, 115.3, 51.9. Elemental Anal. Calcd. for C19H12ClFO5 (374.75): C, 60.90; H, 3.23% Found: C, 60.90; H, 3.23%.

4. Conclusions

A set of 4-(methoxycarbonyl)phenyl 5-arylfuran-2-carboxylates (5a5e) were synthesized in this study via SMC reactions. DFT computations were employed to elucidate the geometric characteristics and other physical properties of synthetic compounds (5a5e). These qualities encompass NMR and other chemical reactivity descriptors, such as ionization potential, electron affinity, chemical hardness, electron chemical potential, and electrophilic index. NMR calculations demonstrated a higher level of agreement with the experimental results, hence enhancing confidence in predicting NMR data. Molecule stability and NLO behavior were examined by calculations of energy gaps of HOMO-LUMO orbitals and hyperpolarizability. 5b was discovered to be highly reactive molecule, with the least energy gap of 4.52 eV. Compound 5b demonstrates a high degree of electronic polarizability and an enhanced intramolecular charge transfer character, leading to a first hyperpolarizability value of 4396.08 Hartree. These findings highlight the potential of 5b as a promising candidate for future nonlinear optical applications. The compound 5e has the highest energy gap (4.60 eV) which shows the most stability and least NLO responses with a value of 3859.80 Hartree. Future DFT-based NLO properties supported by experimental validation can further reveal the scope of these compounds as advanced materials in the field of nonlinear optics.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15080713/s1, Table S1: Energy values of all conformers of compound 5a; Table S2: Comparison of 1H NMR data of synthesized derivative 5a; Table S3: Comparison of 1H NMR data of synthesized derivative 5b; Table S4: Comparison of 1H NMR data of synthesized derivative 5c; Table S5: Comparison of 1H NMR data of synthesized derivative 5d; Table S6: Comparison of 1H NMR data of synthesized derivative 5e; Figure S1: Conformer 1 of compound 5a; Figure S2: 1H NMR of 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3); Figure S3: 13C NMR of 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3); Figure S4: 1H NMR of 4-(methoxycarbonyl)phenyl 5-(3,5-dimethylphenyl)furan-2-carboxylate (5a); Figure S5: 13C NMR of 4-(methoxycarbonyl)phenyl 5-(3,5-dimethylphenyl)furan-2-carboxylate (5a); Figure S6: 1H NMR of 4-(methoxycarbonyl)phenyl 5-(4-chlorophenyl)furan-2-carboxylate (5b); Figure S7: 13C NMR of 4-(methoxycarbonyl)phenyl 5-(4-chlorophenyl)furan-2-carboxylate (5b); Figure S8: 1H NMR of 4-(methoxycarbonyl)phenyl 5-(3, 4-dichlorophenyl)furan-2-carboxylate (5c); Figure S9: 13C NMR of 4-(methoxycarbonyl)phenyl 5-(3, 4-dichlorophenyl)furan-2-carboxylate (5c); Figure S10: 1H NMR of 4-(methoxycarbonyl)phenyl 5-(thiophen-3-yl)furan-2-carboxylate (5d); Figure S11: 13C NMR of 4-(methoxycarbonyl)phenyl 5-(thiophen-3-yl)furan-2-carboxylate (5d); Figure S12: 1H NMR of 4-(methoxycarbonyl)phenyl 5-(3-chloro-4-fluorophenyl)furan-2-carboxylate (5e); Figure S13: 13C NMR of 4-(methoxycarbonyl)phenyl 5-(3-chloro-4-fluorophenyl)furan-2-carboxylate (5e) [62].

Author Contributions

Conceptualization, N.R. and M.E.A.Z.; Data curation, M.S. and A.M. (Ayesha Malik); Formal analysis, A.K., M.A. and A.M. (Ayesha Malik); Funding acquisition, M.E.A.Z.; Investigation, S.A.A.-H.; Methodology, A.M. (Adeel Mubarik), N.R., M.A. and A.M. (Ayesha Malik); Project administration, N.R., S.A.A.-H. and M.E.A.Z.; Resources, M.S.; Software, A.M. (Adeel Mubarik); Supervision, N.R.; Validation, M.A.; Writing—original draft, M.F.U.Z. and A.M. (Adeel Mubarik); Writing—review & editing, M.F.U.Z. and A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article and the Supplementary Materials.

Acknowledgments

The present data are part of Muhammad Fakhar U Zaman. The authors acknowledge PCSIR (Ministry of Science and Technology) for providing access to NMR facilities used in this study. The authors also express their gratitude to the Deanship of Scientific Research, Imam Mohammad Ibn Saud Islamic University (IMSIU), Saudi Arabia to support this research work.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Nazeer, W.; Qamar, M.U.; Rasool, N.; Taibi, M.; Salamatullah, A.M. Synthesis of 2-Ethylhexyl 5-Bromothiophene-2-Carboxylates; Antibacterial Activities against Salmonella Typhi, Validation via Docking Studies, Pharmacokinetics, and Structural Features Determination through DFT. Molecules 2024, 29, 3005. [Google Scholar] [CrossRef] [PubMed]
  2. Jensen, F. Introduction to Computational Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 2017. [Google Scholar]
  3. Goh, G.B.; Hodas, N.O.; Vishnu, A. Deep learning for computational chemistry. J. Comput. Chem. 2017, 38, 1291–1307. [Google Scholar] [CrossRef] [PubMed]
  4. Hales, J.M.; Matichak, J.; Barlow, S.; Ohira, S.; Yesudas, K.; Brédas, J.-L.; Perry, J.W.; Marder, S.R. Design of polymethine dyes with large third-order optical nonlinearities and loss figures of merit. Science 2010, 327, 1485–1488. [Google Scholar] [CrossRef] [PubMed]
  5. Tian, X.; Wei, R.; Guo, Q.; Zhao, Y.J.; Qiu, J. Reverse Saturable Absorption Induced by Phonon-Assisted Anti-Stokes Processes. Adv. Mater. 2018, 30, 1801638. [Google Scholar] [CrossRef]
  6. Guo, Q.; Yao, Y.; Luo, Z.-C.; Qin, Z.; Xie, G.; Liu, M.; Kang, J.; Zhang, S.; Bi, G.; Liu, X. Universal near-infrared and mid-infrared optical modulation for ultrafast pulse generation enabled by colloidal plasmonic semiconductor nanocrystals. ACS Nano 2016, 10, 9463–9469. [Google Scholar] [CrossRef]
  7. Zhang, Q.; Li, L.; Zhang, M.; Liu, Z.; Wu, J.; Zhou, H.; Yang, J.; Zhang, S.; Tian, Y. Nonlinear optical response and biological applications of a series of pyrimidine-based molecules for copper (II) ion probe. Dalton Trans. 2013, 42, 8848–8853. [Google Scholar] [CrossRef]
  8. González-Urbina, L.; Baert, K.; Kolaric, B.; Pérez-Moreno, J.; Clays, K. Linear and nonlinear optical properties of colloidal photonic crystals. Chem. Rev. 2012, 112, 2268–2285. [Google Scholar] [CrossRef] [PubMed]
  9. Liu, J.; Ouyang, C.; Huo, F.; He, W.; Cao, A. Progress in the enhancement of electro-optic coefficients and orientation stability for organic second-order nonlinear optical materials. Dyes Pigments 2020, 181, 108509. [Google Scholar] [CrossRef]
  10. Xu, H.; Liu, F.; Elder, D.L.; Johnson, L.E.; de Coene, Y.; Clays, K.; Robinson, B.H.; Dalton, L.R. Ultrahigh electro-optic coefficients, high index of refraction, and long-term stability from Diels–Alder cross-linkable binary molecular glasses. Chem. Mater. 2020, 32, 1408–1421. [Google Scholar] [CrossRef]
  11. Khan, S.A.; Razvi, M.A.; Bakry, A.H.; Afzal, S.M.; Asiri, A.M.; El-Daly, S.A. Microwave assisted synthesis, spectroscopic studies and non linear optical properties of bis-chromophores. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 137, 1100–1105. [Google Scholar] [CrossRef]
  12. Sun, W.-M.; Li, X.-H.; Wu, D.; Li, Y.; He, H.-M.; Li, Z.-R.; Chen, J.-H.; Li, C.-Y. A theoretical study on superalkali-doped nanocages: Unique inorganic electrides with high stability, deep-ultraviolet transparency, and a considerable nonlinear optical response. Dalton Trans. 2016, 45, 7500–7509. [Google Scholar] [CrossRef]
  13. Ullah, F.; Kosar, N.; Ayub, K.; Mahmood, T. Superalkalis as a source of diffuse excess electrons in newly designed inorganic electrides with remarkable nonlinear response and deep ultraviolet transparency: A DFT study. Appl. Surf. Sci. 2019, 483, 1118–1128. [Google Scholar] [CrossRef]
  14. Israr, H.; Rasool, N.; Rizwan, K.; Hashmi, M.A.; Mahmood, T.; Rashid, U.; Hussein, M.Z.; Akhtar, M.N. Synthesis and reactivities of triphenyl acetamide analogs for potential nonlinear optical material uses. Symmetry 2019, 11, 622. [Google Scholar] [CrossRef]
  15. Wu, C.L.; Su, S.P.; Lin, G.R. All-optical modulation based on silicon quantum dot doped SiOx: Si-QD waveguide. Laser Photonics Rev. 2014, 8, 766–776. [Google Scholar] [CrossRef]
  16. Lin, G.-R.; Su, S.-P.; Wu, C.-L.; Lin, Y.-H.; Huang, B.-J.; Wang, H.-Y.; Tsai, C.-T.; Wu, C.-I.; Chi, Y.-C. Si-rich SiNx based Kerr switch enables optical data conversion up to 12 Gbit/s. Sci. Rep. 2015, 5, 9611. [Google Scholar] [CrossRef]
  17. Wu, C.-L.; Lin, Y.-H.; Su, S.-P.; Huang, B.-J.; Tsai, C.-T.; Wang, H.-Y.; Chi, Y.-C.; Wu, C.-I.; Lin, G.-R. Enhancing optical nonlinearity in a nonstoichiometric SiN waveguide for cross-wavelength all-optical data processing. ACS Photonics 2015, 2, 1141–1154. [Google Scholar] [CrossRef]
  18. Verbiest, T.; Houbrechts, S.; Kauranen, M.; Clays, K.; Persoons, A. Second-order nonlinear optical materials: Recent advances in chromophore design. J. Mater. Chem. 1997, 7, 2175–2189. [Google Scholar] [CrossRef]
  19. Vaz, W.F.; Custodio, J.M.; Silveira, R.G.; Castro, A.N.; Campos, C.E.; Anjos, M.M.; Oliveira, G.R.; Valverde, C.; Baseia, B.; Napolitano, H.B. Synthesis, characterization, and third-order nonlinear optical properties of a new neolignane analogue. RSC advances 2016, 6, 79215–79227. [Google Scholar] [CrossRef]
  20. Zhao, B.; Lu, W.-Q.; Zhou, Z.-H.; Wu, Y. The important role of the bromo group in improving the properties of organic nonlinear optical materials. J. Mater. Chem. 2000, 10, 1513–1517. [Google Scholar] [CrossRef]
  21. Zhang, M.; Deng, G.; Zhang, A.; Xu, H.; Huang, H.; Peng, C.; Bo, S.; Liu, X.; Zhen, Z.; Qiu, L. Synthesis and properties of a new second-order NLO chromophore containing the benzo [b] furan moiety for electro-optical materials. RSC Adv. 2014, 4, 33312–33318. [Google Scholar] [CrossRef]
  22. Shehzadi, K.; Tariq, A.; Zubair, M.; Mahmood, T.; Kosar, N.; Karakaya, I.; Rasool, N.; Rizwan, K. Synthesis of pyridine and furan based arylated ketones through palladium catalyst with DFT study of their static and frequency dependent NLO response. Inorg. Chem. Commun. 2023, 151, 110566. [Google Scholar] [CrossRef]
  23. Muhammad, S.; Bibi, J.; Alqurashy, B.A.; Alarfaji, S.S.; Kalam, A.; Chaudhry, A.R.; Al-Sehemi, A.G. Molecular Tailoring by furan and thiophene Heterocycles to optimize nonlinear optical and Photovoltaic Efficiency. J. Photochem. Photobiol. A Chem. 2025, 468, 116496. [Google Scholar] [CrossRef]
  24. Chaudhry, A.R.; Haq, B.U.; Muhammad, S.; Laref, A.; Irfan, A.; Algarni, H. Structural, electronic and optical properties of furan based materials at bulk level for photovoltaic applications: A first-principles study. Comput. Theor. Chem. 2019, 1147, 20–28. [Google Scholar] [CrossRef]
  25. Sun, H. Unraveling the structure–property relationship of novel thiophene and furan-fused cyclopentadienyl chromophores for nonlinear optical applications. J. Comput. Chem. 2024, 45, 2612–2623. [Google Scholar] [CrossRef]
  26. Mujahid, A.; Rasool, N.; Qamar, M.U.; Zubair, M.; Ahmad, F.; Altaf, A.A.; Akhtar, A.; Shah, S.A.A.; Alqahtani, F.; Alsanea, S. Arylation of halogenated thiophene carboxylate via Suzuki–Miyaura reaction: Anti-bacterial study against clinically isolated extensively drug resistant Escherichia coli sequence type 405 and computational investigation. Arab. J. Chem. 2022, 15, 103662. [Google Scholar] [CrossRef]
  27. Imran, H.M.; Rasool, N.; Kanwal, I.; Hashmi, M.A.; Altaf, A.A.; Ahmed, G.; Malik, A.; Kausar, S.; Khan, S.U.-D.; Ahmad, A. Synthesis of halogenated [1,1′-biphenyl]-4-yl benzoate and [1,1′: 3′,1″-terphenyl]-4′-yl benzoate by palladium catalyzed cascade C–C coupling and structural analysis through computational approach. J. Mol. Struct. 2020, 1222, 128839. [Google Scholar] [CrossRef]
  28. Simonetti, M.; Cannas, D.M.; Larrosa, I. Biaryl synthesis via C–H bond activation: Strategies and methods. In Advances in Organometallic Chemistry; Elsevier: Amsterdam, The Netherlands, 2017; Volume 67, pp. 299–399. [Google Scholar]
  29. Biajoli, A.F.; Schwalm, C.S.; Limberger, J.; Claudino, T.S.; Monteiro, A.L. Recent progress in the use of Pd-catalyzed CC cross-coupling reactions in the synthesis of pharmaceutical compounds. J. Braz. Chem. Soc. 2014, 25, 2186–2214. [Google Scholar]
  30. Miyaura, N.; Suzuki, A. Palladium-catalyzed cross-coupling reactions of organoboron compounds. Chem. Rev. 1995, 95, 2457–2483. [Google Scholar] [CrossRef]
  31. Ahmad, F.; Altaf, H.; Mubarik, A.; Rasool, N.; Parveen, B.; Imran, M.; Ahmad, G. Facile Synthesis of (S)-2-Aryl-N-(1-phenylethylisonicotinamides) Derivatives via SMC Reaction: Their Thermodynamic and Spectroscopic Features via DFT Approach. ChemistrySelect 2024, 9, e202403173. [Google Scholar] [CrossRef]
  32. Hafeez, J.; Sabir, A.; Rasool, N.; Hafeez, U.; Siddique, F.; Bilal, M.; Kanwal, A.; Ahmad, G.; Alqahtani, F.; Imran, I. Synthesis of N, N-Bis ([1,1′-Biphenyl]-4-ylmethyl)-4-morpholinoaniline derivatives via SMC reaction: Assessing their anti-seizure potential through electroencephalogram evaluation and molecular docking studies. Arab. J. Chem. 2024, 17, 105889. [Google Scholar] [CrossRef]
  33. Dennington, R.; Keith, T.A.; Millam, J.M. GaussView. 6.0., Version 16; Semichem Inc.: Shawnee Mission, KS, USA, 2016. [Google Scholar]
  34. Frisc, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. Gaussian 09, Revision C. 01; Gaussian Inc.: Wallingford, UK, 2010. [Google Scholar]
  35. Ahmad, G.; Rasool, N.; Mubarik, A.; Zahoor, A.F.; Hashmi, M.A.; Zubair, M.; Bilal, M.; Hussien, M.; Akhtar, M.S.; Haider, S. Facile Synthesis of 5-Aryl-N-(pyrazin-2-yl) thiophene-2-carboxamides via Suzuki Cross-Coupling Reactions, Their Electronic and Nonlinear Optical Properties through DFT Calculations. Molecules 2021, 26, 7309–7324. [Google Scholar] [CrossRef]
  36. Barone, G.; Duca, D.; Silvestri, A.; Gomez-Paloma, L.; Riccio, R.; Bifulco, G. Determination of the relative stereochemistry of flexible organic compounds by ab initio methods: Conformational analysis and Boltzmann-averaged GIAO 13C NMR chemical shifts. Chem.-Eur. J. 2002, 8, 3240–3245. [Google Scholar] [CrossRef] [PubMed]
  37. Ikram, H.M.; Rasool, N.; Hashmi, M.A.; Anjum, M.A.; Ali, K.G.; Zubair, M.; Ahmad, G.; Mahmood, T. Density functional theory-supported studies of structural and electronic properties of substituted-phenol derivatives synthesized by efficient O-or C-arylation via Chan--Lam or Suzuki cross-coupling reactions. Turk. J. Chem. 2019, 43, 1306–1321. [Google Scholar] [CrossRef]
  38. Mubarik, A.; Rasool, N.; Hashmi, M.A.; Mansha, A.; Zubair, M.; Shaik, M.R.; Sharaf, M.A.; Awwad, E.M.; Abdelgawad, A. Computational study of structural, molecular orbitals, optical and thermodynamic parameters of thiophene sulfonamide derivatives. Crystals 2021, 11, 211. [Google Scholar] [CrossRef]
  39. Zheng, B.; Huo, L. Recent advances of furan and its derivatives based semiconductor materials for organic photovoltaics. Small Methods 2021, 5, 2100493. [Google Scholar] [CrossRef] [PubMed]
  40. Migalska-Zalas, A.; El Korchi, K.; Chtouki, T. Enhanced nonlinear optical properties due to electronic delocalization in conjugated benzodifuran derivatives. Opt. Quantum Electron. 2018, 50, 389. [Google Scholar] [CrossRef]
  41. Qin, Y.; Mu, X.; Feng, X.; Xu, K.; Ma, J.; Cao, L.; Teng, B. Design, synthesis and characterization of two new nonlinear optical crystals based on chalcone containing furan ring: FTB and FTC. J. Mol. Struct. 2025, 1332, 141720. [Google Scholar] [CrossRef]
  42. Satheeshchandra, S.; Namratha, W.; Haleshappa, D.; Jayarama, A.; Shetty, N.; Pinto, R. Third order non linear optical properties of novel furan based organic crystal. Mater. Today: Proc. 2021, 35, 427–430. [Google Scholar] [CrossRef]
  43. Khan, M.U.; Hussain, S.; Asghar, M.A.; Munawar, K.S.; Khera, R.A.; Imran, M.; Ibrahim, M.M.; Hessien, M.M.; Mersal, G.A. Exploration of nonlinear optical properties for the first theoretical framework of non-fullerene DTS (FBTTh2) 2-based derivatives. ACS Omega 2022, 7, 18027–18040. [Google Scholar] [CrossRef]
  44. Kaur, P.; Singh, K. Second-order nonlinear polarizability of “Push-Pull” chromophores. A decade of progress in donor-π-acceptor materials. Chem. Rec. 2022, 22, e202200024. [Google Scholar] [CrossRef]
  45. Zaręba, J.K.; Nyk, M.; Samoć, M. Nonlinear optical properties of emerging nano-and microcrystalline materials. Adv. Opt. Mater. 2021, 9, 2100216. [Google Scholar] [CrossRef]
  46. Bullo, S.; Jawaria, R.; Faiz, I.; Shafiq, I.; Khalid, M.; Asghar, M.A.; Baby, R.; Orfali, R.; Perveen, S. Efficient synthesis, spectroscopic characterization, and nonlinear optical properties of novel salicylaldehyde-based thiosemicarbazones: Experimental and theoretical studies. ACS Omega 2023, 8, 13982–13992. [Google Scholar] [CrossRef]
  47. Palatnikov, M.; Kadetova, A.; Smirnov, M.; Sidorova, O.; Vorobev, D. Nonlinear optical properties of lithium niobate crystals doped with alkaline earth and rare earth elements. Opt. Mater. 2022, 131, 112631. [Google Scholar] [CrossRef]
  48. Raza Shah, A.; Rasool, N.; Bılal, M.; Mubarık, A.; Alı Hashmı, M.; Nadeem Akhtar, M.; Imran, M.; Ahmad, G.; Siddiqa, A.; Adnan Alı Shah, S. Efficient Synthesis of 4-Bromo-N-(1-phenylethyl) benzamide, Arylation by Pd (0) Catalyst, Characterization and DFT Study. ChemistrySelect 2022, 7, e202200861. [Google Scholar] [CrossRef]
  49. Srivastava, A.K.; Pandey, A.K.; Jain, S.; Misra, N. FT-IR spectroscopy, intra-molecular C− H⋯ O interactions, HOMO, LUMO, MESP analysis and biological activity of two natural products, triclisine and rufescine: DFT and QTAIM approaches. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 136, 682–689. [Google Scholar] [CrossRef]
  50. Bayoumy, A.M.; Ibrahim, M.; Omar, A. Mapping molecular electrostatic potential (MESP) for fulleropyrrolidine and its derivatives. Opt. Quantum Electron. 2020, 52, 346. [Google Scholar] [CrossRef]
  51. Kumar, K.; Vishwkarma, A.K.; Pratap, R.; Garai, S.; Pathak, A.; Sharma, R.; Bhattacharya, S. Synthesis, crystal structures, and theoretical studies on bio-pertinent disulfides derived from methyl thiosalicylate and furan-2-thiocarboxylic acid and their interaction with native human lysozymes. Chem. Pap. 2024, 78, 3721–3733. [Google Scholar] [CrossRef]
  52. Fatima, A.; Khanum, G.; Verma, I.; Butcher, R.J.; Siddiqui, N.; Srivastava, S.K.; Javed, S. Synthesis, characterization, crystal structure, Hirshfeld surface, electronic excitation, molecular docking, and DFT studies on 2-amino thiophene derivative. Polycycl. Aromat. Compd. 2023, 43, 1644–1675. [Google Scholar] [CrossRef]
  53. Domingo, L.R.; Ríos-Gutiérrez, M.; Pérez, P. Applications of the conceptual density functional theory indices to organic chemistry reactivity. Molecules 2016, 21, 748. [Google Scholar] [CrossRef]
  54. Chattaraj, P.; Poddar, A. Molecular reactivity in the ground and excited electronic states through density-dependent local and global reactivity parameters. J. Phys. Chem. A 1999, 103, 8691–8699. [Google Scholar] [CrossRef]
  55. Pearson, R.G. Chemical hardness and density functional theory. J. Chem. Sci. 2005, 117, 369–377. [Google Scholar] [CrossRef]
  56. Vektariene, A.; Vektaris, G.; Svoboda, J. A theoretical approach to the nucleophilic behavior of benzofused thieno [3, 2-b] furans using DFT and HF based reactivity descriptors. Ark. Online J. Org. Chem. 2009, vii, 311–329. [Google Scholar] [CrossRef]
  57. Guthrie, G., Jr.; Scott, D.; Hubbard, W.; Katz, C.; McCullough, J.; Gross, M.; Williamson, K.; Waddington, G. Thermodynamic properties of furan. J. Am. Chem. Soc. 1952, 74, 4662–4669. [Google Scholar] [CrossRef]
  58. Pérez, P.; Domingo, L.R.; Aizman, A.; Contreras, R. The electrophilicity index in organic chemistry. In Theoretical and Computational Chemistry; Elsevier: Amsterdam, The Netherlands, 2007; Volume 19, pp. 139–201. [Google Scholar]
  59. Parr, R.G.; Szentpály, L.v.; Liu, S. Electrophilicity index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  60. Hansen-Felby, M.; Henriksen, M.L.; Pedersen, S.U.; Daasbjerg, K. Postfunctionalization of self-immolative poly (dithiothreitol) using Steglich esterification. Macromolecules 2022, 55, 5788–5794. [Google Scholar] [CrossRef]
  61. Afzal, U.; Mahmood, A.; Zubair, M.; Rasool, N.; Kanwal, A.; Sohail, M.; Ahmad, G. Synthesis, hydrolysis, and COX-2/15-LOX inhibitory evaluation of 4-Acetamidophenyl 4-Bromobenzoates. Med. Chem. Res. 2025, 34, 982–995. [Google Scholar] [CrossRef]
  62. Johnson, B.A. Using NMRView to Visualize and Analyze the NMR Spectra of Macromolecules. In Protein NMR Techniques; Humana Press: Totowa, NJ, USA, 2004; pp. 313–352. [Google Scholar]
Scheme 1. Synthesis of 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3).
Scheme 1. Synthesis of 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3).
Catalysts 15 00713 sch001
Scheme 2. Arylation of 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylates (5a5e).
Scheme 2. Arylation of 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylates (5a5e).
Catalysts 15 00713 sch002
Figure 1. 3D Optimized structures of synthesized derivatives (5a5e) by the PBE0-D3BJ/def2-TZVP/SMD1,4-dioxane theory level.
Figure 1. 3D Optimized structures of synthesized derivatives (5a5e) by the PBE0-D3BJ/def2-TZVP/SMD1,4-dioxane theory level.
Catalysts 15 00713 g001
Figure 2. Molecular Orbitals (HOMO-LUMO) of (5a5e) at PBE0-D3BJ/def2-TZVP/SMD1,4-dioxane level of theory.
Figure 2. Molecular Orbitals (HOMO-LUMO) of (5a5e) at PBE0-D3BJ/def2-TZVP/SMD1,4-dioxane level of theory.
Catalysts 15 00713 g002
Figure 3. Molecular electrostatic potentials maps of (5a5f).
Figure 3. Molecular electrostatic potentials maps of (5a5f).
Catalysts 15 00713 g003
Table 1. Optimization of % age yield of 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3).
Table 1. Optimization of % age yield of 4-(methoxycarbonyl)phenyl 5-bromofuran-2-carboxylate (3).
EntrySolventCatalystTemp.Time% Age Yield
1.DCMDCC/DMAP0 °C2 h21
2.THFDCC/DMAP0 °C2 h41
3.CH3CNDCC/DMAP0 °C2 h88
Table 2. Energies of HOMO, LUMO, their energy gap, and hyperpolarizability β values.
Table 2. Energies of HOMO, LUMO, their energy gap, and hyperpolarizability β values.
CompoundsEHOMOELUMOELUMO-EHOMOHyperpolarizability β
5a−6.29−1.744.553897.42
5b−6.42−1.904.524396.08
5c−6.54−2.024.524080.83
5d−6.33−1.734.604242.53
5e−6.52−1.924.603859.80
Table 3. Reactivity Parameters (I, A, η, μ, ω, σ) of substituted arylated furan carboxylates 5a5e.
Table 3. Reactivity Parameters (I, A, η, μ, ω, σ) of substituted arylated furan carboxylates 5a5e.
CompoundsI (eV)A (eV)η (eV)μ (eV)ω (eV)σ (eV−1)
5a6.291.742.27–4.023.540.44
5b6.421.902.25–4.163.840.44
5c6.542.022.26–4.284.050.44
5d6.331.732.30–4.033.540.43
5e6.521.922.30–4.223.870.43
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fakhar U. Zaman, M.; Mubarik, A.; Kanwal, A.; Rasool, N.; Ahmad, M.; Sohail, M.; Malik, A.; Al-Hussain, S.A.; Zaki, M.E.A. Facile Synthesis of 4-(Methoxycarbonyl)phenyl 5-Arylfuran-2-Carboxylates via Readily Available Pd Catalyst–Their Thermodynamic, Spectroscopic Features and Nonlinear Optical Behavior. Catalysts 2025, 15, 713. https://doi.org/10.3390/catal15080713

AMA Style

Fakhar U. Zaman M, Mubarik A, Kanwal A, Rasool N, Ahmad M, Sohail M, Malik A, Al-Hussain SA, Zaki MEA. Facile Synthesis of 4-(Methoxycarbonyl)phenyl 5-Arylfuran-2-Carboxylates via Readily Available Pd Catalyst–Their Thermodynamic, Spectroscopic Features and Nonlinear Optical Behavior. Catalysts. 2025; 15(8):713. https://doi.org/10.3390/catal15080713

Chicago/Turabian Style

Fakhar U. Zaman, Muhammad, Adeel Mubarik, Aqsa Kanwal, Nasir Rasool, Matloob Ahmad, Maria Sohail, Ayesha Malik, Sami A. Al-Hussain, and Magdi E. A. Zaki. 2025. "Facile Synthesis of 4-(Methoxycarbonyl)phenyl 5-Arylfuran-2-Carboxylates via Readily Available Pd Catalyst–Their Thermodynamic, Spectroscopic Features and Nonlinear Optical Behavior" Catalysts 15, no. 8: 713. https://doi.org/10.3390/catal15080713

APA Style

Fakhar U. Zaman, M., Mubarik, A., Kanwal, A., Rasool, N., Ahmad, M., Sohail, M., Malik, A., Al-Hussain, S. A., & Zaki, M. E. A. (2025). Facile Synthesis of 4-(Methoxycarbonyl)phenyl 5-Arylfuran-2-Carboxylates via Readily Available Pd Catalyst–Their Thermodynamic, Spectroscopic Features and Nonlinear Optical Behavior. Catalysts, 15(8), 713. https://doi.org/10.3390/catal15080713

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop