You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/370374084

Antimalarial potential, LC-MS secondary metabolite profiling and


computational studies of Zingiber officinale

Article in Journal of biomolecular Structure & Dynamics · April 2023


DOI: 10.1080/07391102.2023.2205949

CITATIONS READS

2 161

10 authors, including:

Faloye Kolade Manish Kumar Tripathi


Obafemi Awolowo University All India Institute of Medical Sciences
23 PUBLICATIONS 84 CITATIONS 51 PUBLICATIONS 708 CITATIONS

SEE PROFILE SEE PROFILE

Stephen Adeleke Adesida Yemisi Oyetunde


Obafemi Awolowo University University of Ibadan
6 PUBLICATIONS 18 CITATIONS 2 PUBLICATIONS 6 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Stephen Adeleke Adesida on 01 May 2023.

The user has requested enhancement of the downloaded file.


Journal of Biomolecular Structure and Dynamics

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/tbsd20

Antimalarial potential, LC–MS secondary


metabolite profiling and computational studies of
Zingiber officinale

Kolade O. Faloye, Manish K. Tripathi, Stephen A. Adesida, Samuel A.


Oguntimehin, Yemisi M. Oyetunde, Adetola H. Adewole, Ifeoluwa I.
Ogunlowo, Esther A. Idowu, Uduak I. Olayemi & Olamide D. Dosumu

To cite this article: Kolade O. Faloye, Manish K. Tripathi, Stephen A. Adesida, Samuel A.
Oguntimehin, Yemisi M. Oyetunde, Adetola H. Adewole, Ifeoluwa I. Ogunlowo, Esther A. Idowu,
Uduak I. Olayemi & Olamide D. Dosumu (2023): Antimalarial potential, LC–MS secondary
metabolite profiling and computational studies of Zingiber officinale, Journal of Biomolecular
Structure and Dynamics, DOI: 10.1080/07391102.2023.2205949

To link to this article: https://doi.org/10.1080/07391102.2023.2205949

View supplementary material

Published online: 28 Apr 2023.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tbsd20
JOURNAL OF BIOMOLECULAR STRUCTURE AND DYNAMICS
https://doi.org/10.1080/07391102.2023.2205949

Antimalarial potential, LC–MS secondary metabolite profiling and computational


studies of Zingiber officinale
Kolade O. Faloyea, Manish K. Tripathib, Stephen A. Adesidac, Samuel A. Oguntimehind, Yemisi M. Oyetunded,
Adetola H. Adewolee , Ifeoluwa I. Ogunlowoc, Esther A. Idowuc, Uduak I. Olayemic and Olamide D. Dosumuf
a
Department of Chemistry, Faculty of Science, Obafemi Awolowo University, Ile-Ife, Nigeria; bPharmaceutical Chemistry Research Laboratory,
Department of Pharmaceutical Engineering & Technology, Indian Institute of Technology (Banaras Hindu University), Varanasi, India;
c
Department of Pharmacognosy, Faculty of Pharmacy, Obafemi Awolowo University, Ile-Ife, Nigeria; dDepartment of Pharmacognosy, Faculty
of Pharmacy, University of Ibadan, Ibadan, Nigeria; eDepartment of Chemistry, University of Pretoria, Pretoria, South Africa; fDepartment of
Botany, Faculty of Science, Obafemi Awolowo University, Ile-Ife, Nigeria
Communicated by Ramaswamy H. Sarma

ABSTRACT ARTICLE HISTORY


Malaria is among the top-ranked parasitic diseases that pose a threat to the existence of the human Received 25 October 2022
race. This study evaluated the antimalarial effect of the rhizome of Zingiber officinale in infected mice, Accepted 17 April 2023
performed secondary metabolite profiling and detailed computational antimalarial evaluation through
KEYWORDS
molecular docking, molecular dynamics (MD) simulation and density functional theory methods. The
Density functional theory;
antimalarial potential of Z. officinale was performed using the in vivo chemosuppressive model; sec- Liquid chromatography–
ondary metabolite profiling was carried out using liquid chromatography–mass spectrometry (LC–MS). mass spectrometry; malaria;
Molecular docking was performed with Autodock Vina while the MD simulation was performed with molecular docking;
Schrodinger desmond suite for 100 ns and DFT calculations with B3LYP (6-31G) basis set. The extract molecular dynamics
showed 64% parasitaemia suppression, with a dose-dependent increase in activity up to 200 mg/kg. simulation; Zingiber
The chemical profiling of the extract tentatively identified eight phytochemicals. The molecular dock- officinale
ing studies with plasmepsin II and Plasmodium falciparum dihydrofolate reductase-thymidylate syn-
thase (PfDHFR-TS) identified gingerenone A as the hit molecule, and MMGBSA values corroborate the
binding energies obtained. The electronic parameters of gingerenone A revealed its significant anti-
malarial potential. The antimalarial activity elicited by the extract of Z. officinale and the bioactive
chemical constituent supports its usage in ethnomedicine.

1. Introduction plasmepsins (Bhaumik, Gustchina, and Wlodawer 2012; Liu


2017). Four of the ten plasmepsins including plasmepsin II,
Malaria is a parasitic disease that manifests in humans and is
encoded by the genome of P. falciparum are associated with
caused by Plasmodium parasites transmitted to humans
its food vacuole, which is directly related to the role of the
through the bite of an infected female Anopheles mosquito
parasite in the pathophysiology of malaria (Chugh et al.
(Cator et al. 2012; Yadav, Tripathi, and Yadav 2021). It is a 2013). Another important enzyme implicated in malaria is P.
severe and often deadly disease, especially when the parasite falciparum dihydrofolate reductase-thymidylate synthase
involved is Plasmodium falciparum (Jensen, Adams, and Hviid (PfDHFR-TS) which converts dihydrofolate to tetrahydrofolate
2020). Despite various interventions in treatment and preven- and is involved in the synthesis of purines and thymidylate,
tion, about 241 million new cases and 627,000 deaths were all of which are important in DNA/RNA synthesis and repair,
attributed to malaria in 2020 (WHO, 2022). This same year, cell division and protein synthesis (Sharma and Chauhan
African countries accounted for 96% of the reported deaths 2012; Shamshad, Bakri, and Mirza 2022). These actions by
due to malaria, 31.9% of which occurred in Nigeria (WHO, PfDHFR-TS make it a prime target in antifolate antimalarial
2021). Clinical symptoms of malaria, such as high fever, chills therapy (Yuthavong et al. 2005). Therefore, inhibiting the
and rigours, flu-like illness, head, muscle and joint aches, activities of PfDHFR-TS and Plasmepsin II is an effective strat-
among others, occur when the parasite is in the blood stage egy in the research and development of novel drug mole-
of its cycle in the human host (Rudrapal and Chetia, 2017). cules for antimalarial therapy.
The roles of some enzymes in the pathophysiology of Z. officinale Roscoe is a plant in the family Zingiberaceae.
malaria cannot be underestimated. The initial degradation of The rhizome of the plant is used as a spice and for its medi-
hemoglobin in the food vacuole of the malaria parasite to cinal properties. It is used to treat cough and cold, head-
obtain nutrients is catalyzed by aspartic proteases known as aches, nausea, flatulence, malaria, and vomiting (Bhowmik

CONTACT Kolade O. Faloye kollintonx1@gmail.com Department of Chemistry, Faculty of Science, Obafemi Awolowo University, Ile-Ife, Nigeria; Uduak I.
Olayemi uduakolayemi@gmail.com Department of Pharmacognosy, Faculty of Pharmacy, Obafemi Awolowo University, Ile-Ife, Nigeria
Supplemental data for this article can be accessed online at https://doi.org/10.1080/07391102.2023.2205949.
ß 2023 Informa UK Limited, trading as Taylor & Francis Group
2 K. O. FALOYE ET AL.

et al. 2010; Vandebroek and Picking 2020). Its antioxidant, anti- in aluminum cages with wood shavings used as beddings and
inflammatory, antimicrobial, and anticancer activities have been allowed free access to water and food under 12 h day/night
reported (Otunola et al. 2010; Yusuf et al. 2018). Kaushik et al. cycle. They were acclimatized and cared for according to stand-
(2013) reported the in vitro assay of the ethyl acetate solvent ard procedures in the handling of animals (NIH Publication NO
extract of the rhizome of Z. officinale as showing excellent anti- 82-85, Revised 1985). The animals were divided into six groups
plasmodial activity (IC50 10 lg/mL); however, Biruksew et al. (I–VI) of five mice each. P. berghei strain NK65 (Chloroquine sen-
(2018) reported the poor antiplasmodial activity elicited by the sitive) was obtained from the Institute of Advanced Medical
methanol root extract of the plant. Nonetheless, neither the Research and Training (IMRAT), University College Hospital,
in vivo antiplasmodial activity of the methanol extract of Z. offi- Ibadan. The donor mouse was euthanized and blood was with-
cinale rhizome nor the computational investigation of the drawn through the cardiac puncture into a heparinized bottle.
inhibitory action of its chemical constituents on some key mal- The blood was diluted such that 0.2 mL of the inoculum con-
aria enzymes has been reported. tained 107 parasitized red blood cells. Each mouse was inocu-
The drug discovery process takes years of effort and huge lated with 0.2 mL of the inoculum via the intraperitoneal route.
investment to accomplish. Drug targets could be enzymes,
receptors, proteins, or other such molecules that are import-
2.3. Preparation of solutions of the extract and
ant to the disease pathophysiology (Mohs and Greig 2017).
standard drug
Computational drug discovery is an economical and faster
method used in drug discovery (Faloye et al. 2021). Appropriate weights of 25, 50, 100, and 200 mg of the
Molecular docking, pharmacophore modeling and mapping, extract of Z. officinale were separately dissolved in 5.0 mL of
de novo design, and molecular simulation are the methods normal saline to give the following doses: 50, 100, 200, and
of computational drug discovery that have found application 400 mg/kg, respectively. The standard drug (chloroquine) was
in target identification and validation, lead discovery and prepared by dissolving 8.33 mg of the chloroquine tablet in
optimization, as well as preclinical tests (Ou-Yang et al. 2012; 5 mL of normal saline to give a dose of 10 mg/kg.
Sliwoski et al. 2014; Schaduangrat et al. 2020).
With the unavailability of an effective vaccine for the pre-
2.4. In vivo antimalarial assays
vention and treatment of malaria and the development of
resistance by P. falciparum to the available antimalarial drugs, The chemosuppressive activities were performed by oral admin-
the search for new drug molecules from natural sources is istration of the extract (50 mg/kg, 100 mg/kg, 200 mg/kg and
inevitable (Plowe et al. 1995; Etim et al. 2018). There is a dire 400 mg/kg) to groups I to IV, and groups V and VI were admin-
need to evaluate the antimalarial claims of medicinal plants istered chloroquine (CQ, 10 mg/kg) and normal saline, respect-
used in folklore medicine for the treatment of malaria. This ively, two hours after infection. These were repeated daily for
study aims to investigate the in vivo antimalarial activity of 3 days (Peters 1965). The level of parasitaemia was determined
the methanol extract of Z. officinale, perform the chemical for each mouse on Day 4 (D4) by taking a thin blood smear
profiling of the methanol extract of Z. officinale, elucidate from the tail of each mouse on a slide. The blood smear was
the inhibitory action of the chemical constituents on some fixed with methanol, stained with Giemsa and counting of para-
enzymes in malaria pathophysiology using molecular docking sitized blood cells in 5 fields of view on the microscope was car-
and molecular dynamics simulation methods, and estimate ried out (Ryley and Peters 1970). The average parasitaemia in
the quantum chemical properties of the hit molecule. each group was calculated in order to determine the percent-
age chemosuppressive activity of the extract using the formula:

2. Materials and methods ðA  BÞ=A  100


where A and B are the mean parasitaemia in the negative
2.1. Plant collection, identification, and extraction
control and the test groups, respectively (Tona et al. 2001).
The rhizome of Z. officinale was collected at the Medicinal The extract’s antimalarial chemosuppressive activity was
Plants Garden, Department of Pharmacognosy, Obafemi determined by the percentage reduction of parasitaemia in
Awolowo University, Ile-Ife, in November 2016. The sample treated groups compared with the untreated infected group.
was identified, and voucher specimens were deposited at
the Forestry Herbarium, Ibadan and the Faculty of Pharmacy
2.5. Survival time
Herbarium, Ife with the codex FHI 110934 and FPI 2356,
respectively. The rhizomes were washed, air-dried, pulverized The animals were further observed for 28 days post-treatment
and 25 g of the plant material was extracted with methanol for mortality to determine the survival time of each mouse. The
(750 mL) for 72 h. The methanol extract was filtered and con- average survival time for each dose was then calculated in
centrated in vacuo to afford the crude extract (3 g). days ± SEM (Adesida, Odediran, and Elujoba 2021).

2.2. Animals and parasite 2.6. Estimation of the median effective doses ED50 and ED90
Thirty Swiss mice of either sex (18–22 g) were obtained from A graph of the test doses in mg/kg was automatically plot-
the Animal House, Faculty of Pharmacy, OAU. They were housed ted against chemosuppression in percentage using Microsoft
JOURNAL OF BIOMOLECULAR STRUCTURE AND DYNAMICS 3

Excel 2007. From the graph, the median effective doses ncbi.nlm.nih.gov/). The ligands were thereafter loaded on the
(ED50 and ED90), which are the doses that would give 50% open babel interface of PyRx 0.8 software, and its energy was
and 90% chemosuppression, were forecast and recorded as minimized under the MMFF94x forcefield and converted to PDB
mg/kg ± SEM. and later to PDBQT for docking purposes (Faloye et al. 2021).

2.7. Statistical analysis 2.10. Validation of docking methodology and molecular


docking
Values were expressed as mean ± SEM and analyzed statistic-
ally using One-way Analysis of Variance (ANOVA) followed by Before molecular docking was performed, the docking pro-
Student Newmann Keul’s post-hoc for comparisons to deter- cedure was validated by re-docking the co-crystallized
mine the source of significant difference for all values. Values ligands in the binding pocket by considering residues with
of p < 0.05 were of statistical significance. 5 Å (Table 1).
To perform docking studies of the identified phytochemi-
cals against the Plasmodium Plasmepsin II (PDBID: ILEE) and
2.8. LC-MS profiling of methanol extract Z. officinale
P. falciparum dihydrofolate reductase-thymidylate synthase
rhizome
(PfDHFR-TS) (PDBID: 1J3I) enzymes, the clean enzymes were
A linear trap quadrupole-(LTQ) Orbitrap spectrometer converted to PDBQT format using the MGL Tool. The pro-
(Thermo Scientific, USA) was used to carry out Liquid chro- teins were then loaded into the Autodock Vina interface of
matography–mass spectrometry (LC–MS) analysis. The instru- PyRx 0.8 software (Trott and Olson 2010), and amino acid
ment is equipped with an Agilent 1200 HPLC system (Santa residues within 5 Å were selected for each enzyme.
Clara, CA, USA), and connected to a photodiode array (PDA) Thereafter, the grid box center and size were adjusted as
detector. Sample preparation was done by making the rhi- appropriate (Table 2). Then the PDBQT format of the ligand
zome extract into a final concentration of 2 mg/mL in metha- was loaded, and docking was performed against each
nol and was centrifuged for 5 min at 6600 rpm and loaded enzyme at an exhaustiveness of 60. After the docking pro-
for analysis. A reverse phase Luna C18 column (60  3 mm, cedure was completed, the nine binding poses for each pro-
3 lm) [Phenomenex, Torrance CA, US], was used to carry out tein-ligand complex were ranked based on their RMSD
HPLC analysis of the sample. The mobile phase consists of values, and the best pose with the lowest RMSD value was
water (þ0.1% formic acid) A and methanol (þ0.1% formic considered the best prediction. The dock poses were
acid) B at a flow rate of 360 lL/min. The gradient was config- imported into the Discovery Studio visualizer and interac-
ured to be a linear gradient from 96% A to 100% B over tions such as hydrogen bonding, pi, and hydrophobic inter-
14 min, followed by 100% B for 4 min, then a return to the
actions were analyzed.
initial concentration of 96% A in 0.6 min, and allowed to
equilibrate for 4.6 min. The column oven condition was kept
at 30  C, and the injection volume was 6 lL. Spectrometry 2.11. Molecular dynamics simulation of protein–ligand
analysis was carried out in positive mode with a nominal complexes
mass resolving power of 60,000 at 400 m/z, spray voltage of
The docked complex of promising compound and control
6 kV, and a scan rate of 1 Hz. The spectrometer was run with
a capillary temperature of (300  C), a tube lens of 100 V, colli- drug were furthermore optimized using computational
sion gas was Argon and Nitrogen as sheath gas (66 arbitrary molecular dynamics simulations in the Desmond module of
units), and auxiliary gas (8 arbitrary units) (Bedane, Zu €hlke, the Schrodinger Maestro 2018.1 program. Firstly, a system
and Spiteller 2020; Adeyoju et al. 2022). Xcalibur software builder tool was used to prepare a simulation system built
2.2.48 was used for data analysis. Compounds were proposed with a TI3P model and a cubic box around the docked com-
by comparison of acquired MS data with literature. plex’s system. The built system was further energy minimized
using a simulated annealing algorithm with 2000 iterations
and a 1.0 kcal/mol/Å convergence limit. Finally, the produc-
2.9. Protein and ligand preparation tion run was done for the 100 ns simulation time, and the
The three-dimensional crystallographic structure of plasmepsin obtained trajectories were analyzed using the simulation
II (PDBID: ILEE) and P. falciparum dihydrofolate reductase-thymi- interaction programme (Saraf et al. 2022)
dylate synthase (PfDHFR-TS) (PDBID: 1J3I) were downloaded
from the protein data bank (www.rcsb.org). To obtain clean 2.12. DFT studies of the hit molecule and chloroquine
proteins suitable for docking, the co-factors, ions and water
molecules were removed. Also, residues interacting with each DFT studies of gingerenone A (hit molecule) and chloroquine
co-crystallized ligand were viewed on PyMol before the co-crys- (standard drug) were performed on Spartan 14 programme
tallized ligand was removed to obtain a bare protein. using B3LYP functional (Lee-Yang-Parr exchange-correlation
Thereafter, the clean proteins were saved in PDB format. The functional method) and 6-31G basis set (Becke 1993). During
three-dimensional conformers of secondary metabolites identi- the calculations, the values of the frontier orbital energies
fied from the rhizome extract of Z. officinale were retrieved in were computed from the most established conformation of
SDF format from the PubChem database (https://pub-chem. the compounds (See supplementary material).
4 K. O. FALOYE ET AL.

Table 1. Enzymes, predicted active residues, and RMSD values.


PDB ID Co-crystallized ligand Active site residues within 5 Å RMSD value
1LEE 4-amino-n-f4-[2-(2,6-dimethyl-phenoxy)- Ile32, Asp34, Gly36, Ser37, Ala38, Met75, Asn76, Tyr77, 1.98
acetylamino]-3-hydroxy-1-isobutyl-5-phenyl- Val78, Ser79, Phe111, Phe120, Ile123, Leu131, Ile133,
pentylg-benzamide Tyr192, Ile212, Asp214, Gly216, Thr217, Ser218,
Ala219, Thr221, Ile290, Leu292, Phe294, Ile300
1J3I 6,6-dimethyl-1-[3-(2,4,5-trichlorophenoxy)propoxy]- Ile14, Cys15, Ala16, Leu46, Trp48, Lys49, Asp54, Met55, 0.96
1,6-dihydro-1,3,5-triazine-2,4-diamine Tyr57, Phe58, Ser108, Ser111, Ile112, Pro113, Phe116,
Leu119, Ile164, Gly165, Tyr170, Thr187

Table 2. Grid box dimension of plasmepsin II and PfDHFR-TS. bark of Plumeria alba Linn (Apocynaceae) cultivated in
Centre Size Nigeria (Adesida, Odediran, and Elujoba 2021).
Enzymes x y z x y z
The average survival time elicited by Z. officinale rhizome
1LEE 32.4490 33.3825 13.6370 28.0068 23.0537 24.1060
extract at all tested doses was comparable to that of the
1J3I 29.9029 9.2940 58.5605 25.5791 25.8596 23.0218 negative control and significantly lower than that of chloro-
quine, the positive control. It has been reported that the sur-
vival of experimental animals beyond 12 days of malarial
2.12.1. ADME studies
infection is regarded as an activity (Ajaiyeoba et al. 2006). It
The ADME studies of gingerenone A were performed by con-
has also been reported that the survival time for treated
sidering descriptors like solubility, blood-brain barrier permeabil-
mice in antiplasmodial studies which doubled that of the
ity, hepatotoxicity, human intestinal absorption, carcinogenicity
negative control, was evidence of good antiplasmodial activ-
and acute oral toxicity using an online server (http://biosig. ities of the extract (Mukherjee 2008; Odediran, Elujoba, and
unimelb.edu.au/pkcsm/prediction) (Awotuya et al. 2022). Adebajo 2014). Hence, in this present study, the near-doubled
survival time of 12 days given by the extract at 50 mg/kg com-
pared with the negative control (6.5 days) could be considered
3. Results and discussion as an indication of good antiplasmodial activity. Moreover, the
3.1. Dose-response study of the methanol extract of Z. extract at 100 mg/kg did not elongate the life span of the mice
officinale rhizome using the 4-day test beyond that of the negative control.
The median effective doses (ED50 and ED90) are the
Medicinal plants have been used from time past in the man- respective doses that will reduce the parasitaemia levels of
agement of different diseases (Zareen et al. 2021). The plant the untreated mice by 50 and 90%, respectively, under
kingdom and the forest have been referred to as “the sleep- standard experimental conditions (McNaught and Wilkinson
ing giant of drug development” and “God’s own pharmacy” 1997). Both values have been used to describe and rank the
(Treben 1986; Verpoorte et al. 1987). The forest needs to be relative antiplasmodial activities of crude extracts, partitioned
further exploited for useful drugs. Both developing and fractions and isolated chemical constituents of medicinal
developed countries utilize traditional herbal remedies in plants (Odediran, Elujoba, and Adebajo 2014). The ED50
their primary health care (Ganesan 2008). Many plants have (164.2546 ± 6.94) and ED90 (265.0975 ± 27.21) values recorded
also been reported for their pharmacological properties for Z. officinale were significantly lower than values recorded
(Aye et al. 2019). This present study evaluates the in vivo from previous studies on Plumeria alba with ED50 and ED90
antiplasmodial activity of the methanol extract of Z. officinale of 305.82 ± 9.99 and 389.74 ± 29.59, respectively (Adesida,
rhizome using the 4-day suppressive test. Percentage chemo- Odediran, and Elujoba 2021), and on Artocarpus altilis with
suppression and survival time were used as parameters in ED50 and ED90 values of 227.17 ± 0.3 and 372.96 ± 0.3,
respectively (Aladesanmi et al. 2022).
the assessment of the activity of the extract.
It can be inferred that the extract elicited good antiplas-
The percentage reduction in parasitaemia at all doses of
modial activity ( 50%) at all doses except at the highest
Z. officinale rhizome extract was significantly lower (p < 0.05)
tested dose (400 mg/kg). This is an indication that increasing
than the negative control which gave 7.15 ± 0.92% and com-
the doses of the extract had no significant effect on its che-
parable to that of chloroquine (2.0005 ± 0.09%). At doses of
mosuppressive activity. On the other hand, 50 mg/kg of Z.
50, 100, 200 and 400 mg/kg BWT, the extract elicited
officinale rhizome extract was found to give optimum anti-
3.237 ± 0.37, 2.8445 ± 0.17, 2.483 ± 0.13 and 3.704 ± 0.56% in plasmodial activity with better survival time than other
parasitaemia, respectively. tested doses.
The extract of Z. officinale gave the highest % chemosup- Secondary metabolites from plants are known to be
pression of 63.64 ± 4.73% elicited at 200 mg/kg was fairly responsible for the medicinal properties of plants (Sofowora
comparable to that of chloroquine (Table 3). This result was 1993). Many secondary metabolites (such as phenolic and
comparable to the chemosuppressive activity of the extract terpene compounds), which might be responsible for the
at 50 mg/kg and 100 mg/kg. This result was significantly pharmacological properties of Z. officinale have been identi-
higher than the value of 32.83 ± 1.03% at 1000 mg/kg, which fied (Mao et al. 2019). Some studies have reported that high
was reported for Z. officinale root extract by Biruksew et al. levels of parasitemia suppression might result from mecha-
(2018). The result was also similar to the earlier results nisms of biomarker actions that indirectly affect the immune
reported on the in vivo antimalarial activities of the stem system (Hymete et al. 2005; Biruksew et al. 2018). In vitro
JOURNAL OF BIOMOLECULAR STRUCTURE AND DYNAMICS 5

Table 3. Average percentage parasitaemia, percentage chemosuppression and average survival time in the in vivo chemo-
suppressive antimalarial activities of methanol extract of Z. officinale rhizome in mice.
Doses (mg/kg) % Parasitaemia ± SEM % Suppression ± SEM Average survival time ± SEM
NC 7.1455 ± 0.92a 0 ± 0.00a 6.5 ± 0.96a
50 3.237 ± 0.37b 52.95 ± 7.46b,c 12.00 ± 0.4a
100 2.8445 ± 0.17b 58.70 ± 4.49b,c 5.50 ± 3.18a
200 2.483 ± 0.13b 63.64 ± 4.73b,c 9.50 ± 0.29a
400 3.704 ± 0.56b 47.88 ± 6.10b 8.00 ± 1.00a
CQ 2.0005 ± 0.09b 70.75 ± 3.56c 24.25 ± 2.18b
ED50 164.2546 ± 6.94
ED90 265.0975 ± 27.21
Keys: Data show the mean ± SEM, n ¼ 5: NC (negative control): Tween 80 in normal saline; CQ ¼ Chloroquine (10 mg/kg).
Only values with different superscripts within columns are significantly different (p < 0.05, one-way analysis of variance fol-
lowed by the Student–Newman–Keul’s post hoc test). ED50 and ED90 are doses of the extracts that gave 50 and 90% activity,
respectively.

Figure 1. LC–MS chromatogram of secondary metabolite profiling of Z. officinale.

antiplasmodial activities of some species in the family, [M þ H]þ, was tentatively identified as 4-methoxybenzalde-
Zingiberaceae have been reported (Biruksew et al. 2018), hyde (Zhong et al. 2022). Entry 1 (m/z 153.0547 [M þ H]þ),
which support the antiplasmodial effect of the methanol consistent with the formula C8H8O3 was identified as vanillin
extract of Z. officinale rhizome reported in this current (Brillatz et al. 2020).
findings.

3.3. Molecular docking analysis


3.2. Profiling of secondary metabolites in the rhizome
The molecular docking procedure validation showed that the
of Z. officinale methanolic extract
RMSD values calculated between the co-crystallized and re-
An LC–MS method was developed, and the chromatogram docked ligands were < 2 Å (Table 1). Also, the structure of
obtained is presented in Figure 1. the co-crystallized and re-docked ligands were almost super-
Seven compounds were tentatively identified from the imposed on each other, thereby confirming the correctness
extract of Z. officinale on the basis of the molecular ions, of the docking procedure adopted for the study (see supple-
mass fragments and literature comparison (Table 4). mentary file (Figure 1))
Ginger is rich in phenolic compounds (Tohma et al. 2017; Molecular docking is widely used for screening and identi-
Ali, El-Nour, and Yagi 2018). Shogaols are a class of phenolic fying drug candidates from synthetic and natural sources. In
compounds commonly found in ginger (Prasad and Tyagi this study, the phytochemicals identified from the metha-
2015; Mao et al. 2019). They are formed via the conversion nolic extract of the rhizome of Z. officinale were docked
of the thermally unstable gingerols (an abundant class of against antimalarial receptors (plasmepsin II and PfDHFR-TS).
phenolic compound in ginger) at high temperatures. Entries Also, chloroquine was selected as the standard drug, and its
4, 5 and 7 eluting at retention times (rt) 23.52, 25.85 and binding energy was set as the cut-off point to identify the
27.62 were tentatively identified as 6-shogaol (m/z 277.1801 hit molecule.
[M þ H]þ), 8-shogaol (m/z 305.2112 [M þ H]þ) and 10-shogaol Gingerenone A was selected as the hit molecule when
(m/z 333.2425 [M þ H]þ), respectively. These compounds the phytochemicals were screened against plasmepsin II and
have been previously isolated and identified in literature their binding energies were compared with chloroquine
(Chen et al. 1986). (6.4 kcal/mol). The hit molecule elicited a considerably high
The peak at rt 16.82 min (entry 2) was tentatively identi- binding energy of 7.0 kcal/mol but was not as high as the
fied as gingerenone A (Endo, Kanno, and Oshima 1990; Peng synthetic co-crystallized ligand (9.6 kcal/mol). It established
et al. 2012). It showed a molecular ion at m/z 357.1698 hydrophobic interactions with Met15, Ile32, Tyr77, Ile123,
[M þ H]þ with the formula consistent with C21H24O5. Entry 6 and Leu131 compared to chloroquine that formed hydropho-
at rt 26.36 min was tentatively identified as the fatty acid, bic interactions with Met15, Ile32, Tyr77, Val78, Ile100, Ile123,
linolenic acid. Linolenic acid was previously identified with Leu292, and Ile300. The hydrogen atom of the hydroxyl
the aid of an HPLC by Lee et al. (2014). The peak at entry 3, group on the aromatic rings of gingerenone A participated
eluting at rt 20.89 with a molecular ion at m/z 137.0598 in hydrogen bonding interaction with Asn76 at a bond
6 K. O. FALOYE ET AL.

distance of 2.18 Å, while hydrogen bonding interaction was organic compounds’ binding affinity and bioactive potential
established between chloroquine and plasmepsin II at a lon- (Chen et al. 2016; Abelian et al. 2021). Hydrogen bonds play
ger bond distance of 2.21 Å (Figure 2a and b). Furthermore, a significant role in protein-drug interactions and are com-
the stability of gingerenone A was strengthened in the bind- monly found between protein-ligand complexes (Grabowski
ing pocket of plasmepsin II by forming pi-alkyl and pi-sigma 2020). Gingerenone A formed a hydrogen bond with the
interactions with Ile32 and Tyr77, while chloroquine formed amino acid residues of plasmepsin II and PfDHFR-TS at a
pi-alkyl interactions with Tyr77 and Val78. Gingerenone A considerable bond distance that is strong enough to
and the co-crystallized ligand formed common hydrogen strengthen its direct interaction with the receptors in con-
bonding interaction with Asp76, hydrophobic interactions trast to chloroquine.
with Ile32, Tyr77, Ile123, Leu131, and pi-interactions with Hydrophobic interactions are established between the
Ile32 and Tyr77. lipophilic groups of chemical compounds and non-polar side
In the molecular docking studies of identified phytochem- chains of amino acids (Chen et al. 2016). It contributes to the
icals and chloroquine against PfDHFR-TS, gingerenone A was binding affinity of chemical compounds and provides
selected as the hit molecule. The compound elicited a bind- insights into their bioactivity potentials (Abelian et al. 2021).
ing energy of 8.0 kcal/mol which is higher than chloroquine The lipophilic atoms on gingerenone A and chloroquine
(7.2 kcal/mol) but lower than that of the co-crystallized ligand formed hydrophobic interactions with non-polar atoms of
(8.9 kcal/mol). An oxygen atom on the gingerenone A moi- the amino acid residues of the receptors’ active site, and
ety formed a hydrogen bonding interaction with Ser111 at a these interactions are good enough for the hit molecule to
bond distance of 2.62 Å, while chloroquine also formed elicit considerably high binding energy than the latter and
hydrogen bonding interaction with Ile164 at a bond distance also reveal it as a good inhibitor against the receptors. The pi-
of 3.02 Å. Gingerenone A also established hydrophobic inter- interactions are established when pi-electron cloud establishes
action with Ala16, Leu40, Leu46, and Phe58, while chloro- interaction between the aromatic group and pi-electron of phy-
quine had five hydrophobic interactions with Cys15, Leu40, tochemicals to facilitate their stability in the receptor’s binding
Phe58, Ile112, Leu119, and Ile164. Furthermore, pi-alkyl inter- site (Mousavi et al. 2021). Both gingerenone A and chloroquine
action was observed between gingerenone A-Leu40 and have aromatic ring systems and formed pi-interaction at the
chloroquine-Phe58 and Ile164 (Figure 3a and b). The hit mol- receptors’ binding sites.
ecule established only two hydrophobic interactions with A similar study conducted on quercetin against plasmepsin
Ile112 and Leu119. II showed that gingerenone elicited better binding energy than
Non-covalent interactions and chemical bonding is the the former which gave 6.36 kcal/mol (Hasan et al. 2022). In
major route through which chemical compounds binds to another study conducted by Olanlokun et al. (2019) whereby a
receptors (Grabowski 2020). Hydrogen bonding and hydro-
novel phytochemical (tetrahydro-4-((E)-7-hydroxy-10-methoxy-6,
phobic interactions are major determinants in evaluating
14-dimethyl-15-m-tolylpentadec-13-enyl) pyran-2-one) was
screened against plasmepsin II, a binding energy comparable
Table 4. Secondary metabolite profiling of extract of Z. officinale.
to gingerenone A was obtained. Also, Kyei et al. (2022)
m/z [Adduct]
Entry Rt (min) [M þ H]þ Formula Name screened pyrimethamine against PfDHFR-TS and obtained a
1 10.96 153.0547 C8H8O3 Vanillin binding energy of 7.8 kcal/mol which is lower than that of
2 16.82 357.1698 C21H24O5 Gingerenone A gingerenone A. Also, the binding energy of gingerenone A
3 20.89 137.0598 C8H9O2 4-Methoxybenzaldehyde against PfDHFR-TS is higher than values reported for lascu-
4 23.52 277.1801 C17H24O3 6-Shogaol
5 25.85 305.2112 C19H28O3 8-Shogaol floxacin (6.59 kcal/mol) and moxifloxacin (5.65 kcal/mol)
6 26.36 279.2320 C18H30O2 a-Linolenic acid (Balogun, Omoboyowa, and Saibu 2020). Gingerenone A has
7 27.62 333.2425 C21H32O3 10-Shogaol
previously been reported to possess interesting biological

Figure 2. Interaction diagrams of the docked complexes of (a) ILEE-gingerenone A and (b) ILEE-chloroquine.
JOURNAL OF BIOMOLECULAR STRUCTURE AND DYNAMICS 7

Figure 3. Interaction diagrams of the docked complexes of (a) IJ3I-gingerenone A and (b) IJ3I-chloroquine.

activities, such that it attenuated TNF-a, LPS-induced mono- the protein fluctuated slightly between 0.6 and 3.4 Å, while
cyte-endothelial adhesion, and inhibited Staphylococcus aur- that of chloroquine-1LEE ranged between 0.5 and 3.2 Å in
eus SaHPPK in silico (Kim et al. 2018; Rampogu et al. 2018). the 100 ns gingerenone A-ILEE simulation period (Figure 4c
Other studies revealed gingerenone A as an excellent thera- and d). Also, the contact plot showed good hydrogen bonds,
peutic agent with strong antimicrobial, anticancer and anti- hydrophobic interactions and water bridge. Gingerenone A
diabetic activities (Chen et al. 2018; Rampogu et al. 2018; Yu participated in hydrogen bonding interaction with Ala38,
et al. 2022). The binding energy obtained for gingerenone A Asn39, Val78, Ser79, Thr114, and Ile133, while hydrophobic
as compared to chloroquine in the molecular docking studies interaction was observed with Tyr41, Tyr77, Phe111, Ile123,
against the malaria enzymes is a strong indication that the and Ile133. Furthermore, a water bridge was formed with
hit molecule may be responsible for the antimalarial activity Asp34, Val78 and Ser79 (Figure 4e). Comparing the contact
elicited by the methanolic extract of Z. officinale. plot of chloroquine with gingerenone A, chloroquine had
hydrogen bonding interaction with Asp34, Gln142, Asp162,
hydrophobic interactions with Tyr77, Phe111, Phe241 and
3.4. Molecular dynamics simulation analysis
water bridge with Asp214 (Figure 4f). Overall, the RMSD plot
Molecular dynamics simulation is a vital study to investigate showed that gingerenone A was more stable than chloro-
the structural dynamics, conformation, interactions and sta- quine in the binding pocket of plasmepsin II. Also, common
bility of protein-ligand complexes (Yadav, Tripathi, and Yadav hydrophobic interactions were established with Tyr77 and
2021; Iqbal et al. 2022). In this study, the 1LEE-gingerenone Phe111.
A and 1J3I-gingerenone A complexes were subjected to a The 1J31-gingerenone A complex RMSD increased for
100 ns simulation study to gain better insight into any about 10 ns followed by stability between 1.6 and 2.6 Å
change in the protein conformational study that may arise throughout the simulation period. Also, the ligand RMSD for
from the gingerenone A binding. The stability of gingere- the gingerenone A fitted on the 1J3I backbone showed ini-
none A in 1LEE-gingerenone and 1J3I-gingerenone A com- tial stability from 0 to 9 ns followed by a sharp increase to
plexes was estimated by considering their RMSD (root mean 15 ns. Stability was observed between 16 and 43 ns followed
square deviation) and RMSF (root mean square fluctuations). by a sharp increase and fluctuation between 44 and 55 ns
The protein RMSD plot of the Ca backbone (blue colour) and stability for the remaining simulation period, while initial
of 1LEE and reference frame backbone obtained for the deviation was observed for chloroquine and the system was
100 ns simulation is presented in Figure 4. The RMSD plot stable till 90 ns followed by a slight deviation to 100 ns
showed an initial increase from 0 to 5 ns followed by a stable (Figure 5a and b). In the RMSF plot, slight fluctuation was
system from 6 to 60 ns and a slope that ranged between 1.7 observed with the protein around 0.6–4.8 Å, while fluctua-
and 2.4 Å, while the Ca backbone of chloroquine-1LEE com- tions were observed for the protein in the chloroquine-1J3I
plex was stable up to 63 ns followed by a slight deviation simulation (Figure 5c and d).
and stability was observed to 100 ns. In the ligand RMSD Furthermore, the protein–ligand interaction chart showed
plot for the gingerenone A fitted on 1LEE backbone, serial the presence of hydrogen bonds, hydrophobic interactions,
fluctuation was observed between 0 and 40 ns followed by a and water bridge. In the contact plot, gingerenone A estab-
stable system 41–90 ns and a sharp increase that range lished hydrogen bonding interaction with Ala16, Cys50,
between 9 and13 Å, while chloroquine was stable up to Asp54, Ser111, and Tyr185, while hydrophobic interaction
28 ns followed by consistent deviation to 100 ns (Figure 4a was observed between the phytochemical and Ala16, Leu40,
and b). The RMSF plot of gingerenone A-1LEE showed that Leu46, Met55, Phe58, Ile112, and Pro113. Additionally, water
8 K. O. FALOYE ET AL.

Figure 4. Illustration diagram of (a) RMSD of ILEE-gingerenone A, (b) RMSD of ILEE-chloroquine, (c) RMSF of ILEE-gingerenone A, (d) RMSF of ILEE-chloroquine, (e)
contact plot of ILEE-gingerenone A and (f) contact plot of ILEE-chloroquine.

bridge interaction was formed with Gly44, Tyr48, Lys49, and also corroborated the findings in the molecular docking
Cys50, Asp54, Thr107, Ser111, Arg122, and Ile164 (Figure 5e), studies.
while chloroquine established interactions with fewer amino
acids (hydrogen bonding interactions with Asp54, hydropho-
3.5. MMGBSA analysis
bic interactions with Phe58, Leu46 and water bridge with
Ile164) (Figure 5f). The MMGBSA is a less expensive computational method of
The RMSD and RMSF plots help to understand the flexibil- estimating the binding affinity of docked protein-ligand com-
ity of amino acid residues and the stability of ligands in the plexes compared to thermodynamic integration and free
protein’s binding pocket (Mousavi et al. 2021). Also, interac- energy perturbation methods (Iqbal et al. 2022).
tions like hydrogen bonding, hydrophobic, and water bridges In this study, the MMGBSA results showed that gingere-
contribute to the stability and binding energy elicited by none A had good binding free energy with plasmepsin II
ligands (Yadav, Tripathi, and Yadav 2021). Gingerenone A (37.07 kcal/mol) and PfDHFR-TS (-49.09 kcal/mol) (Table 5).
was more stable than chloroquine and established good Also, the covalent, Van der Waal, solvation, and coulomb
interactions with the residues at the proteins’ active sites energy value computed for the ILEE-gingerenone A and 1J3I-
JOURNAL OF BIOMOLECULAR STRUCTURE AND DYNAMICS 9

Figure 4. Continued.

gingerenone A complex contributed to the binding and sta- study, chloroquine has a higher EHOMO value (5.49 eV) as
bility of the ligand throughout the simulation period. compared to gingerenone A (5.35 eV), indicating that
chloroquine has a greater electron donating property than
gingerenone A. Conversely, gingerenone A (1.50 eV) has a
3.6. Frontier molecular orbital analysis of the hit
better electron withdrawing property than chloroquine
molecule and standard drug
(1.04 eV). The EHOMO and ELUMO values obtained were as a
The frontier molecular orbital analysis of phytochemicals result of the electric charges delocalized over the p-network
gives useful information about their drug candidacy through and aromatic ring system of gingerenone A and chloroquine
the estimation of their physiochemical properties like highest (Janjua et al. 2014).
occupied molecular orbital (HOMO), lowest unoccupied The energy gap of a molecule is the excitation energy
molecular orbital (LUMO), energy gap, hardness, softness, required to cause an electron transfer. It is used to study the
electrophilicity index, and electronegativity. chemical reactivity, stability, and bioactive properties of sec-
The EHOMO of a chemical compound is a measure of its ondary metabolites (Subramanian, Sundaraganesan, and
electron donating potential while the ELUMO is used to meas- Jayabharathi 2010; Obi-Egbedi et al. 2011). Chloroquine
ure its electron accepting property. Therefore, the higher the (4.45 eV) has a higher energy gap as compared to gingere-
EHOMO value, the greater the electron donating tendency of none A (3.85 eV), signifying that chloroquine is more stable
the molecule, while a higher ELUMO value indicates a higher but less reactive, at the active site of the antimalarial
electron accepting tendency (Bhavani et al. 2015). In this enzymes (Figure 6). Also, the decreased energy gap of
10 K. O. FALOYE ET AL.

Figure 5. Illustration diagram of (a) RMSD of 1J3I-gingerenone A, (b) RMSD of 1J3I-chloroquine, (c) RMSF of 1J3I-gingerenone A, (d) RMSF of 1J3I-chloroquine, (e)
contact plot of 1J3I-gingerenone A and (f) contact plot of 1J3I-chloroquine.

gingerenone A indicates that it will undergo easy excitation charges at the binding pocket of enzymes (Bulbul et al.
of electrons from the ground state (Raheem, Al-Shejyri, and 2021; Owoseeni et al. 2022). Therefore, they are essential
Al-Bermany 2012). electronic parameters in predicting the pharmacological
The chemical hardness and softness of chemical com- potential of secondary metabolites. Gingerenone A (1.93 eV;
pounds are important in accurately predicting their stability 0.52 eV) has close hardness and softness values to
and reactivity (Ayeni et al. 2020). Chemical compounds with chloroquine (2.23 eV; 0.45 eV). Hence, it could be said that
a lower hardness value have a better tendency to exchange gingerenone is a promising antimalarial agent. Additionally,
JOURNAL OF BIOMOLECULAR STRUCTURE AND DYNAMICS 11

Figure 5. Continued.
12 K. O. FALOYE ET AL.

chemical potential helps to predict the reactivity and stability 3.7. ADMET prediction
of secondary metabolites. The higher the chemical potential
The ADMET properties of gingerenone A were evaluated to
of a molecule, the higher its reactivity and the lower its sta-
understand its metabolism and safety when taken by malaria
bility (Subramanian et al. 2010; Awotuya et al. 2022).
subjects. The solubility and human intestinal absorption
The electronegativity and electrophilicity index of chloro-
property of a drug candidate provide information on the
quine and gingerenone A were also calculated. The electro-
ease of formulation, mode of administration and rate of
negativity of a chemical compound is a measure of its
absorption (Ogboye et al. 2022). Blood-brain barrier (BBB)
electron withdrawing property. It helps in evaluating the bio-
permeability property of a drug candidate measures its abil-
logical property of secondary metabolites (Domingo et al.
ity to cross the BBB layer without any hindrance and it is
2016; Hussein and Azeez 2021). Gingerenone A (3.44 eV) has
important for central nervous system (CNS) drug design
a better electron withdrawing property due to the presence
(Sharma et al. 2020). The carcinogenicity property of a drug
of more electronegative elements as compared to chloro-
candidate elucidates its ability to either speed up the forma-
quine (3.27 eV). The electrophilicity index is directly related
tion of cancerous cells or not, while hepatotoxicity study
to the binding affinity elicited by a secondary metabolite at
helps to predict the possible toxicity of the chemical com-
the active site of a receptor. It also measures the stability of
pound to the liver (Owoseeni et al. 2022). The acute oral tox-
a chemical compound when it obtains more electron density
icity studies of a drug candidate provide information about
from its environment (Domingo et al. 2016). A strong electro-
its safety when taken orally by malarial subjects.
phile possesses a high electrophilicity index value and high
In this study, gingerenone A gave a solubility value of
binding energy against a receptor (Mumit et al. 2020;
4.66 and a human intestinal absorption value of 93%. A
Olajubutu et al. 2022). Gingerenone A (3.16 eV) had a higher
drug candidate with good absorptivity, when administered,
electrophilicity index than chloroquine (2.39 eV) which also
must have a solubility value ranging between 6.5 and 0.5,
corroborated the binding energy elicited by the hit molecule
while human intestinal absorption should be more than 30%
in the molecular docking and MMGBSA studies.
(Sharma et al. 2020; Awotuya et al. 2022). The hit molecule
Table 5. MMGBSA DG binding scores of the top-ranked molecules.
also showed good a BBB permeability property, signifying
that it is capable of acting as a good CNS drug. In terms of
DG bind
DG bind DG bind solvation DG bind toxicity, gingerenone A is a non-carcinogenic compound and
P-L complex DG bind coulomb covalent energy vdW not toxic to the liver. The oral acute and chronic toxicity
1LEE-gingerenone A 37.07 24.98 8.05 41.78 38.10 studies showed that gingerenone A is not toxic when admin-
1J3I-gingerenone A 49.09 8.84 7.20 33.53 52.51
istered orally to malaria subjects. The overall toxicity profile

Figure 6. Illustration diagram of the HOMO and LUMO of (a) gingerenone A and (b) chloroquine.
JOURNAL OF BIOMOLECULAR STRUCTURE AND DYNAMICS 13

of the hit molecule shows it is safe for further clinical and Funding
pharmacokinetic investigations.
The authors did not receive any external funding.

4. Conclusion ORCID
This study evaluated the antimalarial activity of the methanolic Adetola H. Adewole http://orcid.org/0000-0003-4983-245X
extract of Z. officinale rhizome, identified the phytoconstituents
in the extract, performed molecular docking and molecular
dynamics simulation studies against plasmepsin II and PfDHFR- References
TS receptors to identify the hit molecule that may be respon- Abelian, A., Dybek, M., Wallach, J., Gaye, B., & Adejare, A. (2021).
sible for the antimalarial activity. It also carried out density Pharmaceutical chemistry. In Remington (pp. 105–128). Academic Press.
functional theory calculations to obtain the electronic properties Adesida, S. A., Odediran, S. A., & Elujoba, A. A. (2021). Investigation on
of the hit molecule and predicted the ADMET profile to esti- the antimalarial properties of Plumeria alba Linn (apocynaceae) culti-
vated in Nigeria. Nigerian Journal of Natural Products and Medicine,
mate the safety and drug candidacy of the top-ranked phyto- 25(1), 34–42. https://doi.org/10.4314/njnpm.v25i1.2
chemical. The results obtained showed that the extract elicited Adeyoju, E. O., Ajayi, C. O., Adepiti, A. O., & Elujoba, A. A. (2022).
a dose-dependent chemosuppressive activity up to 200 mg/kg. Comparative in vivo antimalarial activities of aqueous and methanol
At the lowest dose of 50 mg/kg, the extract gave antiplasmodial extracts of MAMA powder-A herbal antimalarial preparation. Journal of
Ethnopharmacology, 283, 114686. https://doi.org/10.1016/j.jep.2021.114686
activity >50% and the highest survival time of 12 days. The
Ajaiyeoba, E. O., Abiodun, O. O., Falade, M. O., Ogbole, N. O., Ashidi, J. S.,
highest antimalarial activity of 63% elicited by the methanolic Happi, C. T., & Akinboye, D. O. (2006). In vitro cytotoxicity studies of
extract at 200 mg/kg confirms the usage of Z. officinale rhizome 20 plants used in Nigerian antimalarial ethnomedicine. Phytomedicine
in ethnomedicine. The LC-MS secondary metabolite profiling : International Journal of Phytotherapy and Phytopharmacology, 13(4),
identified vanillin, gingerenone A, 4-methoxybenzaldehyde, 6- 295–298. https://doi.org/10.1016/j.phymed.2005.01.015
Aladesanmi, A. J., Odiba, O. E., Odediran, S. A., & Oriola, A. O. (2022).
shagaol, 8-shogaol, 10-shogaol and a-linolenic acid as the
Antiplasmodial activities of the stem bark extract of Artocarpus altilis
chemical constituents present in the methanolic extract of the Forsberg. African Journal of Infectious Diseases, 16(2S), 33–45. https://
Z. officinale rhizome. The molecular docking study identified doi.org/10.21010/Ajidv16i2S.5
gingerenone A as the hit molecule against plasmepsin II and Ali, A. M. A., El-Nour, M. E. M., & Yagi, S. M. (2018). Total phenolic and
PfDHFR-TS with a binding energy higher than that of chloro- flavonoid contents and antioxidant activity of ginger (Zingiber offici-
nale Rosc.) rhizome, callus and callus treated with some elicitors.
quine. Gingerenone A established good hydrophobic, hydrogen
Journal, Genetic Engineering & Biotechnology, 16(2), 677–682. https://
bonding and pi-interactions which contributed significantly to doi.org/10.1016/j.jgeb.2018.03.003
its good binding energy and stability at the enzymes’ binding Awotuya, I. O., Fakola, E. G., Olusola, A. J., Olanudun, E. A., Bello, O. I.,
pocket. The RMSD plots obtained in the 100 ns molecular Ogunremi, B. I., Gboyero, F. O., Adesida, S. A., & Faloye, K. O. (2022).
dynamics simulation showed that gingerenone A exhibited bet- Exploring the protein tyrosine phosphatase 1B inhibitory potentials of
naturally occurring Brazilin-type homoisoflavonoids: A computational
ter stability at the binding pocket of plasmepsin II and PfDHFR-
approach. Chemistry Africa, 5(5), 1493–1502. https://doi.org/10.1007/
TS receptors, while the RMSF plots suggested slight fluctuations s42250-022-00415-3
were observed with the amino acid residues of the malaria Aye, M., M., Aung, H. T., Sein, M. M., & Armijos, C. (2019). A review on
enzymes. The contact plots showed that gingerenone A con- the phytochemistry, medicinal properties and pharmacological activ-
solidated its stability during the 100 ns molecular dynamics ities of 15 selected Myanmar medicinal plants. Molecules, 24(2), 293.
https://doi.org/10.3390/molecules24020293
simulation by forming water bridge, hydrogen bonding and Ayeni, A. O., Akinyele, O. F., Hosten, E. C., Fakola, E. G., Olalere, J. T.,
hydrophobic interactions. Also, the electrophilicity index and Egharevba, G. O., & Watkins, G. M. (2020). Synthesis, crystal structure,
HOMO-LUMO energy gap values obtained from density func- experimental and theoretical studies of corrosion inhibition of 2-((4-
tional theory calculations for gingerenone A supported the (2-hydroxy-4-methylbenzyl) piperazin-1-yl) methyl)-5-methylphenol–A
binding energy elicited by gingerenone A against plasmepsin II Mannich base. Journal of Molecular Structure, 1219, 128539. https://
doi.org/10.1016/j.molstruc.2020.128539
and PfDHFR-TS enzymes. In addition, the toxicity profile gener- Balogun, T. A., Omoboyowa, D. A., & Saibu, O. A. (2020). In silico anti-
ated from the ADMET profile showed that gingerenone A has malaria activity of quinolone compounds against Plasmodium falcip-
good solubility, human intestinal absorption properties and safe arum dihydrofolate reductase (pfDHFR). International Journal of
for malaria subjects. The isolation and extensive antimalarial Biochemistry Research & Review, 29(8), 10–17.
Becke, A. D. (1993). A new mixing of Hartree–Fock and local density-
studies of gingerenone A is thereby recommended.
functional theories. The Journal of Chemical Physics, 98(2), 1372–1377.
https://doi.org/10.1063/1.464304
Bedane, K. G., Z€ uhlke, S., & Spiteller, M. (2020). Bioactive constituents of
Acknowledgement Lobostemon fruticosus: Anti-inflammatory properties and quantitative
S.A.O, K.O.F, U.I.O, I.I.O and O.D.D are thankful to Ms. Eva Wieczorek analysis of samples from different places in South Africa. South African
(Institute of food and environmental research, Technical University of Journal of Botany, 131, 174–180. https://doi.org/10.1016/j.sajb.2020.02.016
Dortmund) for her assistance with LC-MS data acquisition. Bhaumik, P., Gustchina, A., & Wlodawer, A. (2012). Structural studies of
vacuolar plasmepsins. Biochimica et biophysica acta, 1824(1), 207–223.
https://doi.org/10.1016/j.bbapap.2011.04.008
Disclosure statement Bhavani, K., Renuga, S., Muthu, S., & Sankara Narayanan, K. (2015).
Quantum mechanical study and spectroscopic (FT-IR, FT-Raman, 13C,
The authors declare that they have no conflicts of interest with the con- 1H) study, first order hyperpolarizability, NBO analysis, HOMO and
tents of this article. LUMO analysis of 2-acetoxybenzoic acid by density functional
14 K. O. FALOYE ET AL.

methods. Spectrochimica Acta Part A: Molecular and Biomolecular analysis of Quercetin compound with its evaluation of broad-spec-
Spectroscopy, 136, 1260–1268. https://doi.org/10.1016/j.saa.2014.10.012 trum therapeutic potential against particular diseases. Informatics in
Bhowmik, D., Tripathi, K. K., Chandira, M. R., & Kumar, K. P. (2010). Medicine Unlocked, 29, 100894. https://doi.org/10.1016/j.imu.2022.
Zingiber officinale the herbal and traditional medicine and its thera- 100894
peutically importance. Research Journal of Pharmacognosy and Hymete, A., Iversen, T. H., Rohloff, J., & Erko, B. (2005). Screening of
Phytochemistry, 2(2), 102–110. Echinops ellenbeckii and Echinops longisetus for biological activities
Biruksew, A., Zeynudin, A., Alemu, Y., Golassa, L., Yohannes, M., Debella, and chemical constituents. Phytomedicine : international Journal of
A., Urge, G., De Spiegeleer, B., & Suleman, S. (2018). Zingiber Phytotherapy and Phytopharmacology, 12(9), 675–679. https://doi.org/
Officinale Roscoe and Echinops Kebericho Mesfin showed antiplasmo- 10.1016/j.phymed.2004.01.013
dial activities against Plasmodium berghei in a dosedependent man- Hussein, Y. T., & Azeez, Y. H. (2021). DFT analysis and in silico exploration
ner in Ethiopia. Ethiopian Journal of Health Sciences, 28(5). https://doi. of drug-likeness, toxicity prediction, bioactivity score, and chemical
org/10.4314/ejhs.v28i5.17 reactivity properties of the urolithins. Journal of Biomolecular Structure
Brillatz, T., Kubo, M., Takahashi, S., Jozukuri, N., Takechi, K., Queiroz, E. F., and Dynamics, 51(10), 640–653.
Marcourt, L., Allard, P.-M., Fish, R., Harada, K., Ishizawa, K., Crawford, Iqbal, S., Potharaju, R., Naveen, S., Lokanath, N. K., Mohanakrishnan, A. K.,
A. D., Fukuyama, Y., & Wolfender, J.-L. (2020). Metabolite profiling of & Gunasekaran, K. (2022). Design, crystal structure determination,
Javanese ginger Zingiber purpureum and identification of antiseizure molecular dynamic simulation and MMGBSA calculations of novel
metabolites via a low-cost open-source Zebrafish bioassay-guided iso- p38-alpha MAPK inhibitors for combating Alzheimer’s disease. Journal
lation. Journal of Agricultural and Food Chemistry, 68(30), 7904–7915. of Biomolecular Structure & Dynamics, 40(13), 6114–6127. https://doi.
https://doi.org/10.1021/acs.jafc.0c02641 org/10.1080/07391102.2021.1877197
Bulbul, M. Z., Hosen, M. A., Ferdous, J., Chowdhury, T. S., Misbah, M. M., Janjua, M. R. S. A., Mahmood, A., Nazar, M. F., Yang, Z., & Pan, S. (2014).
& Kawsar, S. (2021). DFT study, physicochemical, molecular docking, Electronic absorption spectra and nonlinear optical properties of ruthe-
and ADMET predictions of some modified uridine derivatives. nium acetylide complexes: A DFT study toward the designing of new
International Journal of New Chemistry, 8(1), 88–110. high NLO response compounds. Acta Chimica Slovenica, 61(2), 382–390.
Cator, L. J., Lynch, P. A., Read, A. F., & Thomas, M. B. (2012). Do malaria Jensen, A. R., Adams, Y., & Hviid, L. (2020). Cerebral Plasmodium falcip-
parasites manipulate mosquitoes? Trends in Parasitology, 28(11), 466– arum malaria: The role of PfEMP1 in its pathogenesis and immunity,
470. https://doi.org/10.1016/j.pt.2012.08.004 and PfEMP1-based vaccines to prevent it. Immunological Reviews,
Chen, C. C., Kuo, M. C., Wu, C. M., & Ho, C. T. (1986). Pungent com- 293(1), 230–252. https://doi.org/10.1111/imr.12807
pounds of ginger (Zingiber officinale Roscoe) extracted by liquid car- Kaushik, N. K., Bagavan, A., Rahuman, A. A., Mohanakrishnan, D., Kamaraj,
bon dioxide. Journal of Agricultural and Food Chemistry, 34(3), 477– C., Elango, G., Zahir, A. A., & Sahal, D. (2013). Antiplasmodial potential
480. https://doi.org/10.1021/jf00069a027 of selected medicinal plants from eastern Ghats of South India.
Chen, J., Sun, J., Prinz, R. A., Li, Y., & Xu, X. (2018). Gingerenone A sensi- Experimental Parasitology, 134(1), 26–32. https://doi.org/10.1016/j.
tizes the insulin receptor and increases glucose uptake by inhibiting exppara.2013.01.021
the activity of p70 S6 kinase. Molecular Nutrition & Food Research, Kim, H. J., Son, J. E., Kim, J. H., Lee, C. C., Yang, H., Yaghmoor, S. S.,
62(23), 1800709. https://doi.org/10.1002/mnfr.201800709 Ahmed, Y., Yousef, J. M., Abualnaja, K. O., Al-Malki, A. L., Kumosani,
Chen, D., Oezguen, N., Urvil, P., Ferguson, C., Dann, S. M., & Savidge, T. A., Kim, J. H., Yoon Park, J. H., Lee, C. Y., Kim, J.-E., & Lee, K. W.
T. C. (2016). Regulation of protein-ligand binding affinity by hydrogen (2018). Gingerenone A attenuates monocyte-endothelial adhesion via
bond pairing. Science Advances, 2(3), e1501240. https://doi.org/10. suppression of I Kappa B kinase phosphorylation. Journal of Cellular
1126/sciadv.1501240 Biochemistry, 119(1), 260–268. https://doi.org/10.1002/jcb.26138
Chugh, M., Sundararaman, V., Kumar, S., Reddy, V. S., Siddiqui, W. A., Kyei, L. K., Gasu, E. N., Ampomah, G. B., Mensah, J. O., & Borquaye, L. S.
Stuart, K. D., & Malhotra, P. (2013). Protein complex directs hemoglo- (2022). An in silico study of the interactions of alkaloids from
bin-to-hemozoin formation in Plasmodium falciparum. Proceedings of Cryptolepis sanguinolenta with Plasmodium falciparum dihydrofolate
the National Academy of Sciences of the United States of America, reductase and dihydroorotate dehydrogenase. Journal of Chemistry,
110(14), 5392–5397. https://doi.org/10.1073/pnas.1218412110 2022, 1–26. https://doi.org/10.1155/2022/5314179
Domingo, L. R., Ríos-Gutiérrez, M., & Pérez, P. (2016). Applications of the Lee, H. R., Lee, J. H., Park, C. S., Ra, K. R., Ha, J. S., Cha, M. H., Kim, S. N.,
conceptual density functional theory indices to organic chemistry Choi, Y., Hwang, J., & Nam, J. S. (2014). Physicochemical properties
reactivity. Molecules, 21(6), 748. https://doi.org/10.3390/ and antioxidant capacities of different parts of ginger (Zingiber offici-
molecules21060748 nale Roscoe). Journal of the Korean Society of Food Science and
Endo, K., Kanno, E., & Oshima, Y. (1990). Structures of antifungal diary- Nutrition, 43(9), 1369–1379. https://doi.org/10.3746/jkfn.2014.43.9.1369
lheptenones, gingerenones A, B, C and isogingerenone B, isolated Liu, P. (2017). Plasmepsin: Function, characterization and targeted anti-
from the rhizomes of Zingiber officinale. Phytochemistry, 29(3), 797– malarial drug development. In Natural remedies in the fight against
799. https://doi.org/10.1016/0031-9422(90)80021-8 parasites (183–218).
Etim, O. E., Bassey, U. E., Akpakpan, E. I., & Udofia, I. E. (2018). Mao, Q. Q., Xu, X. Y., Cao, S. Y., Gan, R. Y., Corke, H., Beta, T., & Li, H. B.
Prophylactic, Suppressive and Curative Antiplasmodial Potentials of (2019). Bioactive compounds and bioactivities of ginger (Zingiber offi-
Methanol Root Extract of Napoleona imperialis in Plasmodium berghei cinale Roscoe). Foods, 8(6), 185. https://doi.org/10.3390/foods8060185
berghei Infected Male Albino Mice. IOSR Journal of Pharmacy and McNaught, A. D., & Wilkinson, A. (1997). IUPAC: Compendium of chemical
Biological Sciences (IOSR-JPBS), 13(1), 73–77. terminology. 2nd ed. "Gold Book" Blackwell Scientific Publications.
Faloye, K. O., Bekono, B. D., Fakola, E. G., Ayoola, M. D., Bello, O. I., Mohs, R. C., & Greig, N. H. (2017). Drug discovery and development: Role
Olajubutu, O. G., Owoseeni, O. D., Mahmud, S., Alqarni, M., Al Awadh, of basic biological research. Alzheimer’s & Dementia (New York, N. Y.),
A. A., Alshahrani, M. M., & Obaidullah, A. J. (2021). Elucidating the glu- 3(4), 651–657. https://doi.org/10.1016/j.trci.2017.10.005
cokinase activating potentials of naturally occurring prenylated flavo- Mousavi, S. S., Karami, A., Haghighi, T. M., Tumilaar, S. G., Idroes, R.,
noids: An explicit computational approach. Molecules, 26(23), 7211. Mahmud, S., Celik, I., Ag  ag€und€uz, D., Tallei, T. E., Emran, T. B.,
https://doi.org/10.3390/molecules26237211 Capasso, & R. Fatimawali. (2021). In silico evaluation of Iranian medi-
Ganesan, A. (2008). The impact of natural products upon modern drug cinal plant phytoconstituents as inhibitors against main protease and
discovery. Current Opinion in Chemical Biology, 12(3), 306–317. https:// the receptor-binding domain of SARS-CoV-2. Molecules, 26(18), 5724.,
doi.org/10.1016/j.cbpa.2008.03.016 https://doi.org/10.3390/molecules26185724
Grabowski, S. J. (2020). Intramolecular hydrogen bond energy and its Mukherjee, P., Pradhan, A., Shah, F., Tekwani, B. L., & Avery, M. A. (2008).
decomposition—O–H O interactions. Crystals, 11(1), 5. https://doi. Structural insights into the Plasmodium falciparum histone deacety-
org/10.3390/cryst11010005 lase 1 (PfHDAC-1): A novel target for the development of antimalarial
Hasan, M. M., Khan, Z., Chowdhury, M. S., Khan, M. A., Moni, M. A., & therapy. Bioorganic & Medicinal Chemistry, 16(9), 5254–5265. https://
Rahman, M. H. (2022). In silico molecular docking and ADME/T doi.org/10.1016/j.bmc.2008.03.005
JOURNAL OF BIOMOLECULAR STRUCTURE AND DYNAMICS 15

Mumit, M. A., Pal, T. K., Alam, M. A., Islam, M. A. A. A. A., Paul, S., & Rudrapal, M., & Chetia, D. (2017). Plant flavonoids as potential source of
Sheikh, M. C. (2020). DFT studies on vibrational and electronic spectra, future antimalarial leads. Systematic Reviews in Pharmacy, 8(1), 13–18.
HOMO–LUMO, MEP, HOMA, NBO and molecular docking analysis of https://doi.org/10.5530/srp.2017.1.4
benzyl-3-N-(2, 4, 5-trimethoxyphenylmethylene) hydrazinecarbodi- Ryley, J. F., & Peters, W. (1970). The antimalarial activity of some quin-
thioate. Journal of Molecular Structure, 1220, 128715. https://doi.org/ olone esters. Annals of Tropical Medicine and Parasitology, 64(2), 209–
10.1016/j.molstruc.2020.128715 222. https://doi.org/10.1080/00034983.1970.11686683
Obi-Egbedi, N. O., Essien, K. E., Obot, I. B., & Ebenso, E. E. (2011). 1, 2- Saraf, P., Tripathi, P. N., Tripathi, M. K., Tripathi, A., Verma, H., Waiker,
Diaminoanthraquinone as corrosion inhibitor for mild steel in hydro- D. K., Singh, R., & Shrivastava, S. K. (2022). Novel 5, 6-diphenyl-1, 2, 4-
chloric acid: Weight loss and quantum chemical study. International triazine-3-thiol derivatives as dual COX-2/5-LOX inhibitors devoid of
Journal of Electrochemical Science. 6, 913–930. cardiotoxicity. Bioorganic Chemistry, 129, 106147. https://doi.org/10.
Odediran, S. A., Elujoba, A. A., & Adebajo, A. C. (2014). Influence of for- 1016/j.bioorg.2022.106147
mulation ratio of the plant components on the antimalarial properties Schaduangrat, N., Lampa, S., Simeon, S., Gleeson, M. P., Spjuth, O., &
of MAMA decoction. Parasitology Research, 113(5), 1977–1984. https:// Nantasenamat, C. (2020). Towards reproducible computational drug
discovery. Journal of Cheminformatics, 12(1), 30. https://doi.org/10.
doi.org/10.1007/s00436-014-3848-2
1186/s13321-020-0408-x
Olajubutu, O. G., Ogunremi, B. I., Adewole, A. H., Awotuya, O. I., Fakola, E. G.,
Sliwoski, G., Kothiwale, S., Meiler, J., & Lowe, E. W. (2014). Computational
Anyim, G., & Faloye, K. O. (2022). Topical anti-inflammatory activity of
methods in drug discovery. Pharmacological Reviews, 66(1), 334–395.
Petiveria alliacea, chemical profiling and computational investigation of
https://doi.org/10.1124/pr.112.007336
phytoconstituents identified from its active fraction. Chemistry Africa, 5(3),
Sharma, M., & Chauhan, P. M. (2012). Dihydrofolate reductase as a thera-
557–565. https://doi.org/10.1007/s42250-022-00339-y
peutic target for infectious diseases: Opportunities and challenges.
Olanlokun, J. O., Olotu, A. F., David, O. M., Idowu, T. O., Soliman, E. M., &
Future Medicinal Chemistry, 4(10), 1335–1365. https://doi.org/10.4155/
Olorunsogo, O. O. (2019). A novel compound purified from Alstonia
fmc.12.68
boonei inhibits Plasmodium falciparum lactate dehydrogenase and Sharma, P., Joshi, T., Joshi, T., Chandra, S., & Tamta, S. (2020). In silico
plasmepsin II. Journal of Biomolecular Structure & Dynamics, 37(8), screening of potential antidiabetic phytochemicals from Phyllanthus
2193–2200. https://doi.org/10.1080/07391102.2018.1483840 emblica against therapeutic targets of type 2 diabetes. Journal of
Ogboye, R. M., Patil, R. B., Famuyiwa, S. O., & Faloye, K. O. (2022). Novel Ethnopharmacology, 248, 112268. https://doi.org/10.1016/j.jep.2019.
a-amylase and a-glucosidase inhibitors from selected Nigerian antidia- 112268
betic plants: An in silico approach. Journal of Biomolecular Structure & Shamshad, H., Bakri, R., & Mirza, A. Z. (2022). Dihydrofolate reductase,
Dynamics, 40(14), 6340–6349. https://doi.org/10.1080/07391102.2021. thymidylate synthase, and serine hydroxy methyltransferase:
1883480 Successful targets against some infectious diseases. Molecular Biology
Otunola, G. A., Oloyede, O. B., Oladiji, A. T., & Afolayan, A. J. (2010). Reports, 49(7), 6659–6691. https://doi.org/10.1007/s11033-022-07266-8
Comparative analysis of the chemical composition of three Subramanian, N., Sundaraganesan, N., & Jayabharathi, J. (2010).
spices–Allium sativum L. Zingiber officinale Rosc. and Capsicum fru- Molecular structure, spectroscopic (FT-IR, FT-Raman, NMR, UV) studies
tescens L. commonly consumed in Nigeria. African Journal of and first-order molecular hyperpolarizabilities of 1, 2-bis (3-methoxy-
Biotechnology, 9(41), 6927–6931. 4-hydroxybenzylidene) hydrazine by density functional method.
Ou-Yang, S. S., Lu, J. Y., Kong, X. Q., Liang, Z. J., Luo, C., & Jiang, H. Spectrochimica Acta. Part A, Molecular and Biomolecular Spectroscopy,
(2012). Computational drug discovery. Acta Pharmacologica Sinica, 76(2), 259–269. https://doi.org/10.1016/j.saa.2010.03.033
33(9), 1131–1140. https://doi.org/10.1038/aps.2012.109 Tohma, H., G€ ulçin, _I., Bursal, E., Go
€ren, A. C., Alwasel, S. H., & Ko
€ksal, E.
Owoseeni, O. D., Patil, R. B., Phage, P. M., Ogboye, R. M., Ayoola, M. D., (2017). Antioxidant activity and phenolic compounds of ginger
Famuyiwa, S. O., Gboyero, F. O., Ndinteh, D. T., & Faloye, K. O. (2022). (Zingiber officinale Rosc.) determined by HPLC-MS/MS. Journal of Food
Computational assessment of xanthones from African medicinal Measurement and Characterization, 11(2), 556–566. https://doi.org/10.
plants as aldose reductase inhibitors. Computation, 10(9), 146. https:// 1007/s11694-016-9423-z
doi.org/10.3390/computation10090146 Tona, L., Mesia, K., Ngimbi, N. P., Chrimwami, B., Cimanga, K., Bruyne,
Peters, W. (1965). Drug resistance in Plasmodium berghei. I. Chloroquine T. D., Apers, S., Hermans, N., Totte, J., Pieters, L., Vlietinck., &
resistance. Experimental Parasitology, 17(1), 80–89. https://doi.org/10. A. J. Okond’ahoka. (2001). In-vivo antimalarial activity of Cassia occi-
1016/0014-4894(65)90012-3 dentalism Morinda morindoides and Phyllanthus niruri. Annals of
Peng, F., Tao, Q., Wu, X., Dou, H., Spencer, S., Mang, C., Xu, L., Sun, L., Tropical Medicine & Parasitology, 95(1), 47–57., https://doi.org/10.1080/
Zhao, Y., Li, H., Zeng, S., Liu, G., & Hao, X. (2012). Cytotoxic, cytopro- 00034983.2001.11813614
tective and antioxidant effects of isolated phenolic compounds from Treben, M. (1986). Health through God’s pharmacy – advice and experiences
with medicinal plants, 6. Australia: Wilhelm Ennsthater, Steys, Publ.
fresh ginger. Fitoterapia, 83(3), 568–585. https://doi.org/10.1016/j.
Trott, O., & Olson, A. J. (2010). AutoDock Vina: improving the speed and
fitote.2011.12.028
accuracy of docking with a new scoring function, efficient optimization,
Prasad, S., & Tyagi, A. K. (2015). Ginger and its constituents: Role in preven-
and multithreading. Journal of Computational Chemistry, 31(2), 455–461.
tion and treatment of gastrointestinal cancer. Gastroenterology Research
Vandebroek, I., & Picking, D. (2020). Zingiber officinale Roscoe
and Practice, 2015, 1–11. https://doi.org/10.1155/2015/142979
(Zingiberaceae). In Popular medicinal plants in Portland and Kingston,
Plowe, C. V., Djimde, A., Bouare, M., Doumbo, O., & Wellems, T. E. (1995).
Jamaica, pp. 235–245. Springer.
Pyrimethamine and proguanil resistance-conferring mutations in
Verpoorte, R., Frederich, M., Delaude, C., Angenot, L., Dive, G., Thépenier,
Plasmodium falciparum dihydrofolate reductase: Polymerase chain P., Jacquier, M. J., Zèches-Hanrot, M., Lavaud, C., & Nuzillard, J. M.
reaction methods for surveillance in Africa. The American Journal of (2010). Moandaensine, a dimeric indole alkaloid from Strychnos
Tropical Medicine and Hygiene, 52(6), 565–568. https://doi.org/10.4269/ moandaensis (Loganiaceae). Phytochemistry Letters, 3(2), 100–103.
ajtmh.1995.52.565 https://doi.org/10.1016/j.phytol.2010.02.005
Raheem, A. H., Al-Shejyri, K. J., & Al-Bermany, E. D. (2012). Density func- World Health Organization. (2021). World Malaria Report 2013.
tional theory calculations for methylbenzene molecules group. British World Health Organization. (2022, July 26). Malaria. https://www.who.int/
Journal of Science, 5, 57–64. news-room/fact-sheets/detail/malaria
Rampogu, S., Baek, A., Gajula, R. G., Zeb, A., Bavi, R. S., Kumar, R., Kim, Y., Yadav, M. K., Tripathi, M. K., & Yadav, S. (2021). Discovery of novel inhibitors
Kwon, Y. J., & Lee, K. W. (2018). Ginger (Zingiber officinale) phyto- targeting Plasmodium knowlesi dihydrofolate reductase using molecular
chemicals—gingerenone-A and shogaol inhibit SaHPPK: Molecular docking and molecular dynamics simulation. Microbial Pathogenesis,
docking, molecular dynamics simulations and in vitro approaches. 161(Pt A), 105214. https://doi.org/10.1016/j.micpath.2021.105214
Annals of Clinical Microbiology and Antimicrobials, 17(1), 1–15. https:// Yusuf, A. A., Lawal, B., Abubakar, A. N., Berinyuy, E. B., Omonije, Y. O.,
doi.org/10.1186/s12941-018-0266-9 Umar, S. I., Shebe, M. N., & Alhaji, Y. M. (2018). In-vitro antioxidants,
16 K. O. FALOYE ET AL.

antimicrobial and toxicological evaluation of Nigerian Zingiber offici- resistance and development of effective inhibitors. Parasitology, 130(Pt 3),
nale. Clinical Phytoscience, 4(1), 1–8. https://doi.org/10.1186/s40816- 249–259. https://doi.org/10.1017/s003118200400664x
018-0070-2 Zareen, S., Khan, S. N., Adnan, M., Haleem, S., Ali, R., & Alnomasy, S. F.
Yu, T. J., Tang, J. Y., Shiau, J. P., Hou, M. F., Yen, C. H., Ou-Yang, F., Chen, (2021). Antiplasmodial potential of Eucalyptus obliqua leaf methanolic
C. Y., & Chang, H. W. (2022). Gingerenone A Induces Antiproliferation extract against Plasmodium vivax: An in vitro study. Open Chemistry,
and Senescence of Breast Cancer Cells. Antioxidants, 11(3), 587. 19(1), 1023–1028. https://doi.org/10.1515/chem-2021-0091
https://doi.org/10.3390/antiox11030587 Zhong, W., Zhu, J., Yi, J., Zhao, C., Shi, Y., Kang, Q., Huang, J., Hao, L., &
Yuthavong, Y., Yuvaniyama, J., Chitnumsub, P., Vanichtanankul, J., Lu, J. (2022). Biochemical analysis reveals the systematic response of
Chusacultanachai, S., Tarnchompoo, B., Vilaivan, T., & motion sickness mice to ginger (Zingiber officinale) extract’s amelior-
Kamchonwongpaisan, S. (2005). Malarial (Plasmodium falciparum) dihydro- ation effect. Journal of Ethnopharmacology, 290, 115077. https://doi.
folate reductase-thymidylate synthase: Structural basis for antifolate org/10.1016/j.jep.2022.115077

View publication stats

You might also like