HOME JOURNALS CONTACT

International Journal of Pharmacology

Year: 2022 | Volume: 18 | Issue: 2 | Page No.: 262-278
DOI: 10.3923/ijp.2022.262.278
Evaluation of Colorectal Cancer Inhibition Ability of Rosmarinus officinalis L. via Molecular Docking and Pharmacophore Analysis
Thanh-Diem Nguyen, Ly Le, Tu-Binh Vo, Kim-Lan Vo, Hoang-Minh Le and Huyen-Trang Vu

Abstract: Background and Objective: Colorectal cancer is one of the most common cancers in the world. Mutated proteins of certain genes that control cell apoptosis have been identified as the cause of colorectal cancer. Natural compounds that interact and denature these proteins can be used to inhibit the activities of these proteins and help prevent tumour growth with limited side effects. However, searching for such new compounds through in vitro or in vivo tests is time-consuming and costly. Materials and Methods: In this study, 30 known compounds from the herbal plant Rosmarinus officinalis L. were used to study the inhibitory ability of certain types of colorectal cancer-causing proteins using the drug design simulation method. Due to the computer-based drug design simulation method, target disease-causing proteins can be simulated to interact with a variety of compounds from herbal medicinal plants to detect compounds with high affinity and low energy required for interaction. Following that, these potential compounds can be used for anti-cancer drug research. Results: Five compounds i.e., rosmarinic acid, carnosic acid, (E,E)-5,9,13-pentadecatrien-2-one,6,10,14-trimethyl, α-amorphene and α-bis-abolol had high affinity and strong interaction with target proteins which resulted in a high ability to denature and inactivate those unexpected proteins. The docking pharmacophore features were also analyzed for clarifying the affinity results. Conclusion: These potential compounds were proposed for further research on drugs for treating colorectal cancer. The drug design simulation method helps to shorten the time and cost significantly in the selection of drug compounds for testing on living cells and animals.

Fulltext PDF Fulltext HTML

How to cite this article
Thanh-Diem Nguyen, Ly Le, Tu-Binh Vo, Kim-Lan Vo, Hoang-Minh Le and Huyen-Trang Vu, 2022. Evaluation of Colorectal Cancer Inhibition Ability of Rosmarinus officinalis L. via Molecular Docking and Pharmacophore Analysis. International Journal of Pharmacology, 18: 262-278.

Keywords: colorectal cancer, pharmacophore, computer-aided drug design, S. allylcysteine, molecular docking, Rosmarinus officinalis and apoptosis

INTRODUCTION

Colorectal Cancer (CRC) is a tumour that develops in the colon, rectum, or appendix. This is the third most common cancer, the second common cause of death in Western countries1. The incidence of CRC is on the rise worldwide, especially in developing countries2. In Vietnam, CRC is ranked as the fifth group of cancer with the number of new cases causing death at a rate of 4.1% among all types of cancers (https://gco.iarc.fr/). Like other types of cancer, CRC is caused by the changes in the genetic system that leads to uncontrol of cell division. The deletion mutation of gene loci related to tumour suppressor genes in the chromosome was reported to relate to the development of CRC in some previous studies, especially genes relating to cell proliferation and apoptosis such as BRAF, TP53, KRAS and ALK 3-6. The BRAF gene is a proto-oncogene belonging to the Serine/Threonine Kinase family. BRAF protein expressed somatic mutations in a variety of tumours, primarily malignancies1. The mutated β-catenin gene increased cell proliferation and inhibits apoptosis. This gene mutation accounts for up to 10% of all CRC cases7. The TP53 gene encodes a protein that aids in the cell cycle and apoptosis8. The TP53 gene mutation was found in more than 50% of cases of CRC, this is considered a marker in the development of tumours to cancer9. The KRAS gene encoding the Ras protein is responsible for the control of cell growth, differentiation and apoptosis. Some human cancers have been shown to relate to the expression of mutated Ras protein (oncogenic Ras). The appearance of mutant Ras proteins accounts for 15-20% in malignant tumours10 and mutation of the KRAS gene accounts for 25-60% of cases of CRC11. The genetic information of protein Anaplastic Lymphoma Kinase (ALK) which is involved in cell growth is from the gene ALK. Mutations (changed mutation) of the ALK gene and protein have been found in several types of cancer, including neuroblastoma and lung cancer. The appearance of the mutant ALK protein increased the growth of cancer cells12. These genes encode proteins that control cell proliferation and apoptosis and in turn, mutated proteins cause uncontrolled cell proliferation leading to tumour creation. The inactivation of these mutant proteins will help prevent the growth of tumours13.

The common cancer treatments include chemotherapy and radiotherapy. However, these methods often adversely affect the health of patients. Therefore, many studies have suggested the use of natural compounds in tumour suppression. These compounds can interact with mutated proteins that cause cancer, leading to the inhibition of tumour growth but little damage to the human body. Some compounds extracted from aged garlic (Allium sativum), especially S-allylcysteine and S-allylmercapto-L-cysteine have been shown to prevent the growth of certain types of cancer14,15. The flavonoids from papaya seeds also showed positive results when treated on some cancer cell lines in mice16. Rosemary (Rosmarinus officinalis L.) is a popular plant in Vietnam that is often used for ornamental purposes, spice in cooking, or for repelling insects. In 2016, this plant was also proved to inhibit CRC cells in mice17 by the two compounds rosmarinic acid and carnosic acid through in vitro test. However, there are still many other compounds of rosemary which are abundant and have not been put into research. Rosemary essential oil accounts for 27% of the plant, contains camphor (5.0-21%), 1.8-cineole (15-55%), α-pinene (9.0-26%), borneol (1.5-5.0%), camphene (2.5-12%), β-pinene (2.0-9.0%), limonene (1.5-5.0%)18 and other bioactive substances such as rosmarinic acid (8%), carnosic acid (30%), carnosol (17%) and ursolic acid (6%)19, which and can be extracted from different organs i.e., the leaves, stems and flower stalks.

Even so, searching for potential anti-cancer compounds through in vitro and in vivo tests is extremely time-consuming and costly14-17. With the development of computer science, simulation approaches have been effectively applied in many areas of life, including medical science, which can overcome those mentioned problems. The Structure-Based Drug Design (SBDD) method allows the batch simulation of docking between many plant compounds and disease-causing molecules just in hours20. The docking pharmacophores with higher affinity, i.e., lower binding energy required, are potential results for protein denaturation leading to inactivation of the target molecules. From initial docking results, potential compounds can be used to perform further wet experiments which require significantly less time and cost. This Computer-Aided Drug Design (CADD) method, which is a combination of computer science, chemistry, biology has been proven to be important for the development of new drugs from herbal plants. In this study, we simulated the binding affinity between compounds of rosemary and some mutated proteins causing a colorectal tumour.

The study aims to propose potential compounds for inhibiting tumours of CRC, serving for further steps of drug treatment on this dangerous disease.

MATERIALS AND METHODS

Study area: The study was carried out at the Department of Biotechnology, Nguyen Tat Thanh University, Vietnam from July, 2020-June, 2021).

Ligands and proteins preparation: Thirty compounds of rosemary used as ligands in this study (Table 1) were referenced from many published sources18,19,21.

Table 1: Information and 2-D structure downloaded from the ZINC database of thirty studied ligand compounds of the rosemary plant
Number Zinc Name Structure Molecular weight (Dalton) xlogP
1 ZINC00899870 Rosmarinic acid 359.31 1.63
2 ZINC03984016 Carnosic acid 331.432 4.6
3 ZINC12358879 (E,E)- 5, 9, 13- Pentadecatrien -2-one, 6 10, 14-trimethyl 262.437 6
4 ZINC01849759 α –bis-Abolol 222.372 4.68
5 ZINC02083320 Caryophyllene oxide 220.356 4.14
6 ZINC01677809 LINALYL PROPIONATE 210.317 4.28
7 ZINC57988166 Copaene 204.357 5.75
8 ZINC08234282 Caryophyllene 204.357 5.17
9 ZINC30726967 alpha-caryophyllene 204.357 5.31
10 ZINC70455185 α –Amorphene 204.357 5.97
11 ZINC00388664 L-Bornyl acetate 196.29 3.05
12 ZINC00899536 5-Methyl-2-(1-methylethyl)-phenol, acetate 192.258 2.91
13 ZINC00001411 o-Methyl eugenol 164.204 2.1
14 ZINC02510141 di-n-Butylethylamine 158.309 3.59
15 ZINC30724426 Sabinene hydrate 154.253 2.32
16 ZINC00967566 Eucalyptol 154.253 2.72
17 ZINC00968131 4-Thujanol 154.253 2.32
18 ZINC01529819 α –Linalool 154.253 3.21
19 ZINC00967533 L-Borneol 154.253 2.35
20 ZINC03861537 Terpinen-4-ol 154.253 2.6
21 ZINC02034811 3-Pinanone 152.237 2.39
22 ZINC14588455 Carvone 150.221 2.51
23 ZINC00967600 Verbenone 150.221 2.44
24 ZINC33845547 (Z)-Cinerone 150.221 2.06
25 ZINC18157343 Piperitenone 150.221 2.51
26 ZINC00967562 3-Carene 136.238 3.45
27 ZINC00968230 Camphene 136.238 3.33
28 ZINC59586951 2-Carene 136.238 3.45
29 ZINC02003408 oct-7-en-4-ol 128.215 2.53
30 ZINC00901249 3,4-Dimethoxy styrene 120.151 1.74

Molecular information of ligand was downloaded from ZINC database (http://zinc.docking.org/) including chemical structure, xlogP, aromatic rings, number of rotation bonds and was then saved as A Tripos Mol2 format. All the amide bonds of each ligand were made to not rotate using AutoDockTools 1.5.6 software22. The data was then turned into PDBQT (Protein Data Bank (PDB), Partial Charge (Q) and Atom Type (T)) format, which is a supported format for running on the AutoDock 4.0 software and increasing the storage capacity of atomic coordinates, partial charge, atomic types of docking molecules in comparison with previous format (http://autodock. scripps. edu/).

Six mutated proteins involved in causing CRC including a mutated form of each four proteins β-catenin (PDB molecular ID i.e., 1JPW), TP53 (4IBW), KRAS (4TQ9), ALK (5FTO) and two mutated forms of BRAF protein (5HID and 4R5Y)23-28 were considered as receptors for docking in this study (Table 2). Other molecular information and 3D structure of these proteins were also recorded from PDB (http://www.rcsb.org/) including resolutions, chains, existed ligands and determination methods. Each protein was prepared using AutoDockTools software 1.5.6 to achieve optimal simulation through 4 steps: (1) Adding polarized hydrogens, (2) Fusing non-polar hydrogens, (3) Removing water molecules and (4) Creating grid boxes. Adding polarized hydrogen bonds is important for docking since hydrogen bonds play a major role in stabilizing protein-ligand complexes29. As water molecules do not join the docking, the removal of water molecules from proteins makes computational accounts easier and avoids interference in searching for ligand molecules, which can create more favourable contact with protein receptors30. Grid boxes were established for verifying docking regions on 6 target proteins with 30×30×30 dimensions and default spacing at 1.000 Å (Table 3). Creating a grid box helps the program to determine the appropriate binding space between protein and ligand, thereby providing optimal binding results31. The data was then saved in PDBQT format for docking in the next step.

Molecular docking and pharmacophore analysis: One ligand was docked with one receptor in the space of one grid box for each running. The rigid docking simulation between a target protein and ligand was first performed using the AutoDock Vina program32. Result data of docking was converted into PDB (Protein Data Bank) format using OpenBabel program33 and was visualized by BIOVIA Discovery Studio Visualizer software34. Pharmacophore features of the simulation were analyzed based on the affinity and molecular interactions. For further analyses, flexible docking was next conducted. In the flexible docking, besides one protein receptor and one ligand, a flexible amino acid inside the receptor was required as a flexible factor to be included in the running setup35. The amino acids that are tightly bound to ligand from the result of rigid docking were chosen for this flexible docking step. Pharmacophore features of flexible docking were analyzed in comparison with the previous rigid pharmacophore.

Table 2: The information obtained from the PDB database of six mutated proteins involved in causing CRC in the study
PDB accession Resolution Chains Existed ligands Structure
1JPW 2.5 Å A, B, C _
4IBW 1.791 Å A, B A: Zn ion A, B: 1,2-Ethanediol
4R5Y 3.5 Å A, B A, B: C25H17F3N4O3
4TQ9 1.491 Å A, B A, B: GDP, Mg ion
5HID 2.5 Å A, B A, B: B1E, PEG
5FTO 2.22 Å A A: YMX


Table 3: Coordinate and dimension information of 10 grid boxes established for verifying docking regions on 6 target proteins in the study
Protein accession Grid box
Center x
Center y
Center z
Size (Å)
5HID GRID 1
3.056
-13.417
-9.417
30×30×30
  GRID 2-full
3,917
-1,667
-11,861
30×30×30
1JPW GRID 1-full
153,194
-1,861
6,528
30×30×30
4IBW GRID 1
-26,139
-7,5
-23,889
30×30×30
  GRID 2
-23.859
1,53
-15,86
30×30×30
4R5Y GRID 1
19,776
13,364
-15,307
30×30×30
  GRID 2-full
17,361
0,444
-1,361
74×12×18
4TQ9 GRID 1
0,417
-10,028
37,889
30×30×30
  GRID 2-full
-5,502
-22,603
26,972
42×36×14
5FTO GRID 1
6,676
19,601
8,223
30×30×30

RESULTS

Rigid docking results: The rigid docking results of 30 ligands with 6 target proteins at different grid boxes, respectively were shown in detail in Table 4. In general, rosmarinic acid and carnosic acid showed good binding results with all six examined proteins. Rosmarinic acid gave the highest affinity with 4TQ9 protein at the lowest binding energy -10.4 kcal mol–1 and with the remaining proteins at around -9.7 and -8.8 kcal mol–1 (Fig. 1).

Fig. 1:
Free binding energy between 30 ligands of Rosemary plant and 6 CRC carcinogenic proteins by Rigid docking simulation
The lower the free binding energy was required, the higher the binding affinity was

The following was carnosic acid which had the highest affinity for binding to 5FTO protein at -9.5 kcal mol–1 and to other proteins at a range from -9.4 and -8.9 kcal mol–1. Besides rosmarinic acid and carnosic acid, four other compounds i.e., (E, E)-5,9,13-pentadecatrien-2-one,6,10,14-trimethyl; α-caryophyllene, α-amorphene and α-bis-abolol which had the binding energy lower than -8.0 kcal mol1 with some of the mutated protein were also used for further flexible docking and pharmacophore analyzing.

Flexible docking and pharmacophore analysis: The absolute values of free binding energy referred from flexible docking were all better than that of rigid docking (Fig. 2). The differences ranged from 0.1 up to 1.7 kcal mol–1. The details were presented in Table 5. Rosmarinic acid gave the best affinity result with 4TQ9 at 11.1 kcal mol–1 instead of -10.4 kcal mol–1 from rigid docking (Fig. 2). The following was a complex of carnosic acid and 1JPW at -10.7 kcal mol–1, which was better than rigid docking by a distance of 1.7 kcal mol–1. Flexible docking of other three ligands α-abolol, α-amorphene and (E, E)-5,9,13-pentadecatrien-2-one,6,10,14-trimethyl also created a favourable affinity with 4R5Y, 5FTO and 4R5Y respectively at -10.0, -9.9 and -9.4 kcal mol–1, corresponding. Though the docking result was better, the flexible binding energy of α-caryophyllene with both of target proteins 5FTO and 4R5Y were still not reached -9.0 kcal mol–1, this compound was not included in the following analysis.

Fig. 2:

Best free-binding energy of 6 potential ligands with target proteins based on flexible docking and rigid docking

The pharmacophore of some high binding complexes was analyzed for more clarity of the binding mechanism. In comparison with rigid docking (Fig. 3a), flexible docking of rosmarinic acid with 4TQ9 produced 3 additional van der Waals bonds and 1 attractive charge (Fig. 3b). Thus, even though less than 1 Pi-cation and an additional unfavourable bump were present, the interaction affinity of this flexible complex was still better by about 0.7 kcal mol–1. In the complex between carnosic acid and 1JPW (Fig. 3c-d), despite the reduction of 1 van der Waals bond and the appearance of two more unfavourable bumps in flexible docking, there was an increased range of molecular bonds including 2 Hydrogen bonds, 1 Akyl bond, 1 Pi-Alkyl bond and 1 charge bond (Fig. 3d), resulting in a significant increase of the interaction affinity (from -9.0 and -10.7 kcal mol–1). This showed that flexible docking creates more sites of interaction between ligand and protein than rigid docking.

The same happened when comparing rigid and flexible pharmacophore in the complexes of 5FTO with α-amorphene (Fig. 3e-f) and 4R5Y with (E,E)-5.9,13-pentadecatrien-2-one, 6,10,14-trimethyl (Fig. 3g-h), although only 1 alkyl bond was improved and the free bond energy difference was not very high.

Table 4:Free-binding energy (kcal mol–1 ) between 30 ligands of rosemary plant and 6 CRC carcinogenic proteins by rigid docking simulation
Bold cell: Free-binding energy that is lower than -8 kcal mol–1


Fig. 3 (a-j):

Pharmacophore features the highest affinity complexes of 5 potential compounds with target proteins, (a, b): Complex of 4TQ9-Rosmarinic acid, (c, d): Complex of 1JPW and Carnosic acid, (e ,f): Complex of 5FTO and (α) Amorphene, (g, h): Complex of 4R5Y and (E,E)-5,9,13-pentadecatrien-2-one,6,10,14-trimethyl, (i, j): Complex of 5HID and a-bis-Abolol
(a, c, e, g, i): Pharmacophores from rigid docking, ( b, d, f, h, j): Pharmacophores from flexible docking. The hydrocarbon structure of the ligand was shown in a black frame. The global shape was the amino acid of the receptor that has interactions with the ligand. Different interactions of complex were represented in different corresponding colours)



Table 5: Free-binding energy between flexible docking of 6 potential ligands with CRC carcinogen receptors using different flexible amino acids in comparison with rigid docking
      Free-binding energy of docking
Ligand Protein Grid box Flexible docking using different amino acids             Minimum Rigid docking
Carnosic acid 1JPW 1JPW arg661b -8.9 asn362c -9.1 leu286c -9 thr289c -9.4 thr330c -9 tyr331c -10.7 arg582b -8.5 gln322c -9.4     -10.7 -9
  4IBW 4IBW asn268a -8.3 leu111a -8.3 ser269a -8 thr102a -8.8             -8.8 -8.2
    4R5YA ile527a -9.1 ile572a -9 leu505a -10.1 leu514a -9.7 thr529a -8.9 val504a -9.2         -10.1 -8.9
  4R5YA 4R5Y asp555b -8.8 thr589b -8.8 val511b -8.8                 -8.8
  4TQ9 4TQ9 full arg161b -9.5 asp154a -9.5 ile142a -9.7 thr158a -9.2 thr127a -9.4 tyr157b -9.9         -9.9 -9.4
  5FTO 5FTO leu1196a -10.3 leu1256a -10 met1199a -9.6 phe1127a -9.8 val1130a -9.5 val1180a -9.5         -10.3 -9.5
  5HID 5HID ile463a -10.3 leu514a -9.5 phe583a -9.5 phe595a -10.4 thr529a -9.3 trp531a -9.4 val471a -9.4       -10.4 -9.3
    5HID full ile463a -10.4 phe583a -9.5 phe595a -9.6 thr529a -9.1 trp531a -9.2 val471a -9.3           -9.1
Rosmarinic acid 1JPW 1JPW arg587b -9.5 asp249c -9.3 his524b -9.2 lys288c -9.4 thr289c -9.2 val584b -9 4         -9.5 -9.2
  4IBW 4IBW -9.5 arg282a -9.4 gln144a -8.9 his115a -9.3 phe113a -9.3 ser116a -9.7 trp146a         -9.7 -9.3
    4R5Y 4R5YA -9.8 ile527a -9.2 leu505a -9.8 leu514a -9.4 thr529a -9.8 val504a         -9.8 -9.4
    4R5Yfull arg562a -9.3 val511a -9.1                   -8.9
    4TQ9 4TQ9full -11.1 arg161b -10.6 asp154a -10.1 gln131a -10.5 ile142a -10.6 arg135a         -11.1 -10.4
    5FTO 5FTO -9.8 leu1256a -9.8 phe1127a -9.8 val1130a             -9.8 -9.7
    5HID 5HID -10.1 ile463a -9.8 leu514a -10.4 phe595a -9.6 thr529a -9.8 val471a         -10.4 -9.7
    5HIDfull ile463a -8.1 leu514a -8.7 phe595a -8.4 thr529a -8.4 val471a -8.2             -8.3
(E,E)- 5, 9, 13-Pentadecatrien 4R5Y 4R5YA leu505a -9.4 leu514a -8.9 thr529a -8.9 val471a -8.9 val504a -8.8 lys483a -9.1         -9.4 -8.9
-2-one, 6, 10, 14-trimethyl 5FTO 5FTO leu1122a -9.3 leu1256a -9.3 phe1127a -9.3 val1130a -9.1             -9.3 -9
α-Caryophyllene 4R5Y 4R5YA ile463a -8.2 leu514a -8.2 phe583a -8.4 phe595a -8.5 thr529a -8.2 trp531a -8.3 val471a -8.3       -8.5 -8.2
  5FTO 5FTO leu1122a -8.8 leu1196a -8.5 leu1256a -8.5 phe1127a -8.5 val1130a -8.6           -8.8 -8.4
α-Amorphene 4R5Y 4R5YA ile463a -8.6 leu514a -8.3 phe583a -8.7 phe595a -8.4 thr529a -8.4 trp531a -8.6 val471a -8.4       -8.7 -8.4
  5FTO 5FTO leu1122a -9.9 leu1196a -9.9 leu1256a -9.8 phe1127a -9.9 val1130a -9.9 lys1150a -9.7         -9.9 -9.8
  5HID 5HID ile463a -8.6 leu514a -8.5 leu597a -8.4 phe583a -8.7 phe595a -8.5 thr529a -8.3 trp531a -8.5 val471a -8.4     -8.7 -8.3
  5HID full ile463a leu514a -8.6 phe583a -8.5 phe595a -8.7 trp531a -8.5 thr529a -8.3 val471a -8.3 -8.4         -8.3
α-bis-Abolol 4R5Y 4R5YA ile463a -8.8 ile572a -8.8 leu505a -8.7 phe583a -8.8 phe595a -8.9 thr529a -8.7 trp531a -9.1 val471a -8.9 leu514a -10 lys483a -8.9 -10 -8.8
  5FTO 5FTO leu1122a -8.1 leu1196a -8 leu1256a -8.3 phe1127a -8.2 val1130a -8 val1180a -8         -8.3 -8
5HID 5HID ile527a -8.2 leu505a -8.5 leu514a -9.3 phe583a -8.5 thr529a -8.3 trp531a -8.4 val471a -8.8 phe595a -8.1 cys532a -8.2     -9.3 -8.2
    5HID full ile527a -8.2 leu505a -8.4 leu514a -9.3 phe583a -8.5 thr529a -8.3 trp531a -8.4 val471a -8.8         -8.3


Table 6: Molecular interactions of the highest affinity complexes from rigid docking and flexible docking
Complex Docking type
Van der waals
Convetional hydrogen bond
Salt bridge
Alkyl
Pi-Alkyl
Attractive charge
Pi-cation
Pi-sigma
Unfavorable bump
Total
4TQ9-Rosmarinic acid Rigid
1
4
0
0
2
1
1
0
0
9
  Flexible
3
4
0
0
2
2
0
0
15
12
1JPW-Carnosic acid Rigid
5
1
1
2
0
0
0
0
0
9
  Flexible
4
3
0
3
1
1
0
0
2
14
5FTO- Rigid
0
0
0
9
6
0
0
0
0
15
α-Amorphene Flexible
0
0
0
10
6
0
0
0
0
16
4R5Y-(E,E)5,9,13- Rigid
1
0
0
5
0
0
0
0
0
6
Pentadecatrien-2-one Flexible
1
0
0
6
0
0
0
0
0
7
6,10,14-trimethyl
5HID- Rigid
1
0
0
6
5
0
0
1
0
13
α-bis-Abolol Flexible
1
1
0
7
2
0
0
0
3
14

However, this result still recommended the importance of some molecular bonds in the interaction affinities. For the interaction between 5HID and α-bis-abolol, the energy difference was quite different (from -8.2 and -9.3 kcal mol–1) due to the increase of 1 Alkyl bond and 1 Hydrogen bond (Fig. 3i-j). The statistics of intermolecular interactions were detailed in Table 6.

DISCUSSION

The two compounds i.e., carnosic acid and rosmarinic acid showed the best binding with all studied colorectal carcinogenic proteins. Previously, carnosic acid was also tested on CRC Caco-2, HT29 and LoVo cell lines by Barni et al.36. The study found out that this compound had strong inhibition of the tumour growing by inactivating both the carcinogenic mRNA, which encodes the COX-2 cancer-causing pathway and its protein. In 2016, rosmarinic acid and carnosic acid were also proven to have an anti-cancer effect on some colorectal cancer cell lines by Jessy Moore et al.17. However, it took 24 hrs to test in vitro inhibitory ability of these compounds on each cell line. As for in vivo test, the treatment effect on mice was evaluated after 11, 16 weeks using carnosic acid and rosmarinic acid, respectively. For our in silico study, it took only hours to get the docking result and select the best ligands. Although it is necessary to further perform in vitro or in vivo tests for drug development, the computer works significantly reduce cost and time-consuming as the first step for selecting potential subjects from a large number of new compounds of herbal plants37. Besides, our study was completely consistent with the studies of Moore17 and Barni36, which not only reconfirmed the role of these two compounds in inhibition of colorectal cancer but also convincingly demonstrated the reliability of this simulation method for other Computer-Aided Drug Design studies.

The change in interaction energy of flexible docking compared with rigid docking in the complex between rosmarinic acid and 4TQ9 occurred due to the addition of 2 van der Waals bonds and 1 attractive charge bond and the appearance of an unfavourable bump. Van der Waals is an attractive force due to dipole-induced interactions, which is weak in comparison with chemical bonds38. Besides, the existence of unfavourable pumps, which is known as unexpected intermolecular steric clash, have been proved to show unstable interactions and binding between interacting amino acids and drug atoms39 as well. Hence the significant increase of binding capacity, in this case, might be due to the appearance of the attractive charge, which in turn is caused by the existence of the -COO- group in the structure of rosmarinic acid (Fig. 3b). The carboxylic acid functional group plays a cardinal role in the biochemistry of living systems as well as in drug design. Since endogenous substances, such as amino acids, triglycerides and prostanoids, possess the carboxylic acid moiety. The acidity as well as the ability to establish relatively strong electrostatic interactions and hydrogen bonds is the reason why this functional group is often part of drug-target interactions40 and pharmacophore of diverse classes of therapeutic agents41. The two compounds rosmarinic acid and carnosic acid, which gave the best binding results in both rigid and flexible docking on this study, all contain this -COO group.

Furthermore, these two ligands also contained aromatic rings in their structure. Rosmarinic acid had two phenol rings, the greatest number of phenol rings in comparison with other compounds in the study. Polyphenol components have been identified for their ability to prevent various types of cancer, in both experimental and simulated research42,43. These compounds had the potential to change the primary and secondary structures due to methyl, glycosyl and hydroxylation processes44,45, which make it easy to link with amino acids to increase the binding capacity between ligands and receptor proteins. The interactions of Pi-cation and Pi-alkyl were all created due to the existence of a pi-electron cloud over these aromatic groups. Pi-alkyl is the interaction of the aromatic group and electron group of an alkyl group. A large number of pi-sigma (pi-alkyl and pi-cation) interactions were mainly involved in charge transfer, which helps to transfer drugs between receptor binding sites38. Meanwhile, Pi-cation interaction is the binding force between the cations and the pi surface (the face of an electron-rich pi system) of the aromatic structure through a non-covalent force. Pi-cation was important in many proteins that bind ligands or cation substrates46.

Three other potential compounds i.e., α-amorphene and α-bis-abolol and (E,E)-5,9,13-pentadecatrien-2-one,6,10,14-trimethyl mainly consisted of methyl groups (-Me) when they linked to the receptors. The methyl group is non-polar radicals and provided electrons to other groups47 to create alkyl bonds. The addition of a methyl group made a molecule more hydrophobic that supporting linkage with biological molecules48. These hydrophobic interactions were reported to contribute to the binding of many ligand-protein systems before49. Alkyl bonds were also reported to increase the lipophilicity of the drug and created favourable conditions for the drug to penetrate the cell membranes50.

On the other hand, the presence of functional groups as -OH and -CO in the structure of three ligands rosmarinic acid, carnosic acid and α-bisabolol also supported protein binding. The Carbonyl group at the C-ring of flavonoid played an important role in the ligand-target interaction, by hydrogen bond interaction to Ser530A and Arg120A residue51. In contrast with (-Me), hydroxyl and carbonyl groups are polar radicals52 due to the high electronegativity of oxygen. Hence the hydrogen bonds (electrostatic bond between hydrogen and the more electronegative atoms) of these compounds with hydrogen atoms in the environment were created. The free energy for hydrogen bonding can vary between -1.5 and -4.7 kcal mol–1. The best ligand in this study, rosmarinic acid, created four hydrogen bonds with 4TQ9, followed by carnosic acid with three hydrogen binding toward 1JPW. The interaction between the -OH group of α-bisabolol and the amino acid THR A:529 of 5HID, which was not created in rigid docking, contributed to the increase of linking affinity (from -8.2 and-9.3 kcal mol–1) during flexible docking. Hydrogen bonds were intermolecular interactions that were common in biological complexes53 and were contributions to the specificity of molecular recognition54.

From the better results of flexible docking, it has been shown that flexible docking provides more sites of molecular interaction than rigid docking. Otherwise, proteins can change their initial stable structure to fit with the ligands. In living organisms, proteins are flexible objects. However, rigid docking assumed that proteins and ligands were immobilized objects, so the docking was performed only at one coordinate. Therefore, the results were extremely limited. On the other hand, flexible docking tried to simulate receptors and ligands as flexible objects. Hence, the docking was performed at several coordinates55 in which the most durable combination with the least energy required was created. In the flexible docking, a flexible amino acid inside the receptor was required as a flexible factor to be included in the running setup. Hence the ligand could adjust to the most stable protein binding site and the simulation was more reliable and just similar to what happens in vivo process.

CONCLUSION

The ligand-protein docking is to simulate how the ligand competes with substrates inactive regions of carcinogenic proteins for inactivating that protein, leading to the inhibition of the tumour growing. Using molecular docking and pharmacophore analysis, our study has confirmed therapeutic effects and clarified the tumour-inhibition ability of Rosmarinus officinalis L. based on molecular interactions between examined compounds with the carcinogenic proteins. Five compounds i.e., rosmarinic acid, carnosic acid, (E,E)-5,9,13-pentadecatrien-2-one, 6,10,14-trimethyl, α-amorphene and α-bis-abolol from rosemary were proposed as potential compounds in colorectal tumour inhibition. The study strongly confirmed the role and the reliability of computer works in supporting other drug development studies.

SIGNIFICANCE STATEMENT

This study discovers the ability of compounds from the herbal plant Rosmarinus officinalis L. that can be beneficial for developing drugs targeting inhibition of different proteins causing colorectal cancer. This study will help the researcher to uncover the critical areas of drug-based docking and interaction models of potential compounds of Rosmarinus officinalis L. with different target proteins that many researchers were not able to explore and at the same time emphasize the useful role of the docking method in process of drug development.

ACKNOWLEDGMENTS

This work is supported by Nguyen Tat Thanh University under Grant No. 2021.01.54/HĐ-KHCN. We are also grateful to the Computational Biology Center of International University, Vietnam National University for providing the computer resources.

REFERENCES

  • Sameer, A.S., 2013. Colorectal cancer: Molecular mutations and polymorphisms. Frontiers Oncol., Vol. 3.
    CrossRef    


  • Bray, F., J. Ferlay, I. Soerjomataram, R.L. Siegel, L.A. Torre and A. Jemal, 2018. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA: Cancer J. Clin., 68: 394-424.
    CrossRef    Direct Link    


  • Williams, D.S., D. Mouradov, C. Browne, M. Palmieri and M.J. Elliott et al., 2020. Overexpression of TP53 protein is associated with the lack of adjuvant chemotherapy benefit in patients with stage III colorectal cancer. Mod. Pathol., 33: 483-495.
    CrossRef    Direct Link    


  • Porru, M., L. Pompili, C. Caruso, A. Biroccio and C. Leonetti, 2018. Targeting KRAS in metastatic colorectal cancer: Current strategies and emerging opportunities. J. Exp. Clin. Cancer Res., Vol. 37.
    CrossRef    


  • Davies, H., G.R. Bignell, C. Cox, P. Stephens and S. Edkins et al., 2002. Mutations of the BRAF gene in human cancer. Nature, 417: 949-954.
    CrossRef    PubMed    Direct Link    


  • Polakis, P., 2000. Wnt signaling and cancer. Genes Dev., 14: 1837-1851.
    CrossRef    PubMed    


  • Behrens, J., 2005. The role of the Wnt signalling pathway in colorectal tumorigenesis. Biochem. Soc. Trans., 33: 672-675.
    CrossRef    Direct Link    


  • Aubrey, B.J., A. Strasser and G.L. Kelly, 2016. Tumor-suppressor functions of the TP53 pathway. Cold Spring Harbor Perspect. Med., Vol. 6.
    CrossRef    


  • Liebl, M.C. and T.G. Hofmann, 2021. The role of p53 signaling in colorectal cancer. Cancers, Vol. 13.
    CrossRef    


  • Schubbert, S., K. Shannon and G. Bollag, 2007. Hyperactive Ras in developmental disorders and cancer. Nat. Rev. Cancer, 7: 295-308.
    CrossRef    Direct Link    


  • Lee, C.S., I.H. Song, A. Lee, J. Kang and Y.S. Lee et al., 2021. Enhancing the landscape of colorectal cancer using targeted deep sequencing. Sci. Rep., Vol. 11.
    CrossRef    


  • Webb, T.R., J. Slavish, R.E. George, A.T. Look and L. Xue et al., 2009. Anaplastic lymphoma kinase: Role in cancer pathogenesis and small-molecule inhibitor development for therapy. Expert Rev. Anticancer Ther., 9: 331-356.
    CrossRef    Direct Link    


  • Pfeffer, C.M. and A.T.K. Singh, 2018. Apoptosis: A target for anticancer therapy. Int. J. Mol. Sci., Vol. 19.
    CrossRef    


  • Thomson, M. and M. Ali, 2003. Garlic [Allium sativum]: A review of its potential use as an anti-cancer agent. Curr. Cancer Drug. Targets, 3: 67-81.
    CrossRef    PubMed    Direct Link    


  • Shukla, Y. and N. Kalra, 2007. Cancer chemoprevention with garlic and its constituents. Cancer Lett., 247: 167-181.
    CrossRef    Direct Link    


  • Pathak, N., S. Khan, A. Bhargava, G.V. Raghuram and D. Jain et al., 2014. Cancer chemopreventive effects of the flavonoid-rich fraction isolated from papaya seeds. Nutr. Cancer, 66: 857-871.
    CrossRef    Direct Link    


  • Moore, J., M. Yousef and E. Tsiani, 2016. Anticancer effects of rosemary (Rosmarinus officinalis L.) extract and rosemary extract polyphenols. Nutrients, Vol. 8.
    CrossRef    


  • Begum, A., S. Sandhya, S.S. Ali, K.R. Vinod, S. Reddy and D. Banji, 2013. An in-depth review on the medicinal flora Rosmarinus officinalis (Lamiaceae). Acta Scientiarum Polonorum: Technologia Alimentaria, 12: 61-73.
    PubMed    Direct Link    


  • Andrade, J.M., C. Faustino, C. Garcia, D. Ladeiras, C.P. Reis and P. Rijo, 2018. Rosmarinus officinalis L.: An update review of its phytochemistry and biological activity. Future Sci. OA, Vol. 4.
    CrossRef    


  • Yu, W. and A.D. MacKerell, 2017. Computer-Aided Drug Design Methods. In: Antibiotics: Methods and Protocols, Sass, P. (Ed.), Humana Press, New York, ISBN: 978-1-4939-6634-9, pp: 85-106
    CrossRef    Direct Link    


  • Moliner, C., V. López, L. Barros, M.I. Dias, I.C.F.R. Ferreira, E. Langa and C. Gómez-Rincón, 2020. Rosemary flowers as edible plant foods: Phenolic composition and antioxidant properties in Caenorhabditis elegans. Antioxidants, Vol. 9.
    CrossRef    


  • Morris, G.M., R. Huey, W. Lindstrom, M.F. Sanner, R.K. Belew, D.S. Goodsell and A.J. Olson, 2009. AutoDock4 and AutoDock Tools4: Automated docking with selective receptor flexibility. J. Comput. Chem., 30: 2785-2791.
    CrossRef    Direct Link    


  • Poy, F., M. Lepourcelet, R.A. Shivdasani and M.J. Eck, 2001. Structure of a human Tcf4–β-catenin complex.
    Nat. Struct. Biol., 8: 1053-1057.
    CrossRef    Direct Link    


  • Eldar, A., H. Rozenberg, Y. Diskin-Posner, R. Rohs and Z. Shakked, 2013. Structural studies of p53 inactivation by DNA-contact mutations and its rescue by suppressor mutations via alternative protein-DNA interactions. Nucleic Acids Res., 41: 8748-8759.
    CrossRef    Direct Link    


  • Tang, Z., X. Yuan, R. Du, S.H. Cheung and G. Zhang et al., 2015. BGB-283, a novel RAF kinase and EGFR inhibitor, displays potent antitumor activity in BRAF-mutated colorectal cancers. Mol. Cancer Ther., 14: 2187-2197.
    CrossRef    Direct Link    


  • Hunter, J.C., A. Manandhar, M.A. Carrasco, D. Gurbani, S. Gondi and K.D. Westover, 2015. Biochemical and structural analysis of common cancer-associated kras mutations. Mol. Cancer Res., 13: 1325-1335.
    CrossRef    Direct Link    


  • Foster, S.A., D.M. Whalen, A. Özen, M.J. Wongchenko and J. Yin et al., 2016. Activation mechanism of oncogenic deletion mutations in BRAF, EGFR, and HER2. Cancer Cell, 29: 477-493.
    CrossRef    Direct Link    


  • Menichincheri, M., E. Ardini, P. Magnaghi, N. Avanzi and P. Banfi et al., 2016. Discovery of entrectinib: A new 3-aminoindazole as a potent anaplastic lymphoma kinase (ALK), c-ros oncogene 1 kinase (ROS1), and pan-tropomyosin receptor Kinases (Pan-TRKS) inhibitor. J. Med. Chem., 59: 3392-3408.
    CrossRef    Direct Link    


  • Lippert, T. and M. Rarey, 2009. Fast automated placement of polar hydrogen atoms in protein-ligand complexes. J. Cheminf., Vol. 1.
    CrossRef    


  • Wong, S.E. and F.C. Lightstone, 2011. Accounting for water molecules in drug design. Expert Opin. Drug Discovery, 6: 65-74.
    CrossRef    Direct Link    


  • Feinstein, W.P. and M. Brylinski, 2015. Calculating an optimal box size for ligand docking and virtual screening against experimental and predicted binding pockets. J. Cheminf., Vol. 7.
    CrossRef    


  • Trott, O. and A.J. Olson, 2010. AutoDock Vina: Improving the speed and accuracy of docking with a new scoring function, efficient optimization and multithreading. J. Comput. Chem., 31: 455-461.
    CrossRef    PubMed    Direct Link    


  • O'Boyle, N.M., M. Banck, C.A. James, C. Morley, T. Vandermeersch and G.R. Hutchison, 2011. Open babel: An open chemical toolbox. J. Cheminf., Vol. 3.
    CrossRef    


  • Silva, N.S.R., L.K.S. Gonçalves, J.L. Duarte, J.S. Silva and C.F. Santos et al., 2014. Computational analysis of physicochemical, pharmacokinetic and toxicological properties of deoxyhypusine synthase inhibitors with antimalarial activity. Comput. Mol. Biosci., 04: 47-57.
    CrossRef    Direct Link    


  • Nguyen, H.T., L. Le and T.N. Truong, 2009. Top-hits for H1N1 identified by virtual screening using ensemble-based docking. PLoS Curr., Vol. 1.
    CrossRef    


  • Barni, M.V., M.J. Carlini, E.G. Cafferata, L. Puricelli and S. Moreno, 2012. Carnosic acid inhibits the proliferation and migration capacity of human colorectal cancer cells. Oncol. Rep., 27: 1041-1048.
    CrossRef    Direct Link    


  • Klebe, G., 2013. Protein–Ligand Interactions as the Basis for Drug Action. In: Drug Design, Klebe, G. (Ed.)., Springer, Berlin, Heidelberg, pp: 61-88
    CrossRef    Direct Link    


  • Arthur, D.E. and A. Uzairu, 2019. Molecular docking studies on the interaction of NCI anticancer analogues with human phosphatidylinositol 4,5-bisphosphate 3-kinase catalytic subunit. J. King Saud Uni. Sci., 31: 1151-1166.
    CrossRef    Direct Link    


  • Zainab, B., Z. Ayaz, M.S. Alwahibi, S. Khan and H. Rizwana et al., 2020. In-silico elucidation of Moringa oleifera phytochemicals against diabetes mellitus. Saudi J. Biol. Sci., 27: 2299-2307.
    CrossRef    Direct Link    


  • Ballatore, C., D.M. Huryn and A.B. Smith, 2013. Carboxylic acid (bio)isosteres in drug design. ChemMedChem, 8: 385-395.
    CrossRef    Direct Link    


  • Hajduk, P.J., M. Bures, J. Praestgaard and S.W. Fesik, 2000. Privileged molecules for protein binding identified from NMR-based screening. J. Med. Chem., 43: 3443-3447.
    CrossRef    Direct Link    


  • Zhou, Y., J. Zheng, Y. Li, D.P. Xu, S. Li, Y.M. Chen and H.B. Li, 2016. Natural polyphenols for prevention and treatment of cancer. Nutrients, Vol. 8.
    CrossRef    


  • Asensi, M., A. Ortega, S. Mena, F. Feddi and J.M. Estrela, 2011. Natural polyphenols in cancer therapy. Crit. Rev. Clin. Lab. Sci., 48: 197-216.
    CrossRef    Direct Link    


  • Halbwirth, H., 2010. The creation and physiological relevance of divergent hydroxylation patterns in the flavonoid pathway. Int. J. Mol. Sci., 11: 595-621.
    CrossRef    Direct Link    


  • Crozier, A., I.B. Jaganath and M.N. Clifford, 2009. Dietary phenolics: Chemistry, bioavailability and effects on health. Nat. Prod. Rep., 26: 1001-1043.
    CrossRef    Direct Link    


  • Taylor, R., 2016. Progress in the Understanding of Traditional and Nontraditional Molecular Interactions. In: Comprehensive Medicinal Chemistry III, Chackalamannil, S., D. Rotella and S.E. Ward (Eds.)., Elsevier Academic Press, United States, ISBN-13: 978-0-12-803201-5, pp: 67-100
    CrossRef    Direct Link    


  • Teixidor, F., G. Barberà, A. Vaca, R. Kivekäs, R. Sillanpää, J. Oliva and C. Viñas, 2005. Are methyl groups electron-donating or electron-withdrawing in boron clusters? Permethylation of o-carborane. J. Am. Chem. Soc., 127: 10158-10159.
    CrossRef    Direct Link    


  • Leung, C.S., S.S.F. Leung, J. Tirado-Rives and W.L. Jorgensen, 2012. Methyl effects on protein–ligand binding. J. Med. Chem., 55: 4489-4500.
    CrossRef    Direct Link    


  • Mecinović, J., P.W. Snyder, K.A. Mirica, S. Bai and E.T. Mack et al., 2011. Fluoroalkyl and alkyl chains have similar hydrophobicities in binding to the “hydrophobic wall” of carbonic anhydrase. J. Am. Chem. Soc., 133: 14017-14026.
    CrossRef    Direct Link    


  • Ditzinger, F., D.J. Price, A.R. Ilie, N.J. Köhl and S. Jankovic et al., 2019. Lipophilicity and hydrophobicity considerations in bio-enabling oral formulations approaches-a PEARRL review. J. Pharm. Pharmacol., 71: 464-482.
    CrossRef    Direct Link    


  • Herowati, R. and G.P. Widodo, 2017. Molecular Docking Analysis: Interaction Studies of Natural Compounds to Anti-Inflammatory Targets. In: Quantitative Structure-Activity Relationship, Kandemirli, F. (Ed.)., IntechOpen Limited, United States, ISBN-13: 978-953-51-3410-7,
    CrossRef    Direct Link    


  • Hanson, J.R., D. Phillips and J.D. Woollins, 2001. Functional Group Chemistry. 1st Edn., The Royal Society of Chemistry, Pages: 171
    Direct Link    


  • Nittinger, E., T. Inhester, S. Bietz, A. Meyder and K.T. Schomburg et al., 2017. Large-scale analysis of hydrogen bond interaction patterns in protein–ligand interfaces. J. Med. Chem., 60: 4245-4257.
    CrossRef    Direct Link    


  • Schaeffer, L., 2008. The Role of Functional Groups in Drug–Receptor Interactions. In: The Practice of Medicinal Chemistry, Wermuth, C.G., D. Aldous, P. Raboisson and D. Rognan (Eds.)., Elsevier Academic Press, United States, ISBN-13: 978-953-51-3410-7, pp: 359-378
    CrossRef    Direct Link    


  • Huang, S.Y., 2018. Comprehensive assessment of flexible-ligand docking algorithms: Current effectiveness and challenges. Briefings Bioinf., 19: 982-994.
    CrossRef    Direct Link    

  • © Science Alert. All Rights Reserved