Next Article in Journal
i-Propylammonium Lead Chloride Based Perovskite Photocatalysts for Depolymerization of Lignin Under UV Light
Next Article in Special Issue
Design of Curcumin and Flavonoid Derivatives with Acetylcholinesterase and Beta-Secretase Inhibitory Activities Using in Silico Approaches
Previous Article in Journal
Proton Conductive Zr-Phosphonate UPG-1—Aminoacid Insertion as Proton Carrier Stabilizer
Previous Article in Special Issue
Antioxidant and Anticancer Activity of Novel Derivatives of 3-[(4-Methoxyphenyl)amino]propanehydrazide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Selective Arylation of 2-Bromo-4-chlorophenyl-2-bromobutanoate via a Pd-Catalyzed Suzuki Cross-Coupling Reaction and Its Electronic and Non-Linear Optical (NLO) Properties via DFT Studies

1
Department of Chemistry, Government College University, Faisalabad 38000, Punjab, Pakistan
2
College of Chemistry and Molecular Engineering, Zhengzhou University, Kexue road No. 100, Zhengzhou 450001, China
3
Department of Chemistry, COMSATS University Islamabad, Abbottabad Campus, University Road, Tobe Camp, Abbottabad 22060, Pakistan
4
Faculty of Pharmacy, Universiti Teknologi MARA Cawangan Selangor Kampus Puncak Alam, Bandar Puncak Alam 42300, Malaysia
5
Atta-ur-Rahman Institute for Natural Products Discovery (AuRIns), Universiti Teknologi MARA Cawangan Selangor Kampus Puncak Alam, Bandar Puncak Alam 42300, Malaysia
6
Department of Biomedical Science, Faculty of Medicine and Health Sciences, Universiti Putra Malaysia, Serdang 43400, Malaysia
7
Halal Institute Research Institute, Universiti Putra Malaysia, Serdang 43400, Malaysia
8
Faculty of Industrial Sciences & Technology, Universiti Malaysia Pahang, Lebuhraya Tun Razak, Gambang Kuantan 26300, Malaysia
9
Bio-Aromatic Research Center of Excellence, Faculty of Industrial Sciences & Technology, Universiti Malaysia Pahang, Lebuhraya Tun Razak, Gambang Kuantan 26300, Malaysia
*
Authors to whom correspondence should be addressed.
Molecules 2020, 25(15), 3521; https://doi.org/10.3390/molecules25153521
Submission received: 29 May 2020 / Revised: 14 July 2020 / Accepted: 15 July 2020 / Published: 31 July 2020
(This article belongs to the Special Issue 25th Anniversary of Molecules—Recent Advances in Organic Synthesis)

Abstract

:
In the present study, 2-bromo-4-chlorophenyl-2-bromobutanoate (3) was synthesized via the reaction of 2-bromo-4-chlorophenol with 2-bromobutanoyl bromide in the presence of pyridine. A variety of 2-bromo-4-chlorophenyl-2-bromobutanoate derivatives (5af) were synthesized with moderate to good yields via a Pd-catalyzed Suzuki cross-coupling reaction. To find out the reactivity and electronic properties of the compounds, Frontier molecular orbital analysis, non-linear optical properties, and molecular electrostatic potential studies were performed.

1. Introduction

Carboxylic esters are an important class of organic compounds, having various biological activities such as antifungal [1], antioxidant [2], anti-inflammatory [3], antiproliferative [4], and anti-bacterial [5], and they have been widely used in the biofuel [6], plasticizer [7], food, and cosmetic industries [8]. In natural product chemistry, esters are well known common intermediates because of their accessibility for easy interconversion and their stability. Esters are generally synthesized by the condensation of carboxylic acids with alcohols. Moreover, activation of the carboxyl group followed by reaction with suitable alcohol is one of the most familiar methods for ester formation [9]. Similarly, biphenyl compounds are used as central building blocks for basic liquid crystal [10] and fluorescent layers in organic light-emitting diodes (OLEDs) [11] and are also important pharmaceutically as anti-hypertensive [12], anti-inflammatory [13], potential anti-cholinesterase [14], anti-diabetic [15], anti-tumor [16], anti-cancer [17], and anti-leukemia agents [18], and as potential therapeutics for cardiovascular disease [19] and osteoporosis [20].
For the synthesis of biphenyl compounds, the Suzuki–Miyaura reaction is one of the best methods. In organic synthesis, regioselective functionalization plays a progressively important role [21]. In such types of reactions, the reaction takes place at the carbon atom, which is more electron-deficient, while the other active site remains unattached. This concept has been practiced in region-selective Pd-catalyzed cross-coupling reactions based on different rates of oxidative additions of Pd (0) species to carbon halide bonds of different substrates. Therefore, we focused on the synthesis of 2-bromo-4-chlorophenyl-2-bromobutanoate followed by regioselective arylation of 2-bromo-4-chlorophenyl-2-bromobutanoate via a palladium-catalyzed Suzuki–Miyaura cross-coupling reaction to obtain the 2-bromo-4-chlorophenyl-2-bromobutanoate derivatives. Frontier molecular orbital (FMO), non-linear optic properties (NLO) and molecular electrostatic potential (MEP) studies were performed to find out the most reactive site of the compound.

2. Results and Discussion

2.1. Chemistry

In the present study, 2-bromo-4-chlorophenyl-2-bromobutyrate (3) was synthesized by the reaction of 2-bromo-4-chlorophenol (1) with 2-bromobutryl bromide (2) in the presence of pyridine and acetonitrile, and good yield of 2-bromo-4-chlorophenyl-2-bromobutyrate (3) was observed (Scheme 1).
In addition, the Suzuki–Miyaura cross-coupling reaction of 2-bromo-4-chlorophenyl-2-bromo butyrate (3) with different phenyl boronic acids was accomplished, which ultimately led to the formation of the corresponding 2-bromo-4-chlorophenyl-2-bromobutyrates (5af) in moderate to good yields (64–81%) (Scheme 2, Table 1). When 1.1 eq of phenyl boronic acid was used, it was noted that the bromo group of the phenyl group was replaced, while in the presence of Pd (PPh3)4, the other bromo group present at the alpha position could not be replaced. The chloro group at the phenyl ring was not replaced due to the C-Cl bond strength that obstructs the reactivity of aryl chlorides, thus, they are resistant to the oxidative addition to Pd(0) because activation of aryl chlorides is much more difficult than that of aryl bromides. The oxidative addition step is a selectivity-determining step in Suzuki cross-coupling, and it is broadly controlled by the bond dissociation energies of the carbon–halogen bond that predictably differs as a function of the aryl halides: Ar-I > Ar-Br > Ar-Cl > Ar-F [22]. It was noted that phenyl boronic acids having electron-donating substituents gave high yields as compared to electron-withdrawing substituted phenyl boronic acids.

2.2. Frontier Molecular Orbital Analysis

Frontier molecular orbital (FMO) analysis is well known to explain the reactivity and electronic properties of the molecules [23]. The energies of HOMOs (EHOMOs) and LUMOs (ELUMOs) provide information about the electronic transition and the HOMOs–LUMOs gap (GH-L) explains the reactivity and kinetic stability of the compounds. If the GH-L is low, the compound will be less stable and more reactive, but if the GH-L is large, the compound will be highly stable and less reactive [24]. The energies of HOMOs (EHOMOs), energies of LOMOs (ELUMOs), and the highest occupied molecular orbital to lowest unoccupied molecular orbital gap (GH-L) are summarized in Table 2.
Among all compounds, compound 5c has the lowest GH-L (4.27 eV) where the EHOMOs and ELUMOs were −6.06 and −1.79 eV, respectively. Based on the lower GH-L, 5c is considered as kinetically less stable and moderately reactive. Compound 5f has GH-L of 5.19 eV, which showed it is the most stable and least reactive of the whole series (5af). The GH-L of the other compounds were between 4.57 and 5.14 eV, which represent their moderate reactivity and sufficient stability. Although the GH-L of 5c is low, compared to other compounds, it is large.
The iso-density distributions in the FMOs of all compounds were also shown in Figure 1. In all the compounds (except compounds 5c and 5e), the iso-density of HOMOs was distributed on the phenyl rings and halogen (chlorine and fluorine) groups that were present on the phenyl rings. However, this iso-density of HOMOs is shifted from the chlorine-containing phenyl ring towards the MeS- and MeO-groups containing phenyl rings in compounds 5c and 5e, respectively. This specific distribution explains the lower iso-density of 5c and 5e compared to that of the other compounds. The shift of iso-density is more pronounced for 5c in comparison to 5e. The iso-density of LUMOs is mostly distributed on the bromo-ester group attached to the phenyl ring in all compounds (5af).

2.3. Reactivity Descriptor Parameters

For further elucidation of the reactivity of these compounds (5af), some other reactivity descriptor parameters were also studied. The results of the analyses are given in Table 3. These parameters include ionization potential (I), electrophilicity index (Ꙍ), nucleophilicity (N), electron affinity (EA), electronic chemical potential (μ), and chemical hardness (Ƞ). According to Koopman’s theorem, the negative of HOMOs and LUMOs values correspond to the ionization potential (I) and electron affinity (EA), respectively [25].

2.3.1. Ionization Potential, Electron Affinity, and Chemical Hardness (Ƞ)

The Chemical hardness is mathematically expressed as:
Chemical hardness (Ƞ) = (EHOMO − ELUMO)/2
The values obtained for all compounds from different reactivity descriptor are comparable and support each other, which strengthens our findings. The ionization potential of all compounds was in the range of 6.06 eV to 7.06 eV. The most reactive and unstable compound in this series was 5c, and its reactivity was also supported from its lower ionization potential (I = 6.06 eV), electron affinity (EA = 1.79 eV), and chemical hardness (Ƞ = 2.13 eV) values. The compound 5f was the most stable compound, which is also confirmed from its high I (7.06 eV) and Ƞ (2.60 eV) values. The chemical hardness of the other compounds was between 2.13 and 2.57 eV.

2.3.2. Electronic Chemical Potential (μ)

The electronic chemical potential gives us an idea about the charge transfer inside compounds at their ground state, and it is mathematically expressed as:
Electronic chemical potential (μ) = (EHOMO + ELUMO)/2
The electronic chemical potential (μ) of all compounds (5af) was in the range of −3.92 eV to −4.52 eV. The highest μ (−4.52 eV) value was observed for 5d and the lowest μ value (−3.92 eV) was calculated for 5c.

2.3.3. Electrophilicity Index (Ꙍ)

Parr et al. provided the concept of the electrophilicity index (Ꙍ) [26], which is mathematically expressed as:
Electrophilicity index (Ꙍ) = μ2/2Ƞ
According to this concept, the Ꙍ is the stabilization of a compound based on energy when an additional charge is transferred from the surroundings or outside environment. The Ꙍ ranged from 3.52 eV to 4.02 eV for compounds 5af. The compound 5e had the lowest Ꙍ value (3.52 eV) in the series of compounds, which indicates that this compound is less stable for an incoming charge. The reason is the electronic acceptor and donor groups present in the compound are connected through extended conjugation. The incoming charge causes further delocalization of electronic density in the compound and enhances its reactivity.

2.3.4. Nucleophilicity (N)

Nucleophilicity is derived from HOMOs and LUMOs values by using the following mathematical equation:
Nucleophilicity (N) = EHOMO − ELUMO
Nucleophilicity is an important parameter for understanding the reactivity of compounds. All the compounds had nucleophilicity values in the range of 4.27 eV to 5.19 eV. The compound 5c had the lowest N value (4.27 eV), which represents the soft nucleophilic nature of this compound. Whereas compound 3′5-Dichloro-[1,1′-biphenyl]-2-yl 2-bromobutanoate (5f) has the highest N value (5.19 eV), which shows the strong nucleophilic nature of this compound.

2.4. Molecular Electrostatic Potential (MEP)

Molecular electrostatic potential (MEP) is used to evaluate and predict the reactive sites of a molecule [27]. MEP is generated on the surfaces of all optimized compounds to explain the nucleophilic and electrophilic sites. The potentials on the surfaces of complexes are characterized by various colors and are described in increasing order of red < orange < yellow < green < blue. The green colors show the neutral region while red and blue colors are used for the nucleophilic and electrophilic sites of the complex, respectively. The molecular electrostatic surfaces of all compounds are given in Figure 2 and the values are given in Table 4.
Figure 2 indicates that the ester group attached to the phenyl ring is an active site for an incoming electrophile. Whereas, hydrogen attached to the phenyl rings have blue electronic densities that represent the strong nucleophilic site. Yellow densities reside on the S atom of the MeS, O of the MeO, Cl, and F atoms, illustrating the moderately electrophilic sites of these functional groups. The highest electronic density distribution of −4.75 × 10−2 to 4.75 × 10−2 was observed for compound 5d. In contrast, the lowest electronic density distribution of −5.22 × 10−2 to 5.22 × 10−2 was observed for compound 5e.

2.5. Non-linear Optical (NLO) Properties

For the last decade, scientific society has been pursuing the pathways to synthesized non-linear optical (NLO) materials with large second-order nonlinear optical properties. Non-linear optical materials have various applications in laser devices, optical communication, optical data storage, optical limiting, optical computing, and medical imaging [28,29]. The NLO response of these materials is measured by the first hyperpolarizability (βo). In organic molecules, the movement of electrons from donor to acceptor group is responsible for the enhancement of βo, which ultimately results in a strong NLO response.
In the synthesized compounds (5af), the para chloro-phenyl ring (p-CPR) acts as electron-withdrawing group and influences different substitutions on the NLO response were studied at the ortho position of phenyl ring. The ester group is not involved due to the resonance between the two oxygen atoms and the electrons are unable to delocalize inside the phenyl ring. When the electron donating groups are attached to the p-CPR, the electron push full mechanism is the operative one and a large βo value is obtained. However, if the electron-withdrawing groups are attached to the p-CPR, then the electron push full mechanism is not operating and the small βo value is obtained in that case. The compound 5c has a strong electron donating group (MeS-) at the para position of the phenyl ring and we noticed a large βo value (1373.76 au) for it according to the push full mechanism. The lower βo value (218.51 au) was observed for compound 5f where an electron-withdrawing chloro group (Cl) is present on the para position of the phenyl ring. On both sides of the phenyl rings, similar chloro-groups are attached which make these rings strongly electron-withdrawing, and the net effect is zero. The moderate electron donating methoxy group at the para position of the phenyl ring (in 5e) had a reasonable NLO response compared to compound 5b where the acetate group has an internal resonance, which has a negative impact on the NLO response. The βo values of 5e and 5b compounds were 954.47 au and 284.34 au, respectively. Compounds 5a and 5d had almost comparable βo values (in the range of 510 au −550 au). The reason may be that the presence of halogen groups on the meta and para positions of the ring nullified the effects of each other and the phenyl ring acted as moderately electron-donating. Based on these results, we suggest that compound 5c shows a better NLO response and can be used as an active NLO material in the future.
The hyperpolarizability of these compounds (5af) was also supported by the polarizability values (Table 5). Polarizability is defined as the electronic density distribution in a system. In compounds where electron-accepting and electron-donating groups are present at the opposite termini of the phenyl rings, they must have positive and negative centers and their αo is high. In contrast, in those compounds where both termini have similar functional groups, there is an equal distribution of electronic density and αo is low. For example, compound 5c had a high αo value (264 au) whereas 5f had a low αo value (235 au). The rest of the compounds had moderate values of αo in the range of 236 au to 252 au.

Non-linear Refractive Index (n2)

The quadratic non-linear refractive index (n2) was calculated by using degenerated four waves mixing gamma values (γDFWM (ω) = (−ω; ω, −ω, ω) as proposed by Tarazker et al. [30] and it is mathematically expressed as follows:
γDFWM (−ω; ω, −ω, ω) = (1/3) γ (−2ω; ω, ω, 0) + (1/3) γ (−ω; ω, 0, 0) + (1/3) γ (0; 0, 0, 0)
where γ (0; 0, 0, 0), γ (−ω; ω, 0, 0), and γ (−2ω; ω, ω, 0) represent static, dc-Kerr, and EFSHG second hyperpolarizability coefficients, respectively. The γDFWM is further used to measure the quadratic non-linear refractive index (n2) using Equation 6.
n2 (cm2/W) = 8.28 × 10−23γDFWM (au)
The n2 values of all compounds (5af) at various frequencies (0, 532 and 1064 nm) are given in Table 6 and graphically represented in Figure 2. In the spectral curves of n2, the ESHG effect and dc-Kerr effect are similar and demonstrate the dominant contribution of both these effects in enhancing n2 (Figure 3). However, the effect of the ESHG effect is more pronounced as compared to that of the dc-Kerr effect. The n2 values for compounds (5af) were in the range 8.01 × 10−18 au to 9.46 × 10−16 au at ω = 532nm. These compounds have n2 values in the range of 4.43 × 10−18 au to 8.41 × 10−18 au at ω = 1064 nm. Among all compounds, the distinct effect was noticed at a low wavelength (532 nm) where the highest value (9.46 × 10−16 au) was noticed for compound 5c. These n2 values proved the influence of various wavelengths on the amplitude of hyperpolarizability and gamma values of these compounds. Besides the various wavelengths, the structure of the compounds also plays a substantial role in the fluctuation of the NLO response. The structure of compound 5c has electron donating and electron withdrawing groups on both terminals of the main skeleton and an excellent push–pull mechanism occurs due to extended conjugation, and its NLO response is outstanding. This result is also supported by the nonlinear refractive index results.

3. Materials and Methods

3.1. Procedure for Synthesis of 2-Bromo-4-chlorophenyl-2-bromobutyrate (3)

First, 2-Bromo-4-chlorophenol (1 g, 4.820 mmol), MeCN (30 mL), pyridine (0.775 mL, 9.63 mmol), and 2-bromobutryl bromide (0.64 mL, 5.3 mmol) were added to a 250 mL round bottom flask with a magnetic stirrer. The flask’s mouth was closed with a Teflon septum. The flask was put on the magnetic hot plate along with an isotherm to maintain the temperature of the reaction mixture at 0 °C. The progress of the reaction was estimated by thin layer chromatography analysis. After 1 h, the reaction was completed. The separation and purification of the desired product was carried out using the column chromatography technique. The NMR spectroscopic technique was used to find out the structure of the desired compound [31].
2-Bromo-4-chlorophenyl 2-bromobutanoate (3), yellow liquid, 1H-NMR: (600 MHz, CDCl3) δ 7.9 (d, J = 1.5 Hz, 1H), 7.50 (dd, J1 = 7.5 Hz, J2 = 1.5 Hz, 1H), 7.29 (d, J = 7.5 Hz, 1H), 4.29 (t, J = 7.1 Hz, 1H), 2.05–2,10 (m, 2H, Ar), 1.1 (t, J = 8.0 Hz, 3H). 13C-NMR (150 MHz, CDCl3) δ 10.1 (CH3), 24.1 (CH2), 42.7 (CH), 115.2 (C), 117.6 (CH), 121.3 (CH), 129.4 (CH), 139.5 (C), 145.5 (C), 170.2 (C). HR-ESI-MS: m/z Calcd for C10H9Br2ClO, [M]+ 353.8909; found 353.8945.

3.2. General Procedure for Synthesis of (5af)

A reflux condenser flushed with argon was attached to a 100 mL schlenk flask with a magnetic stirrer. Then, 2-bromo-4-chlorophenyl-2-bromobutyrate (1 equiv.) and tetrakis(triphenyl phosphine)palladium(0) (5 equiv.) were added to the schlenk flask and dissolved into 8 mL of 1,4-dioxane and the flask’s mouth was closed with a Teflon septum. The reaction mixture was stirred for about 25–30 min under an inert argon atmosphere. Then, K3PO4 (2 equiv.) and aryl boronic acids and 2 mL distilled water were added and the reflux reaction continued under an inert argon atmosphere for 18–25 h at 80–90 °C. During the reaction, the estimation of the reaction was determined by a thin layer chromatography technique using TLC cards. When the reaction was completed, the reaction mixture was filtered using Whatman filter paper and the solvent was evaporated by a rotary evaporator. The separation and purification of the impure products were carried out using the column chromatography technique. 1H-NMR was used to find out the structure of the newly synthesized compounds [32].
3′5-Dichloro-4′-fluoro-[1,1′-biphenyl]-2-yl 2-bromobutanoate (5a), white crystals, m.p. 247–248 °C. 1H-NMR (600 MHz, CDCl3) 8.04 (d, J = 1.8 Hz, 1H), 8.00 (d, J = 1.5 Hz 1H), 7.55 (dd, J1 = 1.6 Hz, J2 = 7.7 Hz, 1H), 7.41 (dd, J1 = 7.7 Hz, J2 = 8.0 Hz 1H), 7.38 (d, J = 8.1 Hz, 1H), 7.35 (m, 1H), 4.2 (t, J = 7.0 Hz, 1H), 2.11–2.17 (m, 2H, Ar), 1.11 (t, J = 7.5 Hz, 3H). 13C-NMR (150 MHz, CDCl3) δ 10.2 (CH3), 25.4 (CH2), 42.7 (CH), 117.6 (CH), 122.4 (C), 125.6 (CH), 130.5 (CH), 131.6 (CH), 132.5 (CH), 133.2 (CH), 132.9 (C), 133.4 (C), 133.9 (C), 148.1 (C), 162.3 (C), 170.1 (C). HR-ESI-MS: m/z Calcd for C16H12BrCl2FO2, [M]+ 406.0722; found 406.0742.
3′-Acetoxy-5-chloro-[1,1′-biphenyl]-2-yl 2-bromobutanoate (5b), yellow crystals, m.p: 268–269 °C. 1H-NMR (600 MHz, CDCl3) δ 8.00 (d, J = 2.0 Hz, 1H), 7.50 (dd, J1 = 7.9 Hz, J2 = 1.8 Hz, 1H), 7.46 (d, J = 1.5 Hz, 1H), 7.42–7.34 (m, 2H), 7.35 (dd, J1 = 7.7 Hz, J2 = 1.5 Hz, 1H), 7.28 (d, J = 7.5 Hz, 1H), 4.20 (t, J = 7.0 Hz, 1H), 2.35 (s, 3H), 2.10–2.15 (m, 2H, Ar), 1.10 (t, J = 7.9 Hz, 3H). 13C-NMR (150 MHz, CDCl3) δ 10.1 (CH3), 19.4 (CH3), 25.6 (CH2), 45.4 (CH), 120.4 (CH), 121.1 (CH), 122.3 (CH), 128.7 (CH), 129.1 (CH), 131.2 (C), 132.7 (CH), 134.8 (CH), 136.3 (C), 137.5 (C), 145.1 (C), 148.1 (C), 170.1 (C), 171.2 (C). HR-ESI-MS: m/z Calcd for C18H16BrClO4, [M]+ 411.6808; found 411.6892.
5-Chloro-4′-(methylthio)-[1,1′-biphenyl]-2-yl 2-bromobutanoate (5c), brown crystals, m.p: 250–251 °C. 1H-NMR (600 MHz, CDCl3) δ 8.00 (d, J = 1.9 Hz, 1H), 7.65 (d, J = 8.2 Hz, 2H), 7.45 (d, J = 7.9 Hz, 2H), 7.35 (dd, J1= 7.5 Hz, J2 = 2.0 Hz, 1H), 7.25 (d, J = 8.4 Hz, 1H), 4.19 (t, J = 7.0 Hz, 1H), 2.53 (s, 3H), 2.05–2.11 (m, 2H, Ar), 1.10 (t, J = 7.0 Hz, 3H). 13C-NMR (150 MHz, CDCl3) δ 10.1 (CH3), 19.4 (CH3), 25.6 (CH2), 45.4 (CH), 121.1 (CH), 124.3 (CH), 125.4 (2CH), 128.1 (C), 130.2 (CH), 131.1 (C), 132.7 (C), 134.8 (CH), 135.4 (CH), 136.1 (C), 146.1 (C), 170.1 (C). HR-ESI-MS: m/z Calcd for C17H16BrClO2S, [M]+ 399.7145; found 399.7119.
3′,4′,5-Trichloro-[1,1′-biphenyl]-2-yl 2-bromobutanoate (5d), white crystals, m.p: 276–277 °C. 1H-NMR (600 MHz, CDCl3) δ 7.96 (d, J = 2.3 Hz, 1H), 7.83 (d, J = 1.9 Hz, 1H), 7.51–7.43 (m, 2H), 7.40 (dd, J1 = 8.2 Hz, J2 = 2.0 Hz, 1H), 7.33 (d, J = 1.5 Hz, 1H), 4.20 (t, J = 7.0 Hz, 1H), 2.15–2.20 (m, 2H, Ar), 1.10 (t, J = 8.0 Hz, 3H).13C-NMR (150 MHz, CDCl3) δ 10.1 (CH3), 25.6 (CH2), 45.4 (CH), 121.1(CH), 126.5 (CH), 128.1 (CH), 130.2 (CH), 131.2 (CH), 132.7 (C), 135.8 (CH), 136.2 (C), 137.2 (C), 138.5 (C), 141.5 (C), 148.1 (C), 170.1 (C). HR-ESI-MS: m/z Calcd for C16H12BrCl3O2, [M]+ 422.5233; found 422.5289.
5-Chloro-4′-methoxy-[1,1′-biphenyl]-2-yl 2-bromobutanoate (5e), light brown crystal, m.p: 237–238 °C. 1H-NMR (600 MHz, CDCl3) δ 7.98 (d, J = 2.0 Hz, 1H), 7.65 (d, J = 2.5 Hz, 2H) 7.49–7.35 (m, 2H), 7.10 (dd, J1 = 8.2 Hz, J2 = 1.5 Hz, 2H), 4.22 (t, J = 6.9 Hz, 1H), 3.93 (s, 3H), 2.08–2.15 (m, 2H, Ar), 1.10 (t, J = 8.0 Hz, 3H). 13C-NMR (150 MHz, CDCl3) δ 10.1 (CH3), 25.6 (CH2), 45.4 (CH), 60.2 (CH3), 116.1 (CH), 116.5 (CH), 124.5 (CH), 125.3 (C), 128.4 (CH), 130.3 (CH), 131.4 (C), 132.2 (CH), 13.3 (CH), 132.7 (C), 135.8 (C), 151.2 (C), 170.1 (C). HR-ESI-MS: m/z Calcd for C17H16BrClO3, [M]+ 383.4515; found 383.4565.
3′,5-Dichloro-[1,1′-biphenyl]-2-yl 2-bromobutanoate (5f), white crystal, m.p: 232–233 °C. 1H-NMR (600 MHz, CDCl3) δ 8.01 (d, J = 2.5 Hz, 2H), 7.52 (dd, J1 = 7.5 Hz, J2 = 1.9 Hz, 1H), 7.55 (t, J = 2.0 Hz, 1H), 7.42–7.41 (m, 2H), 7.34 (d, J = 7.0 Hz, 1H), 4.22 (t, J = 7.0 Hz, 1H), 2.08–2.15 (m, 2H, Ar), 1.10 (t, J = 8.0 Hz 3H). 13C-NMR (150 MHz, CDCl3) δ 10.1 (CH3), 25.6 (CH2), 45.4 (CH), 121.1 (2CH), 122.4 (CH), 121.9 (CH), 125.2 (CH), 128.7 (2CH), 130.2 (C), 132.7 (C), 133.2 (C), 134.7 (C), 151.3 (C), 171.2 (C). HR-ESI-MS: m/z Calcd for C16H13BrCl2O2, [M]+ 388.0861; found 388.0839.

3.3. Computational Details

All calculations were performed in Gaussian 09 software [33]. The optimized structures were drawn using Gauss View 5.0 software [34]. The structures of synthesized compounds (5af) were optimized with the hybrid functional (B3LYP) of the density functional theory (DFT). B3LYP is a widely used functional for the optimization of organic molecules because the minimization of errors in the geometrical parameters is lower for B3LYP compared to that of other density functionals. The split valence basis set developed by Pople’6-311G (d,p) was implemented for molecular orbital descriptions. For further confirmation of these minimum energy structures obtained from optimization (Figure 1), frequency analysis and frontier molecular orbital analysis were done at the same level of theory. During frequency simulations, the harmonic force constant was calculated, which validated the absence of imaginary frequency and justified these as true minima structures. FMOs, reactivity descriptor parameters (ionization potential, electron affinity, chemical hardness, electrophilicity index, electronic chemical potential, and nucleophilicity), and molecular electrostatic potential (MEP) were calculated with the B3LYP/6-311G (d,p) method. NLO properties including polarizabilities, first hyperpolarizability, and gamma values were calculated with CAM-B3LYP and LC-BLYP with 6-311+G (d,p) levels. The best NLO results were obtained by the CAM-B3LYP/6-31+G (d,p) level, so we discussed our results obtained on this level of theory. The nonlinear refractive index was also calculated a with the CAM-B3LYP/6-31+G (d,p) method.

4. Conclusions

In conclusion, a series of 2-bromo-4-chlorophenyl-2-bromobutyrate derivatives (5af) were synthesized and frontier molecular orbitals (FMOs) and molecular electrostatic potential (MEP) analysis were done. FMOs analysis revealed that among all the derivatives, 5c is most reactive, having a HOMO–LUMI band gap of 4.27 eV, whereas the HOMO–LUMO band-gap for 5f was 5.19 eV, and was the most stable. The chemical hardness of 5f displayed the highest value (2.60 eV), whereas 5c has the lowest value and was chemically soft and reactive. Among all the newly synthesized compounds, 5d displayed the highest chemical potential value. The molecular electrostatic potential (MEP) investigation gave us the idea of electrophilicity and nucleophilicity of the synthesized compounds, and it was envisaged that depression of electron density is highly dependent on the substituent present on an aromatic ring. The first hyperpolarizability values showed that all synthesized derivatives has a potential non-linear optics (NLO) response, but 5c has a significant NLO potential due to its large βo (1373.76 au) value.

Author Contributions

U.N., A.M. (Aqsa Mujahid), A.M. (Asim Mansha), M.Z., A.R.S., T.M. and N.K. conceived and designed the experiment and implemented the experiments; N.R. supervised and wrote, reviewed, and edited the manuscript; S.A.A.S., Z.A.Z., and M.N.A. analyzed the data and wrote the original draft, helped in experiments, and contributed reagents/material and analysis tools. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors would like to acknowledge the Ministry of Higher Education (MOHE) Malaysia for financial support under the Fundamental Research Grant Scheme (FRGS) with sponsorship reference numbers FRGS/1/2019/STG05/UITM/02/9. The authors would also like to acknowledge Universiti Teknologi MARA for the financial support under the reference number 600-IRMI/FRGS 5/3 (424/2019). The data reported herein is part of the M. Phil thesis research work of Usman Nazeer.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tawata, S.; Taira, S.; Kobamoto, N.; Zhu, J.; Ishihara, M.; Toyama, S. Synthesis and antifungal activity of cinnamic acid esters. Biosci. Biotechnol. Biochem. 1996, 60, 909–910. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Harini, S.T.; Kumar, H.V.; Rangaswamy, J.; Naik, N. Synthesis, antioxidant and antimicrobial activity of novel vanillin derived piperidin-4-one oxime esters: Preponderant role of the phenyl ester substituents on the piperidin-4-one oxime core. Bioorg. Med. Chem. Lett. 2012, 22, 7588–7592. [Google Scholar] [CrossRef] [PubMed]
  3. Rainsford, K.D.; Whitehouse, M.W. Anti-inflammatory/anti-pyretic salicylic acid esters with low gastric ulcerogenic activity. Agents Act. 1980, 10, 451–456. [Google Scholar] [CrossRef] [PubMed]
  4. Reddy, Y.S.; Kaki, S.S.; Rao, B.B.; Jain, N.; Vijayalakshm, P. Study on Synthesis, Characterization and Antiproliferative Activity of Novel Diisopropylphenyl Esters of Selected Fatty Acids. J. Oleo Sci. 2016, 65, 81–89. [Google Scholar] [CrossRef] [Green Version]
  5. Tandon, V.K.; Yadav, D.B.; Singh, R.V.; Chaturvedi, A.K.; Shukla, P.K. Synthesis and biological evaluation of novel (l)-α-amino acid methyl ester, heteroalkyl, and aryl substituted 1,4-naphtho quinone derivatives as antifungal and antibacterial agents. Bioorg. Med. Chem. Lett. 2005, 15, 5324–5328. [Google Scholar] [CrossRef] [PubMed]
  6. Rottig, A.; Wenning, L.; Broker, D.; Steinbuchel, A. Fatty acid alkyl esters: Perspectives for production of alternative biofuels. Appl. Microbiol. Biotechnol. 2010, 85, 1713–1733. [Google Scholar] [CrossRef] [PubMed]
  7. Giam, C.S.; Chan, H.S.; Neff, G.S.; Atlas, E.L. Phthalate ester plasticizers: A new class of marine pollutant. Science 1978, 199, 419–421. [Google Scholar] [CrossRef]
  8. Keng, P.S.; Basri, M.; Zakaria, M.R.S.; Rahman, M.A.; Ariff, A.B.; Rahman, R.A.; Salleh, A.B. Newly synthesized palm esters for cosmetics industry. Ind. Crops Prod. 2009, 29, 37–44. [Google Scholar] [CrossRef]
  9. Suarez, G.S.; Massa, N.E.; Jubert, A.H.; Jios, J.L.; Autino, J.C.; Romanelli, G.P. Spectroscopic and theoretical study of 2-acetylphenyl-2-naphthoate. Spectrochim. Acta Part A 2009, 71, 1989–1998. [Google Scholar] [CrossRef]
  10. Yamamura, K.; Ono, S.; Tabushi, I. New liquid crystals having 4, 4′-biphenanthryl core. Tetrahedron Lett. 1988, 29, 1797–1798. [Google Scholar] [CrossRef]
  11. Hohnholz, D.; Schweikart, K.H.; Subramanian, L.R.; Wedel, A.; Wischert, W.; Hanack, M. Synthesis and studies on luminescent biphenyl compounds. Synth. Met. 2000, 110, 141–152. [Google Scholar] [CrossRef]
  12. Bakheit, A.H.H.; Abd-Elgalil, A.A.; Mustafa, B.; Haque, A.; Wani, T.A. Telmisartan. In Profiles of Drug Substances, Excipients and Related Methodology; Academic Press: San Diego, CA, USA, 2015. [Google Scholar]
  13. Aronson, J. Meyler’s Side Effects of Drugs, 16th ed.; Elsevier: Amsterdam, The Netherlands, 2016. [Google Scholar]
  14. Mutahir, S.; Jonczyk, J.; Bajda, M.; Khan, I.U.; Khan, M.A.; Ullah, N.; Ashraf, M.; Riaz, S.; Hussain, S.; Yar, M. Novel biphenyl bis-sulfonamides as acetyl and butyrylcholinesterase inhibitors: Synthesis, biological evaluation and molecular modeling studies. Bioorg. Chem. 2016, 64, 13–20. [Google Scholar] [CrossRef] [PubMed]
  15. Ding, Y.; Mao, L.; Xu, D.; Xie, H.; Yang, L.; Xu, H.; Yang, L.; Xu, H.; Geng, W.; Gao, Y.; et al. C-Aryl glucoside SGLT2 inhibitors containing a biphenyl motif as potential anti-diabetic agents. Bioorgan. Med. Chem. Lett. 2015, 25, 2744–2748. [Google Scholar] [CrossRef] [PubMed]
  16. Cincinelli, R.; Zwick, V.; Musso, L.; Zuco, V.; De Cesare, M.; Zunino, F.; Simoes-Pires, C.; Nussiro, A.; Giannini, G.; Cuendet, M.; et al. Biphenyl-4-yl-acrylohydroxamic acids: Identification of a novel indolyl-substituted HDAC inhibitor with antitumor activity. Eur. J. Med. Chem. 2016, 112, 99–105. [Google Scholar] [CrossRef] [PubMed]
  17. Kong, X.W.; Zhang, Y.H.; Dai, L.; Ji, H.; Lai, Y.S.; Peng, S.X. Synthesis and biological evaluation ofnitric oxide-releasing sixalkoxyl biphenyl derivatives as anticancer agents. Chin. Chem. Lett. 2008, 19, 149–152. [Google Scholar] [CrossRef]
  18. Dong, J.; Pan, X.; Wang, J.; Su, P.; Zhang, L.; Wei, F.; Zhang, J. Synthesis and biological evaluation of novel aromatic-heterocyclic biphenyls as potent anti-leukemia agents. Eur. J. Med. Chem. 2015, 101, 780–789. [Google Scholar] [CrossRef]
  19. Brudeli, B.; Andressen, K.W.; Moltzau, L.R.; Nilsen, N.O.; Levy, F.O.; Klaveness, J. Acidic biphenyl derivatives: Synthesis and biological activity of a new series of potent 5-HT4 receptor antagonists. Bioorg. Med. Chem. 2013, 21, 7134–7145. [Google Scholar] [CrossRef]
  20. Gargano, E.M.; Perspicace, E.; Carotti, A.; Marchais-Oberwinkler, S.; Hartmann, R.W. Addressing cytotoxicity of 1,4-biphenyl amide derivatives: Discovery of new potent and selective 17β-hydroxysteroid dehydrogenase type 2 inhibitors. Bioorg. Med. Chem. Lett. 2016, 26, 21–24. [Google Scholar] [CrossRef]
  21. Rizwan, K.; Zubair, M.; Rasool, N.; Ali, S.; Zahoor, A.F.; Rana, U.A.; Jaafar, H.Z. Regioselective synthesis of 2-(bromomethyl)-5-aryl-thiophene derivatives via palladium (0) catalyzed suzuki cross-coupling reactions: As antithrombotic and haemolytically active molecules. Chem. Cent. J. 2014, 8, 1–8. [Google Scholar] [CrossRef] [Green Version]
  22. Del Rio, N.; Baceiredo, A.; Saffon-Merceron, N.; Hashizume, D.; Lutters, D.; Müller, T.; Kato, T. A Stable Heterocyclic Amino (phosphanylidene-σ4-phosphorane) Germylene. Angew. Chem. Int. Ed. 2016, 55, 4753–4758. [Google Scholar] [CrossRef]
  23. Gull, Y.; Rasool, N.; Noreen, M.; Nasim, F.U.H.; Yaqoob, A.; Kousar, S.; Islam, M. Efficient synthesis of 2-amino-6-arylbenzothiazoles via Pd (0) Suzuki cross coupling reactions: Potent urease enzyme inhibition and nitric oxide scavenging activities of the products. Molecules 2013, 18, 8845–8857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Nayak, P.K.; Periasamy, N. Calculation of electron affinity, ionization potential, transport gap, optical band gap and exciton binding energy of organic solids using ‘solvation’ model and DFT. Org. Electron. 2009, 10, 1396–1400. [Google Scholar] [CrossRef]
  25. Arshad, M.N.; Asiri, A.M.; Alamry, K.A.; Mahmood, T.; Gilani, M.A.; Ayub, K.; Birinji, A.S. Synthesis, crystal structure, spectroscopic and density functional theory (DFT) study of N-[3-anthracen-9-yl-1-(4-bromophenyl)-allylidene]-N-benzenesulfonohydrazine. Spectrochim. Acta Part A 2015, 142, 364–374. [Google Scholar] [CrossRef] [PubMed]
  26. Arshad, M.N.; Bibi, A.; Mahmood, T.; Asiri, A.M.; Ayub, K. Synthesis, crystal structures and spectroscopic properties of triazine-based hydrazone derivatives; A comparative experimental-theoretical study. Molecules 2015, 20, 5851–5874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Mebi, C.A. DFT study on structure, electronic properties, and reactivity of cis-isomers of [(NC5H4-S) 2 Fe (CO)2]. J. Chem. Sci. 2011, 123, 727–731. [Google Scholar] [CrossRef] [Green Version]
  28. Parr, R.G.; Szentpaly, L.V.; Liu, S. Electrophilicity index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  29. Ahmed, M.N.; Yasin, K.A.; Ayub, K.; Mahmood, T.; Tahir, M.N.; Khan, B.A.; Hafeez, M.; Ahmed, M. Click one pot synthesis, spectral analyses, crystal structures, DFT studies and brine shrimp cytotoxicity assay of two newly synthesized 1,4,5-trisubstituted 1,2,3-triazoles. J. Mol. Struct. 2016, 1106, 430–439. [Google Scholar] [CrossRef]
  30. Marder, S.R. Organic nonlinear optical materials: Where we have been and where we are going. Chem. Commun. 2006, 2, 131–134. [Google Scholar] [CrossRef]
  31. Champagne, B.; Plaquet, A.; Pozzo, J.L.; Rodriguez, V.; Castet, F. Nonlinear Optical Molecular Switches as Selective Cation Sensors. J. Am. Chem. Soc. 2012, 134, 8101–8103. [Google Scholar] [CrossRef]
  32. Tarazkar, M.; Romanov, D.A.; Levis, R.J. Theoretical study of second-order hyperpolarizability for nitrogen radical cation. J. Phys. B At. Mol. Opt. Phys. 2015, 48, 094019. [Google Scholar] [CrossRef] [Green Version]
  33. Frisch, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. Gaussian 09, Rev. B. 01; Gaussian Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
  34. Dennington, R.; Todd, K.; John, M. Gauss View, Version 5; Semichem Inc.: Shawnee Mission, KS, USA, 2009. [Google Scholar]
Sample Availability: Samples of the compounds are not available from the authors.
Scheme 1. Synthesis of 2-bromo-4-chlorophenyl-2-bromobutyrate (3). Conditions: (i) 1 (1 g, 4.820 mmol), 2 (0.64 mL, 5.3 mmol), pyridine (0.7 mL, 9.63 mmol), MeCN (30 mL), 0 °C, 1 h.
Scheme 1. Synthesis of 2-bromo-4-chlorophenyl-2-bromobutyrate (3). Conditions: (i) 1 (1 g, 4.820 mmol), 2 (0.64 mL, 5.3 mmol), pyridine (0.7 mL, 9.63 mmol), MeCN (30 mL), 0 °C, 1 h.
Molecules 25 03521 sch001
Scheme 2. Synthesis of 2-aryl-4-chlorophenyl-2-bromobutyrates (5a–f). Conditions: (i) 3 (200 mg, 0.062 mmol, 1 eq), 4 (0.683 mmol, 1.1 eq), Pd (PPh3)4 (35 mg, 5 mol %), K3PO4 (263 mg, 1.24 mmol, 2.2 eq), 1, 4-dioxane (8 mL), water (4 mL), 90 °C, 17 h under argon.
Scheme 2. Synthesis of 2-aryl-4-chlorophenyl-2-bromobutyrates (5a–f). Conditions: (i) 3 (200 mg, 0.062 mmol, 1 eq), 4 (0.683 mmol, 1.1 eq), Pd (PPh3)4 (35 mg, 5 mol %), K3PO4 (263 mg, 1.24 mmol, 2.2 eq), 1, 4-dioxane (8 mL), water (4 mL), 90 °C, 17 h under argon.
Molecules 25 03521 sch002
Figure 1. FMOs plot of all compounds (5af) based on the relative energies; the energies are in eV.
Figure 1. FMOs plot of all compounds (5af) based on the relative energies; the energies are in eV.
Molecules 25 03521 g001
Figure 2. Molecular electrostatic potential surfaces of compounds (5af), the electronic density in red represents nucleophilic sites and in blue represents electrophilic site.
Figure 2. Molecular electrostatic potential surfaces of compounds (5af), the electronic density in red represents nucleophilic sites and in blue represents electrophilic site.
Molecules 25 03521 g002
Figure 3. Graphical representations of dc-Kerr, ESHG coefficients, and refractive index for compounds (5af) at different wavelengths (532 and 1064 nm).
Figure 3. Graphical representations of dc-Kerr, ESHG coefficients, and refractive index for compounds (5af) at different wavelengths (532 and 1064 nm).
Molecules 25 03521 g003
Table 1. Suzuki–Miyaura reaction of 2-bromo-4-chlorophenyl-2-bromobutyrate with different phenyl boronic acids.
Table 1. Suzuki–Miyaura reaction of 2-bromo-4-chlorophenyl-2-bromobutyrate with different phenyl boronic acids.
Sr. No.Phenyl Boronic AcidProductsSolvent/H2O (4:1)Yield %
5a3-Cl-4-F-C6H4 Molecules 25 03521 i0011,4-dioxane64
5b3-(MeCO2)-C6H4 Molecules 25 03521 i0021,4-dioxane76
5c4-(MeS)-C6H4 Molecules 25 03521 i0031,4-dioxane79
5d3,4-Cl2-C6H4 Molecules 25 03521 i0041,4-dioxane68
5e4-MeO-C6H4 Molecules 25 03521 i0051,4-dioxane81
5f3-Cl-C6H4 Molecules 25 03521 i0061,4-dioxane73
Table 2. The energies of HOMOs (EHOMOs, in eV), energies of LUMOs (ELUMOs, in eV,) HOMOs–LUMOs gap (GH-L), and dipole moment (μ, in Debye) of compounds (5af).
Table 2. The energies of HOMOs (EHOMOs, in eV), energies of LUMOs (ELUMOs, in eV,) HOMOs–LUMOs gap (GH-L), and dipole moment (μ, in Debye) of compounds (5af).
CompoundsEHOMOsELUMOsGH-L
5a−7.06−1.935.14
5b−7.03−1.975.06
5c−6.06−1.794.27
5d−7.07−1.985.09
5e−6.30−1.734.57
5f−7.06−1.875.19
Table 3. Ionization potential (I), electron affinity (EA), chemical hardness (Ƞ), electronic chemical potential (μ), electrophilicity index (Ꙍ), and nucleophilicity (N) of compounds (5af). All values are given in eV.
Table 3. Ionization potential (I), electron affinity (EA), chemical hardness (Ƞ), electronic chemical potential (μ), electrophilicity index (Ꙍ), and nucleophilicity (N) of compounds (5af). All values are given in eV.
CompoundsIonization Potential (I)Electron Affinity (EA)Chemical Hardness (Ƞ)Electronic Chemical Potential (μ)Electrophilicity Index (Ꙍ)Nucleophilicity (N)
5a7.061.932.57−4.503.945.14
5b7.031.972.53−4.504.005.06
5c6.061.792.13−3.923.614.27
5d7.071.982.55−4.524.025.09
5e6.301.732.29−4.013.524.57
5f7.061.872.60−4.463.835.19
Table 4. Molecular electrostatic potential (MEP) values of all compounds (5af).
Table 4. Molecular electrostatic potential (MEP) values of all compounds (5af).
Compounds−ve Potential+ve Potential
5a−4.77 × 10−24.77 × 10−2
5b−5.01 × 10−25.01 × 10−2
5c−5.13 × 10−25.13 × 10−2
5d−4.75 × 10−24.75 × 10−2
5e−5.22 × 10−25.22 × 10−2
5f−4.95 × 10−24.95 × 10−2
Table 5. Polarizability (αo) and hyperpolarizability (βo) of compounds 5af with CAM-B3LYP and LC-BLYP methods.
Table 5. Polarizability (αo) and hyperpolarizability (βo) of compounds 5af with CAM-B3LYP and LC-BLYP methods.
CompoundsCAM-B3LYPLC-BLYP
αoβoαoβo
5a236515.43229436.87
5b252284.34244223.17
5c2641373.762551010.54
5d250539.03243445.12
5e244954.47237776.21
5f235218.51228173.78
Table 6. The static, dc-Kerr effect, electric field-induced second harmonic generation coefficients (in au), and quadratic non-linear refractive index (n2 = cm2/W) at different wavelengths (in nm) for compounds (5af).
Table 6. The static, dc-Kerr effect, electric field-induced second harmonic generation coefficients (in au), and quadratic non-linear refractive index (n2 = cm2/W) at different wavelengths (in nm) for compounds (5af).
CompoundsFrequencyγ(0; 0, 0, 0)γ(−ω; ω, ω, 0)γ(−2ω; ω, ω, 0)n2
5a048741.5
532 64320.81995298.63 × 10−18
1064 51988.459645.34.43 × 10−18
5b054142.2
532 73258.12934321.16 × 10−17
1064 58215.367505.44.96 × 10−18
5c086291.9
532 141361340643019.46 × 10−16
1064 95936.81225898.41 × 10−18
5d057794.5
532 78682.23141301.24 × 10−17
1064 6186072208.85.3 × 10−18
5e063232
532 92924.81013010.13.23 × 10−17
1064 69085.683451.75.96 × 10−18
5f049285.2
532 64440.91764088.01 × 10−18
1064 52669.560014.74.47 × 10−18

Share and Cite

MDPI and ACS Style

Nazeer, U.; Rasool, N.; Mujahid, A.; Mansha, A.; Zubair, M.; Kosar, N.; Mahmood, T.; Raza Shah, A.; Shah, S.A.A.; Zakaria, Z.A.; et al. Selective Arylation of 2-Bromo-4-chlorophenyl-2-bromobutanoate via a Pd-Catalyzed Suzuki Cross-Coupling Reaction and Its Electronic and Non-Linear Optical (NLO) Properties via DFT Studies. Molecules 2020, 25, 3521. https://doi.org/10.3390/molecules25153521

AMA Style

Nazeer U, Rasool N, Mujahid A, Mansha A, Zubair M, Kosar N, Mahmood T, Raza Shah A, Shah SAA, Zakaria ZA, et al. Selective Arylation of 2-Bromo-4-chlorophenyl-2-bromobutanoate via a Pd-Catalyzed Suzuki Cross-Coupling Reaction and Its Electronic and Non-Linear Optical (NLO) Properties via DFT Studies. Molecules. 2020; 25(15):3521. https://doi.org/10.3390/molecules25153521

Chicago/Turabian Style

Nazeer, Usman, Nasir Rasool, Aqsa Mujahid, Asim Mansha, Muhammad Zubair, Naveen Kosar, Tariq Mahmood, Ali Raza Shah, Syed Adnan Ali Shah, Zainul Amiruddin Zakaria, and et al. 2020. "Selective Arylation of 2-Bromo-4-chlorophenyl-2-bromobutanoate via a Pd-Catalyzed Suzuki Cross-Coupling Reaction and Its Electronic and Non-Linear Optical (NLO) Properties via DFT Studies" Molecules 25, no. 15: 3521. https://doi.org/10.3390/molecules25153521

Article Metrics

Back to TopTop