Next Article in Journal
Treatment of Wastewaters with Zirconium Phosphate Based Materials: A Review on Efficient Systems for the Removal of Heavy Metal and Dye Water Pollutants
Next Article in Special Issue
LP1 and LP2: Dual-Target MOPr/DOPr Ligands as Drug Candidates for Persistent Pain Relief
Previous Article in Journal
Chondroprotection and Molecular Mechanism of Action of Phytonutraceuticals on Osteoarthritis
Previous Article in Special Issue
Pretreatment with High Mobility Group Box-1 Monoclonal Antibody Prevents the Onset of Trigeminal Neuropathy in Mice with a Distal Infraorbital Nerve Chronic Constriction Injury
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Analgesic and Anticancer Activity of Benzoxazole Clubbed 2-Pyrrolidinones as Novel Inhibitors of Monoacylglycerol Lipase

by
Obaid Afzal
1,*,
Abdulmalik Saleh Alfawaz Altamimi
1,*,
Mir Mohammad Shahroz
2,
Hemant Kumar Sharma
2,
Yassine Riadi
1 and
Md Quamrul Hassan
3
1
Department of Pharmaceutical Chemistry, College of Pharmacy, Prince Sattam Bin Abdulaziz University, Al Kharj 11942, Saudi Arabia
2
Department of Pharmaceutical Chemistry, College of Pharmacy, Sri Satya Sai University of Technology and Medical Sciences, Sehore 466001, Madhya Pradesh, India
3
Department of Pharmacology, School of Pharmaceutical Education and Research, Jamia Hamdard, New Delhi 110062, India
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(8), 2389; https://doi.org/10.3390/molecules26082389
Submission received: 20 March 2021 / Revised: 7 April 2021 / Accepted: 14 April 2021 / Published: 20 April 2021
(This article belongs to the Special Issue Novel Antinociceptive Agent against Persistent Pain)

Abstract

:
Ten benzoxazole clubbed 2-pyrrolidinones (1120) as human monoacylglycerol lipase inhibitors were designed on the criteria fulfilling the structural requirements and on the basis of previously reported inhibitors. The designed, synthesized, and characterized compounds (1120) were screened against monoacylglycerol lipase (MAGL) in order to find potential inhibitors. Compounds 19 (4-NO2 derivative) and 20 (4-SO2NH2 derivative), with an IC50 value of 8.4 and 7.6 nM, were found most active, respectively. Both of them showed micromolar potency (IC50 value above 50 µM) against a close analogue, fatty acid amide hydrolase (FAAH), therefore considered as selective inhibitors of MAGL. Molecular docking studies of compounds 19 and 20 revealed that carbonyl of 2-pyrrolidinone moiety sited at the oxyanion hole of catalytic site of the enzyme stabilized with three hydrogen bonds (~2 Å) with Ala51, Met123, and Ser122, the amino acid residues responsible for the catalytic function of the enzyme. Remarkably, the physiochemical and pharmacokinetic properties of compounds 19 and 20, computed by QikProp, were found to be in the qualifying range as per the proposed guideline for good orally bioactive CNS drugs. In formalin-induced nociception test, compound 20 reduced the pain response in acute and late stages in a dose-dependent manner. They significantly demonstrated the reduction in pain response, having better potency than the positive control gabapentin (GBP), at 30 mg/kg dose. Compounds 19 and 20 were submitted to NCI, USA, for anticancer activity screening. Compounds 19 (NSC: 778839) and 20 (NSC: 778842) were found to have good anticancer activity on SNB-75 cell line of CNS cancer, exhibiting 35.49 and 31.88% growth inhibition (% GI), respectively.

1. Introduction

Endocannabinoids (endogenous ligands), cannabinoid (CB) receptors, and proteins for their biological synthesis and degradation constitute the endocannabinoid system (ECS) [1]. Endocannabinoids are biosynthesized from the membrane phospholipids [2]. Endocannabinoid, N-arachidonoyl ethanolamine (AEA, Anandamide) functions as partial agonist on CB1 and CB2 receptors. It has low affinity for CB2 and moderate affinity for CB1. Endocannabinoids, 2-arachidonoylglycerol (2-AG) function as full agonist and have moderate affinity for both the receptors. Interestingly, 2-AG is the major endocannabinoid and is found to be approximately 170-fold higher in concentration than AEA, in the brain [3]. AEA and 2-AG hydrolysis and degradation are facilitated by fatty acid amide hydrolase (FAAH) and monoacylglycerol lipase (MAGL) enzymes, correspondingly [1]. MAGL (an α/β-hydrolase) hydrolyzes 2-AG into glycerol and free fatty acid (FFA), by the action of the catalytic triad organized with Ser122, Asp239 and His269 amino acids in the active site of the enzyme [4]. MAGL hydrolyzes approximately 85% of total 2-AG in the CNS [5] and generates arachidonic acid (AA), giving rise to neuro-inflammatory PGE2 and PGD2 prostaglandins [6]. MAGL inhibitors ameliorates neuropathic pain by increasing 2-AG and decreasing neuro-inflammatory prostaglandins in the CNS (Figure 1) [7]. In addition, 2-AG demonstrated analgesic activity by acting on CB1 receptors in the CNS and periphery [8,9,10,11]. It is reported that JJKK-048 (MAGL, IC50 363 pM) exhibited analgesic activity in tail immersion and writhing test [12]. MAGL inhibition has shown significant neuroprotective and anti-inflammatory potential in Parkinson’s and Alzheimer’s disease [13,14]. PF-06795071 (IC50 3 nM), a MAGL inhibitor, is reported to have considerable anti-neuroinflammatory potential [15]. Another MAGL inhibitor, ABX-1431 (IC50 14 nM) is under clinical trials for broad range of CNS disorders like Tourette syndrome [16]. Many more research findings encouraged that the inhibitors of MAGL has therapeutic potentials in pain and CNS disorders [17,18,19,20].
The governing importance of MAGL in abnormal lipolysis in cancer has been demonstrated [21]. MAGL was originally known for its lipolytic action on monoacylglycerols from stored triacylglycerols into glycerol and free fatty acids (FFA) [22]. Cancer cells utilizes this lipolytic pathways for their hastened proliferation [23]. The cancer-supporting action of MAGL is due to elevated FFA levels. This MAGL-FFA pathway promotes in vivo tumor growth by increasing FFA-derived oncogenic signaling lipids (PA, LPA, S1P, and PGE2) [18]. These protumorigenic lipid mediators encourage tumor growth, angiogenesis, and metastasis in cancer (Figure 1) [24]. MAGL is reported to expressed vastly in aggressive type of cancer cells and is associated with pathogenesis, proliferation, and in vivo tumor growth. MAGL inhibition disrupts cancer cell proliferation, growth and metastasis [25,26,27]. The anticancer effect of MAGL inhibition in prostate cancer was totally abolished by cotreatment with SR141716 (rimonabant; CB1 receptor antagonist) and fatty acids, signifying that amplified endocannabinoid action and reduced stock of FFA from MAGL inhibition is the reason behind antitumor effect [26].
MAGL inhibitors identified till date includes URB602 (IC50 28 µM) [28], CAY10499 (IC50 144 nM) [29], SPB01403 (IC50 31 µM) [30], JZL184 (IC50 8 nM) [31], SAR629 (IC50 1.1 nM) [32], KML29 (IC50 3.6 nM) [33], ML30 (IC50 0.54 nM) [34], JJKK-048 (IC50 363 pM) [35], ABX1431 (IC50 14 nM) [16], R(3t) (IC50 3.6 nM) [36], PF-06795071 (IC50 3 nM) [15], a benzoylpiperidine derivative (IC50 80 nM) [37], and a benzisothiazolinone derivative (IC50 34.1 nM) [38]. JJKK-048 and KML29 both were reported to be highly selective MAGL inhibitors, and their selectivity is more than 10,000-fold over FAAH. [33,35]. In addition, 3D crystal structures of human MAGL enzyme were elucidated by X-ray crystallography, and are available on protein data bank [32,36,39,40]. A comprehensive description of MAGL crystal structures and inhibitors were reviewed [41]. Our group has also identified nanomolar MAGL inhibitors, viz. ZINC24092691 (IC50 10 nM), ZINC12863377 (IC50 39 nM), a ZINC24092691 analogue (IC50 6.5 nM), a thiazole-5-carboxylate derivative (IC50 37 nM), and a pyrrolidin-2-one linked benzimidazole derivative (compound 25; IC50 9.4 nM) [42,43,44,45]. In continuation of our work on MAGL inhibitors, we have designed novel 2-pyrrolidinone linked benzoxazole derivatives and screened them for analgesic and anticancer effects.

2. Results

The binding pattern of pyrrolidin-2-one derivatives (ZINC12863377, compound 25 and compound R-3t), and the basic structural requirement for MAGL inhibitors was kept in mind to design novel pyrrolidin-2-one linked benzoxazole derivatives (Figure 2). The route of synthesis of compounds (1120) is presented in Figure 3.

2.1. Chemistry

Intermediate compound 1, (1-Benzyl-5-oxopyrrolidine-3-carboxylic acid), was successfully prepared by fusion of benzylamine and methylidenesuccinic acid in water, while compounds 2–10 were synthesized as reported by us in our previous publication [45]. The fusion of synthesized 1-substituted-5-oxopyrrolidine-3-carboxylic acids (1–10) with 2-aminophenol was done by the procedure reported earlier with some minor modifications [46]. The fusion of acids (1–10) and 2-aminophenol was carried out by the use of polyphosphoric acid, giving better yields (57–70%) and purity of benzoxazole derivatives (1120).
The prototype intermediate compound 1, (1-Benzyl-5-oxopyrrolidine-3-carboxylic acid), revealed typical peaks at 1627 cm−1 (carbonyl of acid), 1734 cm−1 (carbonyl of pyrrolidinone), and 3241 cm−1 (O-H of acid) in IR spectrum. In 1H-NMR spectrum of compound 1, two methyl protons were found resonating as singlet at δ 3.62. The multiplets resonating at δ 2.62–2.76 indicated two COCH2 protons, and at δ 4.29–4.44 they indicated two NCH2 and one CH proton of 2-pyrrolidinone ring. Five aromatic protons of the benzyl ring were found resonating at δ 7.20–7.36 as a multiplet. The typical singlet was assigned to COOH proton at δ 11.35 and was found to be D2O exchangeable. Final prototype compound 11 (4-(Benzo[d]oxazol-2-yl)-1-benzylpyrrolidin-2-one) exhibited specific IR bands at 1621 cm−1 (C=N of benzoxazole) and 1703 cm−1 (C=O of pyrrolidin-2-one). The two protons of CH2 of benzyl appeared at δ 3.95–4.01 as a multiplet. Two COCH2 protons of the 2-pyrrolidinone ring was assigned at δ 2.60–2.77 as a multiplet. The multiplets located at δ 3.13–3.18 and 3.81–3.86 were ascribed to one CH proton and two NCH2 protons of the 2-pyrrolidinone ring, correspondingly. Six aromatic protons were found resonating at δ 7.38–7.58 as a multiplet. The doublet resonating at δ 7.73–7.78 was assigned to the aromatic protons of benzoxazole. The other two protons of benzoxazole were found resonating at δ 7.94–7.96 as doublet. The M+ (molecular ion) peak of compound 11 was found at 292.14, validating its successful synthesis.

2.2. Human MAGL Assay

The assay was executed by Cayman’s assay kit by the reported procedure [29]. All the ten compounds (1120) were screened for hMAGL inhibitory potential. The substituted phenyl derivatives (1320) were established to reduce the MAGL activity at 100 μM concentration below 50%. Compound 19 (4-NO2 derivative) and compound 20 (4-SO2NH2 derivative) were the most potent, with an IC50 of 8.4 and 7.6 nM, correspondingly. The structure–function relationship of benzoxazole derivatives is displayed hMAGL inhibitory activity as follows: 4-SO2NH2 > 4-NO2 > 3-Cl,4-F > 4-OCH3 > 4-Cl > 4-OH > 4-CH3 > 2-CH3 > phenyl/benzyl. The IC50 of standard controls, selective MAGL inhibitors, CAY10499 (IC50 = 415 nM), and for JZL184 (IC50 = 10 nM) were comparable to the reported values [29]. The outcomes of the experiments are presented in Table 1.

2.3. Human FAAH Assay

Derivatives having IC50 in nanomolar range (compound 15, 16, 18, 19, and 20) were nominated for further screening against FAAH, an allied hydrolase of MAGL [47]. They displayed micromolar potency, with an IC50 value ranging from 25 to 68 µM against FAAH. The benzoxazole derivatives having 4-NO2 phenyl (19) and 4-SO2NH2 phenyl (20), with an FAAH IC50 value greater than 50 µM, were considered selective MAGL inhibitors. The IC50 of standard control, URB597 (selective FAAH inhibitor), was 5 nM, equivalent to the value reported (IC50 = 4.6 nM) [48]. The outcomes of the experiments are presented in Table 1.

2.4. Molecular Docking Study

Most active and selective compounds identified by MAGL and FAAH inhibition assay (19 and 20), were docked at the catalytic center of MAGL, in order to get an insight of their binding pattern, with the help of XP Glide docking using Maestro (Schrodinger). Compounds 19 and 20, showed comparable docking scores of −9.87 and −9.83, respectively. The binding of compounds 19 and 20 in the active site of MAGL revealed that the carbonyl group of pyrrolidinone is located exactly in the oxyanion hole and stabilized by three hydrogen bonds (~2Å) with alanine 51, serine 122, and methionine 123. Serine 122 is one of the critical amino acid residues of the catalytic triad of MAGL. The benzoxazole moiety is found to be positioned in the amphiphilic pouch, having π-π stacking contact with the amino acid Tyr194. The 4-NO2 (19) and 4-SO2NH2 (20) phenyl ring of the ligands were involved in hydrophobic (van der Waals) attractions with the amino acids, leucine 148, 213, and 241. In addition, the 3D and 2D binding pattern of compounds 19 and 20 in the catalytic location of MAGL is depicted in Figure 4.

2.5. Pharmacokinetic and Physicochemical Characteristics

To investigate the potential of the identified derivatives (19 and 20) to cross the selectively permeable membranes of hematoencephalic barrier (BBB), to develop orally active CNS drugs, their pharmacokinetic and physicochemical features were computed by QikProp (ADMET predictor) of Schrodinger. Guidelines, concerning the validation and optimization of orally active CNS compounds, were developed by Ghose et. al., by analyzing 35 characteristic features of orally bioavailable 317 CNS and 626 non-CNS drugs [49]. This guideline states that in order to design high-quality CNS drugs, the molecule must qualify by the following parameters: TPSA less than 76 Å2 (ideally 25–60 Å2), number of N atoms between 1–2, comprising 1 aliphatic amine, 2–4 side chains on/outside rings, number of polar H atoms ˂ 3 (ideally 0–1), SASA 460–580 Å2, molecular volume 740–970 Å3, and must have +ve QikProp CNS property. Remarkably, most of the properties of compounds 19 and 20, computed by QikProp, were found to be in the qualifying range as per the proposed guideline (Table 2). The properties of compound 19 were found to be within the qualifying range except dipole moment. For compound 20, 5 out of 35 properties is just slightly above the upper qualifying limit. Most importantly, qualifying limits for CNS active drugs in terms of TPSA is from 3.8 to 109, and the calculated TPSA for compound 20 was found to be 119.08. Therefore, the designing of more potent MAGL inhibitors having physicochemical and pharmacokinetic properties within the preferred CNS limits is required.

2.6. In Silico Absorption and Toxicity Profile

The selected compounds (19 and 20) were evaluated for their absorption and toxicity profile by a bioinformatics tool admetSAR [50]. The results suggested that both the compounds have high blood–brain barrier (BBB) penetration properties as well as high chance of human intestinal absorption. In AMES test, compound 19 was found to be mutagenic, while 20 was non-mutagenic. Carcinogenicity test revealed that both the compounds were non-carcinogens. The LD50 values in rat were also evaluated, a compound with high value is considered as less lethal. The LD50 for compounds 19 and 20 were found to be 2.30 and 2.21 mol/kg, respectively. Overall, compound 20 has better toxicity profile as compared to compound 19 (Table 3).

2.7. Analgesic Activity

The formalin-induced analgesic test is an extensively acknowledged animal nociception model. In order to evaluate both central and peripheral effects of the compound (20), formalin-induced nociception model was selected for analgesic activity. The formalin induced behavioral response comprises two typical phases, stage I and II. Stage I persists up to five minutes after formalin injection and is characterized by acute pain with vigorous licking and biting of the injected site. Stage I consists the action formalin on afferent C-fiber nociceptors. While Stage II starts 10–30 min after formalin injection and persists till 60 min, characterized by periodic licking and biting of the injected site. Stage II imitates the action of central sensitization of the spinal dorsal horn neurons [51,52]. Compound 20 was selected for analgesic activity due to its higher potency (IC50 7.6 nM). Compound 20 (suspensions prepared with 0.5% CMC) were administered per oral (p.o, in doses of 5, 10, 30, and 50 mg/kg body weight, 4 h prior to the formalin injection. Gabapentin (GBP), (dissolved in 0.9% normal saline), was chosen as positive control (reference drug) and administered intraperitoneal (i.p) in 100 mg/kg dose. GBP exhibited little analgesic effects in Stage I (acute nociception), in comparison to the control (0.5% CMC). Though, in Stage II, it displayed significant reduction of paw licking and biting, endorsing GBP central effects. However, compound 20, reduced the pain response significantly both in acute (Stage I) and late (Stage II) phases, in a dose-dependent manner. They significantly demonstrated the reduction in pain response, having better potency than the positive control GBP at 30 mg/kg. The duration (in seconds) of paw licking and paw biting throughout Stage I and II is provided in Figure 5.

2.8. Anticancer Activity

Compounds 19 and 20 were supplied to National Cancer Institute (USA), for sulforhodamine B (SRB) assay and anticancer screening [53,54]. Single-dose (10 µM) assay results for compounds 19 and 20 were provided as a mean of percent growth (% G) and growth inhibition (% GI) against 60 cell lines of nine types of cancers and are tabulated in Table 4. Derivatives 19 (NSC: 778839) and 20 (NSC: 778842) were found to have good anticancer activity towards SNB-75 cell line of CNS cancer, having % growth inhibition (% GI) of 35.49 and 31.88, respectively. Compound 20 showed 22.22 and 18.03% GI of HOP-92 and HOP-62 cell lines of non-small cell lung cancer, respectively. Both the compounds 19 and 20 were also found active on UO-31 renal cancer cell line with % GI of 21.18 and 29.95, respectively. Compound 19 showed % GI of 19.99 on T-47D, while compound 20 showed % GI of 19.89 on MDA-MB-231/ATCC cell lines of breast cancer.

3. Discussion

Ten benzoxazole clubbed 2-pyrrolidinone derivatives (1120) as the inhibitors of monoacylglycerol lipase were designed on the criteria fulfilling the structural requirements and on the basis of previously reported inhibitors [36,42,43,44,45]. The designed, synthesized, and characterized compounds (1120) were screened against monoacylglycerol lipase (MAGL) in order to find potential inhibitors. The substituted phenyl derivatives (1320) were established to reduce the MAGL activity at 100 μM concentration below 50%. Compound 19 (4-NO2 derivative) and compound 20 (4-SO2NH2 derivative) were the most potent, with IC50 of 8.4 and 7.6 nM, correspondingly. The benzoxazole derivatives having 4-NO2 phenyl (19) and 4-SO2NH2 phenyl (20), with an FAAH IC50 value greater than 50 µM, were considered selective MAGL inhibitors. In molecular docking studies, compounds 19 and 20 showed comparable docking scores of −9.87 and −9.83, respectively. The binding of compounds 19 and 20 in the active site of MAGL revealed that the carbonyl group of pyrrolidinone is located exactly in the oxyanion hole and stabilized by three hydrogen bonds (~2 Å) with alanine 51, serine 122, and methionine 123. Serine 122 is one of the critical amino acid residues of the catalytic triad of MAGL. The benzoxazole moiety is found to be positioned in the amphiphilic pouch, having π-π stacking contacts with the amino acid tyrosine 194. The 4-NO2 phenyl (19) and 4-SO2NH2 phenyl (20) part of the ligand was engaged in hydrophobic (Van der Waals) attractions with the amino acids leucine 148, 213, and 241. The binding patterns of compounds 19 and 20 in the catalytic site of MAGL were found to be similar as those of the reported inhibitors bound crystal structures [32,36,39,41]. Remarkably, the physiochemical and pharmacokinetic properties of compounds 19 and 20 computed by QikProp were found to be in the qualifying range as per the proposed guideline for good orally bioactive CNS drugs. Moreover, compound 20 showed better toxicity profile than compound 19, as predicted by admetSAR [50]. In formalin-induced analgesic test, compound 20 reduced the pain response significantly both in acute (stage I) and late (stage II) phases in a dose-dependent manner. It significantly demonstrated the reduction in pain response, having better potency than the positive control GBP, at the dose of 30 mg/kg. Moreover, in one dose (10 µM), anticancer screening by SRB assay, compounds 19 (NSC: 778839) and 20 (NSC: 778842) were found to have good anticancer activity towards SNB-75 cell line of CNS cancer, having % growth inhibition (% GI) of 35.49 and 31.88, respectively. Therefore, the present work concluded that compound 20 is the potential lead compounds that can be further manipulated at points 1 and 4 of the 2-pyrrolidinone moiety for the discovery and development of more selective and potent inhibitors of MAGL for neuropathic pain and CNS disorders including cancers.

4. Experimental

4.1. Chemistry

Reagents and solvents were procured from Merck Ltd. (New Delhi, India) and Sigma-Aldrich Ltd. (New Delhi, India). Progress and completion of the reactions was checked by thin-layer chromatography (TLC). Melting points of the derivatives were determined by open tube capillary method and uncorrected. Elemental analysis data were obtained from CHNOS elemental analyzer (Vario EL III, Elementar Analysensysteme GmbH, Langenselbold, Germany). Shimadzu FT-IR spectrometer (Shimadzu Analytical Pvt. Ltd., New Delhi, India) was used for recording IR spectrum (4000–400 cm−1), by preparing KBr pellets. 1H-NMR spectrum of the derivatives were obtained from Bruker 300 MHz NMR instrument (Bruker Avance AV-III type, Billerica, MA, USA) using CDCl3 or DMSO-d6 as solvent. 1H-NMR spectra of compounds 1120 can be found in the Supplementary Material. Molecular mass (m/z) of the derivatives were obtained by UPLC-MS (Q-TOF-ESI) (Waters Corp., Milford, MA, USA).

4.1.1. Synthesis of 1-(Aryl Substituted)-5-Oxopyrrolidine-3-Carboxylic Acids (110)

Method-1 (for compound 1)

Equimolar amount of benzylamine (50 mmol) and itaconic acid (50 mmol, 6.5 g) in 50 mL of tripled distilled water was refluxed for about 45–60 min. The contents were then chilled, filtered, and washed with cold water. The obtained solid was dissolved in minimum quantity of aq. NaOH (10%). After treatment with activated charcoal, the solution was filtered and acidified with dil. HCl in order to obtain the precipitate. The filtered solid was washed with cold water, dried, and purified by recrystallization from ethanol/water mixture.
1-Benzyl-5-Oxopyrrolidine-3-Carboxylic Acid (1), white solid; yield: 75%; m.p. 142–145° C; IR: 1518 (C=C), 1627 (C=OOH), 1734 (C=O), 2947 (sp3 C-H), 3045 (Ar C-H), 3241 (COO-H); 1H-NMR (DMSO-d6) δ (ppm): 2.63–2.76 (m, 2H, COCH2), 3.62 (s, 2H, CH2), 4.29–4.44 (m, 3H, NCH2 and CHpyrr), 7.20–7.36 (m, 5H, Ar-H), 11.35 (s, 1H, COOH, D2O exchangeable); ESI-MS (m/z): 219.12 [M]+; Anal. calcd. For C12H13NO3: C, 65.74; H, 5.98; N, 6.39. Found: C, 65.80; H, 5.85; N, 6.50.

Method-2 (for compounds 2–10)

Intermediate compounds (2–10) were synthesized as per the procedure reported in our previous publication [45].

4.1.2. Synthesis of 4-(Benzoxazolyl)-1-(Aryl Substituted)Pyrrolidin-2-Ones (1120)

Appropriate 1-(aryl substituted)-5-oxopyrrolidine-3-carboxylic acids (20 mmol), 2-aminophenol (20 mmol, 2.18 g) and polyphosphoric acid (20 g) in an RBF were heated to 150–160° C and stirred for 2–3 h. The content of the RBF (round bottom flask) was cooled at RT; 5% NaCO3 (25 mL) was added and heated for 10 min. The content was chilled and transferred in a flask having 100 mL of water, and stirred at RT for 15 min. The filtered solid was washed three times with water (50 mL), dried, and purified by recrystallization with ethanol. The derivatives were then purified by chromatography using ethylacetate:hexane (1:4) as solvent.
4-(Benzo[d]oxazol-2-yl)-1-benzylpyrrolidin-2-one (11), pale-yellow solid; yield: 57%; m.p. 160–162 °C; IR: 1370 (C-N), 1555 (C=C), 1621 (C=N), 1703 (C=O), 3077 (Ar C-H); 1H-NMR (DMSO-d6) δ (ppm): 2.60–2.77 (m, 2H, COCH2), 3.13–3.18 (m, 1H, CHpyrr), 3.81–3.86 (m, 2H, NCH2), 3.95–4.01 (m, 2H, CH2), 7.38–7.58 (m, 6H, Ar-H), 7.73–7.78 (d, 1H, Ar-H, J = 15.6 Hz), 7.94–7.96 (d, 2H, Ar-H, J = 8.4 Hz); ESI-MS (m/z): 292.14 [M]+, 293.14 [M + H]+; Anal. calcd. for C18H16N2O2: C, 73.95; H, 5.52; N, 9.58. Found: C, 74.16; H, 5.70; N, 9.82.
4-(Benzo[d]oxazol-2-yl)-1-phenylpyrrolidin-2-one (12), pale-yellow solid; yield: 67%; m.p. 148–150 °C; IR: 1395 (C-N), 1552 (C=C), 1615 (C=N), 1698 (C=O), 3070 (Ar C-H); 1H-NMR (DMSO-d6) δ (ppm): 2.86–3.20 (m, 2H, COCH2), 4.16–4.40 (m, 3H, NCH2 and CHpyrr), 7.16–8.18 (m, 9H, Ar-H); ESI-MS (m/z): 278.13 [M]+, 279.13 [M + H]+; Anal. calcd. for C17H14N2O2: C, 73.37; H, 5.07; N, 10.07. Found: C, 73.60; H, 5.18; N, 10.33.
4-(Benzo[d]oxazol-2-yl)-1-(o-tolyl)pyrrolidin-2-one (13), pale-yellow solid; yield: 68%; m.p. 162–164 °C; IR: 1404 (C-N), 1568 (C=C), 1607 (C=N), 1691 (C=O), 3079 (Ar C-H); 1H-NMR (DMSO-d6) δ (ppm): 2.18 (s, 3H, CH3), 2.88–3.17 (m, 2H, COCH2), 4.12–4.18 (m, 1H, CHpyrr), 4.26–4.39 (m, 2H, NCH2), 7.18–7.21 (m, 2H, H-5phenyl and H-6phenyl), 7.36–7.54 (m, 4H, H-5benzoxazole, H-6benzoxazole, H-3phenyl and H-4phenyl), 7.98–8.01 (d, 1H, H-7benzoxazole, J = 8.1 Hz), 8.09–8.12 (d, 1H, H-4benzoxazole, J = 7.8 Hz); ESI-MS (m/z): 292.13 [M]+, 293.13 [M + H]+; Anal. calcd. for C18H16N2O2: C, 73.95; H, 5.52; N, 9.58. Found: C, 74.21; H, 5.82; N, 9.77.
4-(Benzo[d]oxazol-2-yl)-1-(p-tolyl)pyrrolidin-2-one (14), pale-yellow solid; yield: 67%; m.p. 164–166 °C; IR: 1411 (C-N), 1562 (C=C), 1598 (C=N), 1697 (C=O), 3077 (Ar C-H); 1H-NMR (DMSO-d6) δ (ppm): 2.28 (s, 3H, CH3), 2.94–3.13 (m, 2H, COCH2), 4.13–4.35 (m, 3H, NCH2 and CHpyrr), 7.17–7.20 (d, 2H, H-3phenyl and H-5phenyl, J = 8.4 Hz), 7.34–7.42 (m, 2H, H-2phenyl and H-6phenyl), 7.54–7.57 (d, 2H, H-4benzoxazole and H-7benzoxazole, J = 8.4 Hz), 7.70–7.74 (m, 2H, H-5benzoxazole and H-6benzoxazole); ESI-MS (m/z): 292.13 [M]+, 293.13 [M + H]+; Anal. calcd. for C18H16N2O2: C, 73.95; H, 5.52; N, 9.58. Found: C, 74.18; H, 5.78; N, 9.80.
4-(Benzo[d]oxazol-2-yl)-1-(4-chlorophenyl)pyrrolidin-2-one (15), pale-yellow solid; yield: 70%; m.p. 174–176 °C; IR: 758 (C-Cl), 1396 (C-N), 1568 (C=C), 1612 (C=N), 1689 (C=O), 3086 (Ar C-H); 1H-NMR (DMSO-d6) δ (ppm): 3.01–3.09 (m, 2H, COCH2), 4.20–4.33 (m, 3H, NCH2 and CHpyrr), 7.35–7.43 (m, 4H, Ar-H), 7.68–7.72 (t, 4H, Ar-H, J = 6.3 Hz); ESI-MS (m/z): 312.11 [M]+, 314.10 [M + 2]+; Anal. calcd. for C17H13ClN2O2: C, 65.29; H, 4.19; N, 8.96. Found: C, 65.52; H, 4.35; N, 9.03.
4-(Benzo[d]oxazol-2-yl)-1-(3-chloro-4-fluorophenyl)pyrrolidin-2-one (16), pale-yellow solid; yield: 68%; m.p. 187–189 °C; IR: 765 (C-Cl), 1388 (C-N), 1545 (C=C), 1613 (C=N), 1690 (C=O), 3098 (Ar C-H); 1H-NMR (DMSO-d6) δ (ppm): 2.92–3.19 (m, 2H, COCH2), 4.13–4.21 (m, 1H, CHpyrr), 4.25–4.41 (m, 2H, NCH2), 7.41–7.60 (m, 5H, Ar-H), 8.05–8.08 (d, 1H, H-7benzoxazole, J = 8.1 Hz), 8.11–8.14 (d, 1H, H-4benzoxazole, J = 8.1 Hz); ESI-MS (m/z): 330.10 [M]+, 332.09 [M + 2]+; Anal. calcd. for C17H12ClFN2O2: C, 61.73; H, 3.66; N, 8.47. Found: C, 62.07; H, 3.85; N, 8.61.
4-(Benzo[d]oxazol-2-yl)-1-(4-hydroxyphenyl)pyrrolidin-2-one (17), pale-yellow solid; yield: 62%; m.p. 242–245 °C; IR: 1407 (C-N), 1535 (C=C), 1597 (C=N), 1698 (C=O), 3092 (Ar C-H), 3410 (O-H); 1H-NMR (DMSO-d6) δ (ppm): 2.91–3.20 (m, 2H, COCH2), 4.12–4.23 (m, 1H, CHpyrr), 4.25–4.39 (m, 2H, NCH2), 5.64 (bs, 1H, OH, D2O exchangeable), 7.41–7.60 (m, 4H, Ar-Hphenyl), 7.72–7.77 (t, 2H, H-5benzoxazole and H-6benzoxazole, J = 7.5 Hz), 7.96–7.99 (d, 1H, H-7benzoxazole, J = 8.4 Hz), 8.08–8.10 (d, 1H, H-4benzoxazole, J = 8.4 Hz); ESI-MS (m/z): 294.12 [M]+, 295.12 [M + H]+; Anal. calcd. for C17H14N2O3: C, 69.38; H, 4.79; N, 9.52. Found: C, 69.66; H, 4.97; N, 9.78.
4-(Benzo[d]oxazol-2-yl)-1-(4-methoxyphenyl)pyrrolidin-2-one (18), pale-yellow solid; yield: 66%; m.p. 177–179 °C; IR: 1412 (C-N), 1546 (C=C), 1600 (C=N), 1690 (C=O), 3081 (Ar C-H); 1H-NMR (DMSO-d6) δ (ppm): 2.88–3.18 (m, 2H, COCH2), 3.67 (s, 3H, OCH3), 4.08–4.16 (m, 1H, CHpyrr), 4.22–4.39 (m, 2H, NCH2), 7.16–7.19 (d, 2H, H-2phenyl and H-6phenyl, J = 8.4 Hz), 7.42–7.58 (m, 4H, H-5benzoxazole, H-6benzoxazole, H-3phenyl and H-5phenyl), 7.95–7.98 (d, 1H, H-7benzoxazole, J = 8.4 Hz), 8.09–8.12 (d, 1H, H-4benzoxazole, J = 8.4 Hz); ESI-MS (m/z): 308.13 [M]+, 309.13 [M + H]+; Anal. calcd. for C18H16N2O3: C, 70.12; H, 5.23; N, 9.09. Found: C, 70.35; H, 5.38; N, 9.18.
4-(Benzo[d]oxazol-2-yl)-1-(4-nitrophenyl)pyrrolidin-2-one (19), yellow solid; yield: 62%; m.p. 276–278 °C; IR: 1405 (C-N), 1550 (C=C), 1595 (C=N), 1695 (C=O), 3089 (Ar C-H); 1H-NMR (DMSO-d6) δ (ppm): 2.89–3.15 (m, 2H, COCH2), 4.11–4.16 (m, 1H, CHpyrr), 4.24–4.37 (m, 2H, NCH2), 6.79–6.82 (d, 2H, H-2phenyl and H-6phenyl, J = 9 Hz), 7.43–7.48 (t, 2H, H-5benzoxazole, H-6benzoxazole, J = 7.5 Hz), 8.01–8.18 (m, 4H, H-4benzoxazole, H-7benzoxazole, H-3phenyl and H-5phenyl); ESI-MS (m/z): 323.11 [M]+, 324.11 [M + H]+; Anal. calcd. for C17H13N3O4: C, 63.16; H, 4.05; N, 13.00. Found: C, 63.35; H, 4.27; N, 13.23.
4-(4-(Benzo[d]oxazol-2-yl)-2-oxopyrrolidin-1-yl)benzenesulfonamide (20), yellow solid; yield: 60%; m.p. 227–229 °C; IR: 1416 (C-N), 1564 (C=C), 1607 (C=N), 1704 (C=O), 3095 (Ar C-H); 1H-NMR (DMSO-d6) δ (ppm): 2.91–3.19 (m, 2H, COCH2), 4.15–4.22 (m, 1H, CHpyrr), 4.25–4.38 (m, 2H, NCH2), 5.67 (s, 2H, SO2NH2, D2O exchangeable), 7.42–7.59 (m, 4H, Ar-H), 8.09–8.16 (m, 4H, Ar-H); ESI-MS (m/z): 357.12 [M]+, 358.12 [M + H]+; Anal. calcd. for C17H15N3O4S: C, 57.13; H, 4.23; N, 11.76. Found: C, 57.36; H, 4.36; N, 11.95.

4.2. Human MAGL Assay

The screening of the synthesized compounds (1120) for their capability to reduce hMAGL activity was performed according to the information leaflet provided with Cayman’s assay kit (Cayman Chemical, Michigan, USA) by the reported method [29], as discussed in detail in our previous publication [45]. The results were compared with standard MAGL inhibitors, CAY10499 and JZL184, and are provided in Table 1.

4.3. Human FAAH Assay

The screening of the selected compounds (15, 16, 18, 19 and 20) for their potential to inhibit hFAAH was performed according to the information leaflet provided with Cayman’s assay kit (Cayman Chemical, Ann Arbor, Michigan, USA) by the reported method [47], as discussed in detail in our previous publication [45]. The results were compared with standard FAAH inhibitor, URB597, and are provided in Table 1.

4.4. Molecular Docking Study

Glide executed on Maestro 9.4 (Schrödinger Inc., New York, NY, USA) was utilized for Glide XP docking of active compounds (19 and 20). The .pdb file of hMAGL X-ray crystal structure was downloaded from protein data bank having ID 5ZUN (crystal structure resolution 1.35 Å) for molecular docking study [36]. The protein structure was refined, optimized, and energy-minimized with the help of preparation wizard in Maestro. A docking grid of 20 × 20 × 20 Å, was created around the catalytic site by defining the cocrystallized ligand. Ligand (compounds 19 and 20) structures were prepared with the help of LigPrep 2.6 with Epik 2.4 at pH 7.0 ± 2.0. The methodology was validated by docking the cocrystallized ligand with Glide XP docking protocol [55].

4.5. Physicochemical and Pharmacokinetic Characteristics

Guidelines, concerning the validation and optimization of orally active CNS compounds, were developed by Ghose et. al. by analyzing 35 characteristic features of orally bioavailable 317 CNS and 626 non-CNS drugs [49]. For computations of these properties of the selected compounds (19 and 20), QikProp 3.6 module of Schrodinger was utilized. The generated data was then matched with the qualifying range as per the suggested guideline for good orally bioactive CNS drugs.

4.6. In Silico Absorption and Toxicity Profile

The selected compounds (19 and 20) were evaluated for their absorption and toxicity profile by a bioinformatics tool admetSAR [50]. Oral bioavailability, intestinal absorption, and BBB penetration properties were calculated. AMES test for mutagenicity, carcinogenicity test, and the calculation of LD50 for both the compounds (19 and 20) were also evaluated.

4.7. Analgesic Activity

Formalin-induced analgesic test was executed by the procedure described by Coderre and Laughlin [51,52] as discussed in detail in our previous publication [45]. Male Wistar rats (180–200 g) were obtained with the permission of IAEC (proposal number 1048) from Jamia Hamdard, New Delhi, India. The results of test compound 20 and reference drug, Gabapentin (GBP), were statistically compared with the control group.

4.8. Anticancer Screening: Sulforhodamine B Assay

Compounds 19 and 20 were supplied to National Cancer Institute (Bethesda, Maryland, USA), for in vitro sulforhodamine B (SRB) assay, anticancer screening on 60 cell lines of cancers of leukemia, melanoma, and tumors of the kidney, brain, breast, lung, colon, ovary, and prostate, as per their standard protocol [53,54]. One dose anticancer results (NCI, USA) of compounds 19 and 20 are provided in the Supplementary Materials.

4.9. Statistical Analysis

The statistical study of the data was accomplished by GraphPad Prism (version 8.0.2; GraphPad Software, San Diego, CA, USA). The dose response of the test compounds was compared with that of control, in formalin-induced analgesic test, by analysis of variance (ANOVA) followed by Dunnett’s test. Outcomes are communicated as mean ± SEM.

5. Conclusions

Ten benzoxazole clubbed 2-pyrrolidinone derivatives (1120) as the inhibitors of MAGL were designed, synthesized, characterized, and assayed against MAGL and FAAH enzymes, in order to find potential small molecule selective MAGL inhibitor. Compounds 19 (4-NO2 derivative) and 20 (4-SO2NH2 derivative) were found most potent and selective, with an IC50 of 8.4 and 7.6 nM, respectively. The binding patterns of compounds 19 and 20 in the catalytic site of MAGL were found as expected and similar as those of the reported inhibitors.
The physiochemical and pharmacokinetic properties of compounds 19 and 20 were found to be almost in the qualifying range as per the proposed guideline for good orally bioactive CNS drugs. Compound 20 significantly demonstrated the reduction in pain response, having better potency than the positive control GBP, at the dose of 30 mg/kg. The present work concluded that compound 20 is the potential lead compounds that can be further studied and optimized at points 1 and 4 of the 2-pyrrolidinone moiety for the discovery and development of more selective and potent inhibitors of MAGL for neuropathic pain.

Supplementary Materials

The following are available online, 1H-NMR spectra of compounds 1120 and one dose anticancer results (NCI, USA) of compounds 19 and 20 is provided in the supporting information.

Author Contributions

Conceptualization, O.A., A.S.A.A. and H.K.S.; data curation, O.A., M.Q.H. and M.M.S.; formal analysis, A.S.A.A., Y.R., M.Q.H. and H.K.S.; funding acquisition, A.S.A.A. and O.A.; investigation, O.A., M.Q.H. and M.M.S.; methodology, O.A. and M.Q.H.; project administration, A.S.A.A., O.A. and Y.R.; resources, A.S.A.A., O.A., M.Q.H. and H.K.S.; software, O.A., M.Q.H. and H.K.S.; supervision, O.A., A.S.A.A. and M.Q.H.; validation, A.S.A.A., Y.R., M.Q.H., M.M.S. and H.K.S.; visualization, Y.R., M.Q.H., M.M.S. and H.K.S.; writing—original draft, O.A.; writing—review and editing, M.Q.H., Y.R., M.M.S. and H.K.S. All authors have read and agreed to the published version of the manuscript.

Funding

Authors would like to thank the Deanship of Scientific Research at Prince Sattam Bin Abdulaziz University for providing financial support for this project (grant no. 2019/03/11064).

Institutional Review Board Statement

The study was conducted according to the guidelines of the Declaration of Helsinki, and approved by the Institutional Animal Ethics Committee (Proposal number 1048) of Jamia Hamdard (Hamdard University, New Delhi-110062 (Registration Number: 173/CPCSEA; 28 January 2000).

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Acknowledgments

Authors would like to thank the Deanship of Scientific Research at Prince Sattam Bin Abdulaziz University for providing financial support for this project (grant no. 2019/03/11064).

Conflicts of Interest

Authors declare that they do not have any conflict of interest.

Sample Availability

The samples of the compounds are available on request from the corresponding author.

References

  1. Di Marzo, V.; Bisogno, T.; De Petrocellis, L. the biosynthesis, fate and pharmacological properties of endocannabinoids. Handb. Exp. Pharmacol. 2005, 168, 147–185. [Google Scholar] [CrossRef]
  2. Sugiura, T.; Kondo, S.; Sukagawa, A.; Nakane, S.; Shinoda, A.; Itoh, K. 2- Arachidonoylglycerol: A possible endogenous can-nabinoid receptor ligand in brain. Biochem. Biophys. Res. Commun. 1995, 215, 89–97. [Google Scholar] [CrossRef] [PubMed]
  3. Matias, I.; Endocannabinoid Research Group; Bisogno, T.; Di Marzo, V. Endogenous cannabinoids in the brain and peripheral tissues: Regulation of their levels and control of food intake. Int. J. Obes. 2006, 30, S7–S12. [Google Scholar] [CrossRef] [Green Version]
  4. Blankman, J.L.; Simon, G.M.; Cravatt, B.F. A comprehensive profile of brain enzymes that hydrolyze the endocannabinoid 2-arachidonoylglycerol. Chem. Biol. 2007, 14, 1347–1356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Savinainen, J.R.; Saario, S.M.; Laitinen, J.T. The serine hydrolases MAGL, ABHD6 and ABHD12 as guardians of 2-arachidonoylglycerol signalling through cannabinoid receptors. Acta Physiol. 2011, 204, 267–276. [Google Scholar] [CrossRef]
  6. Kinsey, S.G.; Long, J.Z.; O’Neal, S.T.; Abdullah, R.A.; Poklis, J.L.; Boger, D.; Cravatt, B.F.; Lichtman, A.H. Blockade of endo-cannabinoid-degrading enzymes attenuates neuropathic pain. J. Pharmacol. Exp. Ther. 2009, 330, 902–910. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Kinsey, S.G.; Long, J.Z.; Cravatt, B.F.; Lichtman, A.H. Fatty acid amide hydrolase and monoacylglycerol lipase inhibitors produce anti-allodynic effects in mice through distinct cannabinoid receptor mechanisms. J. Pain 2010, 11, 1420–1428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Hohmann, A.G.; Suplita, R.L.; Bolton, N.M.; Neely, M.H.; Fegley, D.; Mangieri, R.; Krey, J.F.; Walker, J.M.; Holmes, P.V.; Crystal, J.D.; et al. An endocannabinoid mechanism for stress-induced analgesia. Nat. Cell Biol. 2005, 435, 1108–1112. [Google Scholar] [CrossRef] [PubMed]
  9. Connell, K.; Bolton, N.; Olsen, D.; Piomelli, D.; Hohmann, A.G. Role of the basolateral nucleus of the amygdala in endocan-nabinoid-mediated stress-induced analgesia. Neurosci. Lett. 2006, 397, 180–184. [Google Scholar] [CrossRef] [Green Version]
  10. Guindon, J.; Guijarro, A.; Piomelli, D.; Hohmann, A.G. Peripheral antinociceptive effects of inhibitors of monoacylglycerol lipase in a rat model of inflammatory pain. Br. J. Pharmacol. 2010, 163, 1464–1478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Guindon, J.; Lai, Y.; Takacs, S.M.; Bradshaw, H.B.; Hohmann, A.G. Alterations in endocannabinoid tone following chemotherapy-induced peripheral neuropathy: Effects of endocannabinoid deactivation inhibitors targeting fatty-acid amide hydrolase and monoacylglycerol lipase in comparison to reference analgesics following cisplatin treatment. Pharmacol. Res. 2013, 67, 94–109. [Google Scholar] [PubMed] [Green Version]
  12. Aaltonen, N.; Kedzierska, E.; Orzelska-Górka, J.; Lehtonen, M.; Navia-Paldanius, D.; Jakupovic, H.; Savinainen, J.R.; Neva-lainen, T.; Laitinen, J.T.; Parkkari, T.; et al. In vivo characterization of the ultrapotent monoacylglycerol lipase inhibitor {4-[bis-(benzo[d][1,3]dioxol-5-yl)methyl]-piperidin-1-yl}(1H-1,2,4-triazol-1-yl)methanone (JJKK-048). J. Pharm. Exp. Ther. 2016, 359, 62–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Nomura, D.K.; Lombardi, D.P.; Chang, J.W.; Niessen, S.; Ward, A.M.; Long, J.Z.; Hoover, H.H.; Cravatt, B.F. Monoacylglycerol lipase exerts dual control over endocannabinoid and fatty acid pathways to support prostate cancer. Chem. Biol. 2011, 18, 846–856. [Google Scholar] [CrossRef] [Green Version]
  14. Chen, R.; Zhang, J.; Wu, Y.; Wang, D.; Feng, G.; Tang, Y.-P.; Teng, Z.; Chen, C. Monoacylglycerol lipase is a therapeutic target for Alzheimer’s Disease. Cell Rep. 2012, 2, 1329–1339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. McAllister, L.A.; Butler, C.R.; Mente, S.; O’Neil, S.V.; Fonseca, K.R.; Piro, J.R.; Cianfrogna, J.A.; Foley, T.L.; Gilbert, A.M.; Harris, A.R.; et al. Discovery of trifluoromethyl glycol carbamates as potent and selective covalent monoacylglycerol lipase (MAGL) inhibitors for treatment of neuroinflammation. J. Med. Chem. 2018, 61, 3008–3026. [Google Scholar] [CrossRef] [PubMed]
  16. Cisar, J.S.; Weber, O.D.; Clapper, J.R.; Blankman, J.L.; Henry, C.L.; Simon, G.M.; Alexander, J.P.; Jones, T.K.; Ezekowitz, R.A.B.; O’Neill, G.P.; et al. Identification of ABX-1431, a selective inhibitor of monoacylglycerol lipase and clinical candidate for treatment of neurological disorders. J. Med. Chem. 2018, 61, 9062–9084. [Google Scholar] [CrossRef] [Green Version]
  17. Maione, S.; Morera, E.; Marabese, I.; Ligresti, A.; Luongo, L.; Ortar, G.; DiMarzo, V. Antinociceptive effects of tetrazole inhib-itors of endocannabinoid inactivation: Cannabinoid and noncannabinoid receptor-mediated mechanisms. Br. J. Pharmacol. 2008, 155, 775–782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Fowler, C.J. Monoacylglycerol lipase—A target for drug development? Br. J. Pharmacol. 2012, 166, 1568–1585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Mulvihill, M.M.; Nomura, D.K. Therapeutic potential of monoacylglycerol lipase inhibitors. Life Sci. 2013, 92, 492–497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Alhouayek, M.; Masquelier, J.; Muccioli, G.G. Controlling 2-arachidonoylglycerol metabolism as an anti-inflammatory strategy. Drug Discov. Today 2014, 19, 295–304. [Google Scholar] [CrossRef] [PubMed]
  21. Guzman, M. A new age for MAGL. Chem. Biol. 2010, 17, 4–6. [Google Scholar] [CrossRef] [Green Version]
  22. Long, J.Z.; Nomura, D.K.; Cravatt, B.F. Characterization of monoacylglycerol lipase inhibition reveals differences in central and peripheral endocannabinoid metabolism. Chem. Biol. 2009, 16, 744–753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Lass, A.; Zimmermann, R.; Oberer, M.; Zechner, R. Lipolysis—A highly regulated multi-enzyme complex mediates the catabolism of cellular fat stores. Prog. Lipid Res. 2011, 50, 14–27. [Google Scholar] [CrossRef] [Green Version]
  24. Dorsam, R.T.; Gutkind, J.S. G-protein-coupled receptors and cancer. Nat. Rev. Cancer 2007, 7, 79–94. [Google Scholar] [CrossRef] [PubMed]
  25. Nomura, D.K.; Long, J.Z.; Niessen, S.; Hoover, H.S.; Ng, S.-W.; Cravatt, B.F. Monoacylglycerol lipase regulates a fatty acid network that promotes cancer pathogenesis. Cell 2010, 140, 49–61. [Google Scholar] [CrossRef] [Green Version]
  26. Nomura, D.K.; Morrison, B.E.; Blankman, J.L.; Long, J.Z.; Kinsey, S.G.; Marcondes, M.C.G.; Ward, A.M.; Hahn, Y.K.; Lichtman, A.H.; Conti, B.; et al. Endocannabinoid hydrolysis generates brain prostaglandins that promote neuroinflammation. Science 2011, 334, 809–813. [Google Scholar] [CrossRef] [Green Version]
  27. Ye, L.; Zhang, B.; Seviour, E.G.; Tao, K.-X.; Liu, X.-H.; Ling, Y.; Chen, J.-Y.; Wang, G.-B. Monoacylglycerol lipase (MAGL) knockdown inhibits tumor cells growth in colorectal cancer. Cancer Lett. 2011, 307, 6–17. [Google Scholar] [CrossRef]
  28. King, A.R.; Duranti, A.; Tontini, A.; Rivara, S.; Rosengarth, A.; Clapper, J.R.; Astarita, G.; Geaga, J.A.; Luecke, H.; Mor, M.; et al. URB602 inhibits monoacylglycerol lipase and selectively blocks 2-arachidonoylglycerol degradation in intact brain slices. Chem. Biol. 2007, 14, 1357–1365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Muccioli, G.G.; LaBar, G.; Lambert, D.M. CAY10499, a novel monoglyceride lipase inhibitor evidenced by an expeditious MGL assay. ChemBioChem 2008, 9, 2704–2710. [Google Scholar] [CrossRef] [PubMed]
  30. Minkkilä, A.; Savinainen, J.R.; Käsnänen, H.; Xhaard, H.; Nevalainen, T.; Laitinen, J.T.; Poso, A.; Leppänen, J.; Saario, S.M. Screening of various hormone-sensitive lipase inhibitors as endocannabinoid-hydrolyzing enzyme inhibitors. ChemMedChem 2009, 4, 1253–1259. [Google Scholar] [CrossRef]
  31. Long, J.Z.; Jin, X.; Adibekian, A.; Li, W.; Cravatt, B.F. Characterization of tunable piperidine and piperazine carbamates as inhibitors of endocannabinoid hydrolases. J. Med. Chem. 2010, 53, 1830–1842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Bertrand, T.; Augé, F.; Houtmann, J.; Rak, A.; Vallée, F.; Mikol, V.; Berne, P.; Michot, N.; Cheuret, D.; Hoornaert, C.; et al. Structural basis for human monoglyceride lipase inhibition. J. Mol. Biol. 2010, 396, 663–673. [Google Scholar] [CrossRef]
  33. Chang, J.W.; Niphakis, M.J.; Lum, K.M.; Cognetta, A.B.; Wang, C.; Matthews, M.L.; Niessen, S.; Buczynski, M.W.; Parsons, L.H.; Cravatt, B.F. Remarkably selective inhibitors of monoacylglycerol lipase bearing a reactive group that is bioisosteric with endocannabinoid substrates. Chem. Biol. 2012, 19, 579–588. [Google Scholar] [CrossRef] [Green Version]
  34. Morera, L.; LaBar, G.; Ortar, G.; Lambert, D.M. Development and characterization of endocannabinoid hydrolases FAAH and MAGL inhibitors bearing a benzotriazol-1-yl carboxamide scaffold. Bioorg. Med. Chem. 2012, 20, 6260–6275. [Google Scholar] [CrossRef]
  35. Aaltonen, N.; Savinainen, J.R.; Ribas, C.R.; Rönkkö, J.; Kuusisto, A.; Korhonen, J.; Navia-Paldanius, D.; Häyrinen, J.; Takabe, P.; Käsnänen, H.; et al. Piperazine and piperidine triazole ureas as ultrapotent and highly selective inhibitors of monoacylglycerol lipase. Chem. Biol. 2013, 20, 379–390. [Google Scholar] [CrossRef] [Green Version]
  36. Aida, J.; Fushimi, M.; Kusumoto, T.; Sugiyama, H.; Arimura, N.; Ikeda, S.; Sasaki, M.; Sogabe, S.; Aoyama, K.; Koike, T. Design, synthesis, and evaluation of piperazinyl pyrrolidin-2-ones as a novel series of reversible monoacylglycerol lipase inhibitors. J. Med. Chem. 2018, 61, 9205–9217. [Google Scholar] [CrossRef]
  37. Granchi, C.; Lapillo, M.; Glasmacher, S.; Bononi, G.; Licari, C.; Poli, G.; El Boustani, M.; Caligiuri, I.; Rizzolio, F.; Gertsch, J.; et al. Optimization of a benzoylpiperidine class identifies a highly potent and selective reversible monoacylglycerol lipase (MAGL) inhibitor. J. Med. Chem. 2019, 62, 1932–1958. [Google Scholar] [CrossRef]
  38. Castelli, R.; Scalvini, L.; Vacondio, F.; Lodola, A.; Anselmi, M.; Vezzosi, S.; Carmi, C.; Bassi, M.; Ferlenghi, F.; Rivara, S. Benzisothiazolinone derivatives as potent allosteric monoacylglycerol lipase inhibitors that functionally mimic sulfenylation of regulatory cysteines. J. Med. Chem. 2020, 63, 1261–1280. [Google Scholar] [CrossRef]
  39. LaBar, G.; Bauvois, C.; Borel, F.; Ferrer, J.-L.; Wouters, J.; Lambert, D.M. Crystal structure of the human monoacylglycerol lipase, a key actor in endocannabinoid signaling. ChemBioChem 2009, 11, 218–227. [Google Scholar] [CrossRef]
  40. Schalk-Hihi, C.; Schubert, C.; Alexander, R.; Bayoumy, S.; Clemente, J.C.; Deckman, I.; DesJarlais, R.L.; Dzordzorme, K.C.; Flores, C.M.; Grasberger, B.; et al. Crystal structure of a soluble form of human monoglyceride lipase in complex with an inhibitor at 1.35 Å resolution. Protein Sci. 2011, 20, 670–683. [Google Scholar] [CrossRef] [Green Version]
  41. Scalvini, L.; Piomelli, D.; Mor, M. Monoglyceride lipase: Structure and inhibitors. Chem. Phys. Lipids 2016, 197, 13–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Afzal, O.; Kumar, S.; Kumar, R.; Firoz, A.; Jaggi, M.; Bawa, S. Docking based virtual screening and molecular dynamics study to identify potential monoacylglycerol lipase inhibitors. Bioorg. Med. Chem. Lett. 2014, 24, 3986–3996. [Google Scholar] [CrossRef]
  43. Afzal, O.; Akhtar, S.; Kumar, S.; Ali, R.; Jaggi, M.; Bawa, S. Hit to lead optimization of a series of N-[4-(1,3-benzothiazol-2-yl)phenyl]acetamides as monoacylglycerol lipase inhibitors with potential anticancer activity. Eur. J. Med. Chem. 2016, 121, 318–330. [Google Scholar] [CrossRef] [PubMed]
  44. Ali, M.R.; Kumar, S.; Shalmali, N.; Afzal, O.; Azim, S.; Chanana, D.; Alam, O.; Paudel, Y.N.; Sharma, M.; Bawa, S. Development of thiazole-5-carboxylate derivatives as selective inhibitors of monoacylglycerol lipase as target in cancer. Mini Rev. Med. Chem. 2019, 19, 410–423. [Google Scholar] [CrossRef]
  45. Altamimi, A.S.A.; Bawa, S.; Athar, F.; Hassan, Q.; Riadi, Y.; Afzal, O. Pyrrolidin-2-one linked benzofused heterocycles as novel small molecule monoacylglycerol lipase inhibitors and antinociceptive agents. Chem. Biol. Drug Des. 2020, 96, 1418–1432. [Google Scholar] [CrossRef]
  46. Mickevicius, M.; Beresnevicius, Z.J.; Mickevicius, V.; Mikulskiene, G. Condensation products of 1-aryl-4-carboxy-2-pyrroli-dinones with o-diaminoarenes, o-aminophenol, and their structural studies. Heteroat. Chem. 2006, 17, 47–56. [Google Scholar] [CrossRef]
  47. Granchi, C.; Rizzolio, F.; Palazzolo, S.; Carmignani, S.; Macchia, M.; Saccomanni, G.; Manera, C.; Martinelli, A.; Minutolo, F.; Tuccinardi, T. Structural optimization of 4-chlorobenzoylpiperidine derivatives for the development of potent, reversible, and selective monoacylglycerol lipase (MAGL) inhibitors. J. Med. Chem. 2016, 59, 10299–10314. [Google Scholar] [CrossRef]
  48. Mor, M.; Rivara, S.; Lodola, A.; Plazzi, P.V.; Tarzia, G.; Duranti, A.; Tontini, A.; Piersanti, G.; Kathuria, S.; Piomelli, D. Cy-clohexylcarbamic acid 3’- or 4’-substituted biphenyl-3-yl esters as fatty acid amide hydrolase inhibitors: Synthesis, quantitative structure-activity relationships, and molecular modeling studies. J. Med. Chem. 2004, 47, 4998–5008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Ghose, A.K.; Herbertz, T.; Hudkins, R.L.; Dorsey, B.D.; Mallamo, J.P. Knowledge-based, central nervous system (CNS) lead selection and lead optimization for CNS drug discovery. ACS Chem. Neurosci. 2012, 3, 50–68. [Google Scholar] [CrossRef] [Green Version]
  50. Yang, H.; Lou, C.; Sun, L.; Li, J.; Cai, Y.; Wang, Z.; Li, W.; Liu, G.; Tang, Y. admetSAR 2.0: Web-service for prediction and optimization of chemical ADMET properties. Bioinformatics 2018, 35, 1067–1069. [Google Scholar] [CrossRef]
  51. Coderre, T.J.; Vaccarino, A.L.; Melzack, R. Central nervous system plasticity in the tonic pain response to subcutaneous for-malin injection. Brain Res. 1990, 535, 155–158. [Google Scholar] [CrossRef]
  52. Laughlin, T.M.; Tram, K.V.; Wilcox, G.L.; Birnbaum, A.K. Comparison of antiepileptic drugs tiagabine, lamotrigine, and gabapentin in mouse models of acute, prolonged, and chronic nociception. J. Pharmacol. Exp. Ther. 2002, 302, 1168–1175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Grever, M.R.; A Schepartz, S.; A Chabner, B. The National Cancer Institute: Cancer drug discovery and development program. Semin. Oncol. 1992, 19, 622–638. [Google Scholar] [PubMed]
  54. Shoemaker, R.H. The NCI60 human tumour cell line anticancer drug screen. Nat. Rev. Cancer 2006, 6, 813–823. [Google Scholar] [CrossRef] [PubMed]
  55. Friesner, R.A.; Murphy, R.B.; Repasky, M.P.; Frye, L.L.; Greenwood, J.R.; Halgren, T.A.; Sanschagrin, P.C.; Mainz, D.T. Extra precision glide: Docking and scoring incorporating a model of hydrophobic enclosure for protein-ligand complexes. J. Med. Chem. 2006, 49, 6177–6196. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Role of MAGL inhibitors in the alleviation of pain and cancer. PG: prostaglandins; PA: phosphatidic acid; LPA: lysophosphatidic acid; and S1P: sphingosine-1-phosphate.
Figure 1. Role of MAGL inhibitors in the alleviation of pain and cancer. PG: prostaglandins; PA: phosphatidic acid; LPA: lysophosphatidic acid; and S1P: sphingosine-1-phosphate.
Molecules 26 02389 g001
Figure 2. Design of novel pyrrolidin-2-one linked benzoxazole MAGL inhibitors. (A) Binding pattern of reported pyrrolidin-2-one MAGL inhibitors; ZINC12863377 [42], compound 25 [45] and (R)-3t [36]. (B) Designed compounds (1120).
Figure 2. Design of novel pyrrolidin-2-one linked benzoxazole MAGL inhibitors. (A) Binding pattern of reported pyrrolidin-2-one MAGL inhibitors; ZINC12863377 [42], compound 25 [45] and (R)-3t [36]. (B) Designed compounds (1120).
Molecules 26 02389 g002
Figure 3. Scheme for synthesis of the intermediates (110) and target compounds (1120). (a) Reflux in water (Method 1); (b) Fusion at 130–140 °C (Method 2); (c) 2-aminophenol, polyphosphoric acid, 150–160 °C, sodium carbonate.
Figure 3. Scheme for synthesis of the intermediates (110) and target compounds (1120). (a) Reflux in water (Method 1); (b) Fusion at 130–140 °C (Method 2); (c) 2-aminophenol, polyphosphoric acid, 150–160 °C, sodium carbonate.
Molecules 26 02389 g003
Figure 4. Glide XP docking; 3D and 2D representation of the binding pattern of compound 19 (A), and 20 (B) in the catalytic site of MAGL. The picture (2D and 3D) for the docked complexes were obtained from Discovery studio visualizer 2020. The hydrophobicity was calculated from the default option available.
Figure 4. Glide XP docking; 3D and 2D representation of the binding pattern of compound 19 (A), and 20 (B) in the catalytic site of MAGL. The picture (2D and 3D) for the docked complexes were obtained from Discovery studio visualizer 2020. The hydrophobicity was calculated from the default option available.
Molecules 26 02389 g004
Figure 5. Formalin-induced analgesic test; a dose of the test compound 20 (5, 10, 30, and 50 mg/kg, p.o, suspended in 0.5% CMC) was administered 4 h before formalin injection (50 µL, 2.5%). Reference drug, gabapentin (100 mg/kg, i.p, dissolved in 0.9% normal saline) was injected 30 min before formalin injection. Total paw licking and biting duration, Stage I (white bar, 0–5 min) and stage II (black bar, 10–30 min) was recorded as a measure of pain behavior. Data is represented as mean ± SEM from a group of 10 animals. * p < 0.05, ** p < 0.01, p > 0.05 (NS, nonsignificant) vs. control (vehicle). GBP: Gabapentin; D5, D10, D30, andD50 are the dose concentrations of test compound 20.
Figure 5. Formalin-induced analgesic test; a dose of the test compound 20 (5, 10, 30, and 50 mg/kg, p.o, suspended in 0.5% CMC) was administered 4 h before formalin injection (50 µL, 2.5%). Reference drug, gabapentin (100 mg/kg, i.p, dissolved in 0.9% normal saline) was injected 30 min before formalin injection. Total paw licking and biting duration, Stage I (white bar, 0–5 min) and stage II (black bar, 10–30 min) was recorded as a measure of pain behavior. Data is represented as mean ± SEM from a group of 10 animals. * p < 0.05, ** p < 0.01, p > 0.05 (NS, nonsignificant) vs. control (vehicle). GBP: Gabapentin; D5, D10, D30, andD50 are the dose concentrations of test compound 20.
Molecules 26 02389 g005
Table 1. In vitro hMAGL and hFAAH inhibition assay of the synthesized compounds (1120).
Table 1. In vitro hMAGL and hFAAH inhibition assay of the synthesized compounds (1120).
Molecules 26 02389 i001
CompoundRhMAGL; IC50hFAAH; IC50
11 Molecules 26 02389 i002>100 μMND
12 Molecules 26 02389 i003>100 μMND
13 Molecules 26 02389 i00485 ± 1.5 μMND
14 Molecules 26 02389 i00572 ± 2.1 μMND
15 Molecules 26 02389 i00665 ± 3.2 nM28 ± 1.8 μM
16 Molecules 26 02389 i00734 ± 1.7 nM25 ± 2.3 μM
17 Molecules 26 02389 i00862 ± 2.7 μMND
18 Molecules 26 02389 i00942 ± 1.5 nM37 ± 2.2 μM
19 Molecules 26 02389 i0108.4 ± 1.9 nM55 ± 2.7 μM
20 Molecules 26 02389 i0117.6 ± 0.8 nM68 ± 2.1 μM
CAY10499--415 ± 3.2 nM--
JZL184--10 ± 0.8 nM--
URB597----5 ± 0.6 nM
ND: Not determined; CAY10499, JZL184 and URB597 (standard control); IC50 values were calculated from GraphPad Prism (ver. 8.0.2). Results are expressed as mean ± SEM (n = 3).
Table 2. Physicochemical and pharmacokinetic properties of compounds 19 and 20, predicted by QikProp, Schrodinger, for CNS activity.
Table 2. Physicochemical and pharmacokinetic properties of compounds 19 and 20, predicted by QikProp, Schrodinger, for CNS activity.
S. No.PropertyDescriptionRange of Properties in CNS DrugsCompound 19Compound 20
QLPLPUQU
1#starsdrug likeness penalty; the higher the value, the less drug-like the molecule000300
2#amineno. of basic amines011200
3#amidineno. of amidines groups000000
4#acidno. of carboxylic acid groups000000
5#amideno. of amides groups000100
6#rotorno. of rotatable bonds (without CX3, alkene, amide, small ring) 036812
7CNSa qualitative CNS activity parameter−2012−2−2
8dipolecomputed dipole moment 0.671.13.98.99.4710.22
9SASAsolvent accessible surface area348487620798584.85617.51
10FOSASASA on saturated carbon and attached hydrogen 1617831446491.9991.96
11FISASASA on N, O, and H attached to heteroatoms0064176167.04210.06
12PISAπ component of SASA0160292343325.81313.57
13WPSAweakly polar component of the SASA (halogens, P, and S)00012601.94
14volumesolvent accessible volume (Å3)492830110413881002.521065.46
15donorHBestimated no. of hydrogen bonds that would be donated to the solvent water001302
16accptHBestimated no. of hydrogen bonds that would be accepted from the solvent water12.85.28.369.5
17globa globularity descriptor (1 for a sphere)0.770.820.880.930.820.81
18QPpolrzpredicted polarizability (Å3)1428384936.4338.19
19QPlogPo/woctanol−water logP−0.162.54.76.02.130.90
20QPlogSsolubility in log(moles/liter)−6.5−4.6−2.5−0.42−3.99−3.99
21CIQPlogSlog of conformation-independent solubility−6.3−4.2−2.30.36−4.16−3.77
22QPPCacoapparent Caco-2 cell permeability008103269258.09100.92
23QPlogBBbrain/blood partition coefficient−1.2−0.060.751.2−1.12−1.65
24QPPMDCKpredicted apparent MDCK cell permeability (nm/s)006345899114.4342.50
25QPlogKhsaprediction of binding to human serum albumin−10.040.781.04−0.11−0.34
26HumanOralAbsorptionHuman oral absorption233333
27PercentHuman
OralAbsorption
Percent of human oral absorption619510010082.6368.11
28TPSAvan der Waals surface area of polar nitrogen and oxygen atoms3.8125410998.65119.08
29#NandOno. of N and O atoms124777
30RuleOfFiveno. of violations of Lipinski’s rule of five000100
31RuleOfThreeno. of violations of Jorgensen’s rule of three000100
32#in34no. of atoms in three- or four-membered rings000000
33#in56no. of atoms in five- or six-membered rings51117242020
34#nonconno. of atoms not able to form conjugation in nonaromatic rings0041033
35#nonHatmno. of non-H atoms81925302425
Abbreviations: QL, qualifying lower limit; PL, preferred lower limit; QU, qualifying upper limit; PU, preferred upper limit. # QL, PL, QU and PU values for CNS drug criteria were obtained from reference [49].
Table 3. In-silico absorption and toxicity profile of compounds 19 and 20 obtained from admetSAR server [50].
Table 3. In-silico absorption and toxicity profile of compounds 19 and 20 obtained from admetSAR server [50].
CompoundBBBHIAHOBAMES testCarcinogenicityRat Acute Toxicity
(LD50, mol/kg)
19YesYesYes MutagenicNon-carcinogen2.30
20YesYesYes Non-MutagenicNon-carcinogen2.21
BBB: blood–brain barrier; HIA: human intestinal permeability; HOB: human oral bioavailability; AMES test is to detect a probable mutagen; carcinogenicity estimates the cancer causing ability of a molecule; LD50: lethal dose which could kill 50% of the population of the organism (rat) on which it is being tested.
Table 4. In vitro anticancer screening of compound 19 and 20, against NCI60 cell lines at 10 μM concentration.
Table 4. In vitro anticancer screening of compound 19 and 20, against NCI60 cell lines at 10 μM concentration.
PanelCell LineCompound 19
(NSC: 778839)
Compound 20
(NSC: 778842)
% G% GI% G% GI
LeukemiaCCRF-CEM93.586.4294.485.52
HL-60(TB)100.09−0.0996.153.85
K-56298.411.5998.831.17
MOLT-493.646.3692.477.53
RPMI-8226101.28−1.28103.53−3.53
SR88.9211.0893.206.80
Non-Small Cell Lung CancerA549/ATCC100.60−0.6095.284.72
HOP-6285.0414.9681.9718.03
HOP-92104.72−4.7277.7822.22
NCI-H22698.631.3792.757.25
NCI-H2393.076.9392.197.81
NCI-H322M94.185.8299.100.90
NCI-H460102.33−2.33104.20−4.20
Colon CancerCOLO 205103.79−3.79104.05−4.05
HCC-2998102.05−2.05100.24−0.24
HCT-116102.14−2.1495.78−4.22
HCT-1598.101.9101.16−1.16
HT2999.250.75103.23−3.23
KM12105.43−5.43101.16−1.16
SW-620102.52−2.52102.95−2.95
CNS CancerSF-26891.858.1587.4312.57
SF-29598.211.7993.716.29
SF-53995.594.4187.5812.42
SNB-1999.290.7197.022.98
SNB-7564.5135.4968.1231.88
U251100.97−0.9795.214.79
MelanomaLOX IMVI89.0610.9492.997.01
MALME-3M88.5311.4793.296.71
M14101.37−1.3798.581.42
MDA-MB-43595.054.95100.55−0.55
SK-MEL-2102.31−2.31111.54−11.54
SK-MEL-28111.25−11.25101.79−1.79
SK-MEL-598.721.2898.82−1.18
UACC-257106.92−6.92110.78−10.78
UACC-6297.592.4192.787.22
Ovarian CancerIGROV1104.14−4.14101.63−1.63
OVCAR-398.561.4498.531.47
OVCAR-4106.27−6.2799.470.53
OVCAR-598.301.7092.087.92
OVCAR-8101.95−1.9597.492.51
NCI/ADR-RES98.371.63101.45−1.45
SK-OV-388.3411.6694.775.23
Renal Cancer786-0104.06−4.0698.991.01
A498113.46−13.46113.94−13.94
ACHN91.708.389.4810.52
CAKI-197.192.8192.357.65
SN12C97.212.7995.504.50
TK-10110.14−10.14114.82−14.82
UO-3178.8221.1870.0529.95
Prostate CancerPC-391.148.8688.1111.89
DU-145110.09−10.09111.62−11.62
Breast CancerMCF799.190.8192.327.68
MDA-MB-231/ATCC88.4111.5980.1119.89
HS 578T101.81−1.81104.19−4.19
T-47D80.0119.9983.2416.24
MDA-MB-46898.031.97100.34−0.34
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Afzal, O.; Altamimi, A.S.A.; Shahroz, M.M.; Sharma, H.K.; Riadi, Y.; Hassan, M.Q. Analgesic and Anticancer Activity of Benzoxazole Clubbed 2-Pyrrolidinones as Novel Inhibitors of Monoacylglycerol Lipase. Molecules 2021, 26, 2389. https://doi.org/10.3390/molecules26082389

AMA Style

Afzal O, Altamimi ASA, Shahroz MM, Sharma HK, Riadi Y, Hassan MQ. Analgesic and Anticancer Activity of Benzoxazole Clubbed 2-Pyrrolidinones as Novel Inhibitors of Monoacylglycerol Lipase. Molecules. 2021; 26(8):2389. https://doi.org/10.3390/molecules26082389

Chicago/Turabian Style

Afzal, Obaid, Abdulmalik Saleh Alfawaz Altamimi, Mir Mohammad Shahroz, Hemant Kumar Sharma, Yassine Riadi, and Md Quamrul Hassan. 2021. "Analgesic and Anticancer Activity of Benzoxazole Clubbed 2-Pyrrolidinones as Novel Inhibitors of Monoacylglycerol Lipase" Molecules 26, no. 8: 2389. https://doi.org/10.3390/molecules26082389

Article Metrics

Back to TopTop