Next Article in Journal
Molecular Mechanism of Sirtuin 1 Inhibition by Human Immunodeficiency Virus 1 Tat Protein
Next Article in Special Issue
Antimicrobial Peptides: The Production of Novel Peptide-Based Therapeutics in Plant Systems
Previous Article in Journal
Oncological Outcomes of Segmentectomy versus Lobectomy in Clinical Stage I Non-Small Cell Lung Cancer up to Two Centimeters: Systematic Review and Meta-Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phytochemicals as Invaluable Sources of Potent Antimicrobial Agents to Combat Antibiotic Resistance

by
Ragi Jadimurthy
1,
Swamy Jagadish
1,
Siddaiah Chandra Nayak
2,
Sumana Kumar
3,
Chakrabhavi Dhananjaya Mohan
1,* and
Kanchugarakoppal S. Rangappa
4,*
1
Department of Studies in Molecular Biology, University of Mysore, Manasagangotri, Mysore 570006, India
2
Department of Studies in Biotechnology, University of Mysore, Manasagangotri, Mysore 570006, India
3
Department of Microbiology, Faculty of Life Sciences, JSS Academy of Higher Education and Research, Mysore 570015, India
4
Institution of Excellence, Vijnana Bhavan, University of Mysore, Manasagangotri, Mysore 570006, India
*
Authors to whom correspondence should be addressed.
Life 2023, 13(4), 948; https://doi.org/10.3390/life13040948
Submission received: 27 January 2023 / Revised: 4 March 2023 / Accepted: 29 March 2023 / Published: 4 April 2023

Abstract

:

Simple Summary

Many microorganisms develop resistance to drugs through different mechanisms, and this process is called antimicrobial resistance. It is highly essential to discover new antimicrobials to kill pathogenic microbes that have developed antimicrobial resistance. Natural sources, including plants, have been serving as a great source of medicinally important compounds for the past several decades. In this article, we have discussed the antimicrobial properties of plant-derived compounds against drug-resistant human pathogens, including bacteria, fungi, and viruses.

Abstract

Plants have been used for therapeutic purposes against various human ailments for several centuries. Plant-derived natural compounds have been implemented in clinics against microbial diseases. Unfortunately, the emergence of antimicrobial resistance has significantly reduced the efficacy of existing standard antimicrobials. The World Health Organization (WHO) has declared antimicrobial resistance as one of the top 10 global public health threats facing humanity. Therefore, it is the need of the hour to discover new antimicrobial agents against drug-resistant pathogens. In the present article, we have discussed the importance of plant metabolites in the context of their medicinal applications and elaborated on their mechanism of antimicrobial action against human pathogens. The WHO has categorized some drug-resistant bacteria and fungi as critical and high priority based on the need to develope new drugs, and we have considered the plant metabolites that target these bacteria and fungi. We have also emphasized the role of phytochemicals that target deadly viruses such as COVID-19, Ebola, and dengue. Additionally, we have also elaborated on the synergetic effect of plant-derived compounds with standard antimicrobials against clinically important microbes. Overall, this article provides an overview of the importance of considering phytogenous compounds in the development of antimicrobial compounds as therapeutic agents against drug-resistant microbes.

1. Introduction

Antimicrobial agents are drugs that are used to prevent and treat infections caused by bacteria, fungi, viruses, and parasites. Thousands of small molecules and peptides were isolated from natural sources such as plants, bacteria, fungi, and marine invertebrates, and some have demonstrated significant antimicrobial activity in preclinical settings and clinics [1,2]. Therefore, they have been used as standard antimicrobial drugs against microbial infections. The period between 1940 to 1965 is considered the golden era of antibiotics as many new antibiotics were introduced to modern medicine which revolutionized the treatment of bacterial infections [3]. Unfortunately, the phenomenon of antimicrobial resistance is becoming one of the primary health concerns across the globe, in which the pathogens do not respond to existing antimicrobial agents, which complicates the treatment regimen and thereby increases the mortality rate [4]. There is a swift spread of pan-drug-resistant bacteria at an alarming rate.
The World Health Organization (WHO) has declared that antimicrobial resistance is one of the top 10 global public health threats facing humanity. As per the antibiotic resistance threats report (2019) of the Centers for Disease Control and Prevention (CDC, the United States), the annual death rate due to antibiotic-resistant infection is over 35,000 people in the United States alone [5]. Misuse and overuse of antimicrobial agents are the prime reasons for the development of resistance by microbes, which pose a serious health concern to mankind [6,7]. Microorganisms develop resistance to antimicrobials in various ways and we have comprehensively discussed the underlying mechanisms that are involved in the development of antibiotic resistance in bacteria in our previous report [8]. Despite continuous efforts, a marked number of effective antimicrobials have not been discovered in the last three decades. Many pharmaceutical companies are involved in the development of drugs for the treatment of non-communicable metabolic disorders that are of significant economic interest. It is the need of the hour to focus on the discovery and development of novel antibiotics to treat deadly infections caused by antimicrobial-resistant organisms.
Mother Nature is serving as a treasure house of medicinally important compounds that can be used against various human ailments, including cancer, malaria, inflammatory diseases, and microbial infections [9,10,11,12]. We earlier demonstrated the pharmacological activities of various natural compounds in preclinical models of different diseases [13,14,15,16,17]. Extensive screening and research advancements in the previous century led to the discovery of thousands of bioactive secondary metabolites from medicinally important plants. Plants have been serving as a great source of bioactive compounds, which are being tested in preclinical disease models and clinical trials. Natural compounds, or their semi-synthetic derivatives obtained from plants, have contributed to the development of drugs against microbial diseases and various human ailments. For instance, artemisinin, a sesquiterpene lactone obtained from Artemisia annua, is used as a therapeutic agent for the treatment of malaria that is caused by Plasmodium falciparum. Various traditional medicine systems, folklore, codified systems of medicine, ethnopharmacology, ayurvedic classical texts, or zoopharmacognosy propose that some plants can be used against microbial infections. Some of the plant-derived metabolites have shown good antibacterial, antifungal, and antiviral activities in preclinical settings, and they can be considered potential candidates to be examined in clinical trials. In the present article, we have comprehensively discussed the mechanism of action of selected plant metabolites that have shown promising antimicrobial (antibacterial, antifungal, and antiviral) activity against clinically important human pathogens. Although some articles have been published in a similar line, many of them have not focused on the effect of plant-derived natural compounds on pathogenic microorganisms that are listed by the WHO as threats to mankind. We have attempted to provide a holistic view of the effect of selected natural compounds that have shown good growth-inhibitory activity toward clinically prominent microorganisms. We have also emphasized the synergistic effect of natural compounds with standard chemotherapeutic agents against human pathogens.

2. Phytochemicals as a Source of Antimicrobial Compounds

2.1. Antibacterial Agents Derived from Plants

Among microbial infections, bacterial infections pose a huge threat to human life across the globe. The WHO has categorized bacterial pathogens into critical, high, and medium priority depending on the need to develop new drugs against drug-resistant bacteria. The bacteria that are grouped under critical priority encompass carbapenem-resistant-Acinetobacter baumannii and -Pseudomonas aeruginosa, carbapenem-resistant, and third-generation cephalosporin-resistant Enterobacteriaceae. The bacteria that are categorized as high priority include vancomycin-resistant Enterococcus faecium, methicillin-resistant, vancomycin-resistant-Staphylococcus aureus, clarithromycin-resistant Helicobacter pylori, fluoroquinolone-resistant Campylobacter spp., fluoroquinolone-resistant Salmonella spp., and third-generation cephalosporin-resistant, fluoroquinolone-resistant-Neisseria gonorrhoeae. The medium priority list comprises penicillin-non-susceptible Streptococcus pneumoniae, ampicillin-resistant Haemophilus influenzae, and fluoroquinolone-resistant Shigella spp. [18]. Additionally, some of the bacteria included in these lists are also grouped as ESKAPE pathogens, which include E. faecium, S. aureus, Klebsiella pneumoniae, A. baumannii, P. aeruginosa, and Enterobacter spp., as they have adapted the escape mechanisms from the action of antibiotics. In the following section, we have discussed the mechanism of action of plant-derived natural compounds that target the abovementioned clinically important bacteria (Table 1). The structure of plant-derived compounds that are active against bacteria discussed in the text is given in Figure 1.

2.1.1. Apigenin

Apigenin is a flavonoid found in various plants, including Petroselinum crispum, Matricaria chamomilla, Apium graveolens, Basella rubra, Cynara scolymus, Origanum vulgare, and Portulaca oleracea [19]. It was found to have antibacterial activity against P. aeruginosa, K. pneumoniae, Salmonella typhimurium, Enterobacter aerogenes, and Proteus mirabilis. Apigenin was found to inhibit H. pylori-derived D-Alanine:D-alanine ligase with a relatively lesser IC50 value (132.7 μM) than a positive control D-cycloserine (299 μM). Apigenin displayed a binding affinity towards H. pylori-derived D-Alanine:D-alanine ligase (kD value: 22.3 μM), as demonstrated by surface plasmon resonance studies. In functional studies, apigenin displayed moderate antibacterial activity (MIC: 25 μg/mL) against H. pylori [20]. In a mass spectrometry-based assay to measure efflux pump inhibition, apigenin exerted efflux pump inhibition in S. aureus with an IC50 value of 38 µg/mL [21].

2.1.2. 18-β-Glycyrrhetinic Acid

Glycyrrhizic acid is the primary saponin found in Glycyrrhiza glabra L., of the licorice family [22]. Glycyrrhizic acid and its derivatives are endowed with antibacterial, antitumor, antiviral, antibacterial, and anti-inflammatory activities [23]. Notably, glycyrrhizic acid is metabolically inactive and thus, it is metabolized to 18-β-glycyrrhetinic acid by the intestinal microflora upon consumption, leading to its absorption into the bloodstream. It was demonstrated that 18β-glycyrrhetinic acid-induced bactericidal activity against methicillin-resistant Staphylococcus aureus (MRSA) and its topical application significantly reduced staphylococcal skin and soft tissue infection in mice models. At the transcript level, pathogenicity-associated transcripts such as saeR (regulatory gene component of the global virulence regulatory system), hla (codes for α-toxin), sbi (gene essential for evasion of antibodies and complement), and mecA (gene associated with offering resistance to β-lactam antibiotics) were downregulated in 18-β-glycyrrhetinic acid-treated MRSA. A similar alteration in transcript abundance was obtained in bacterial mRNA isolated from the infected tissues of mice infected subcutaneously with MRSA [23].

2.1.3. Honokiol

Honokiol [3′,5-di-(2-propenyl)-1,1′-biphenyl-2,2′-diol] is a bioactive neolignan found in the root bark of many species of the Magnoliaceae family, such as Magnolia officinalis, Magnolia obovata, and Magnolia grandiflora [24]. Honokiol has shown antibacterial potency against a wide range of bacteria from common oral pathogens to some of the ESKAPE organisms. Colistin is a last-line antibiotic that can be implemented in the treatment of multidrug-resistant (MDR) bacterial infections. Unfortunately, the emergence of mcr-1 (a plasmid-mediated colistin resistance gene) is threatening the clinical use of colistin [25]. Guo and colleagues demonstrated that honokiol enhances the sensitivity of MCR-1-positive Enterobacteriaceae infections to polymyxin (polypeptide antibiotics) in vitro and in vivo. Molecular dynamics simulations showed that honokiol establishes hydrogen bonding and hydrophobic interactions with the active region of MCR-1 [26]. In another study, honokiol amphiphiles showed potent antibacterial activity against clinical isolates of MRSA (MIC: 0.5–2 µg/mL) with minimum cytotoxicity towards normal hepatocytes. These honokiol amphiphiles were found to disrupt biofilms and bacterial cell membranes, imparting bactericidal activity [27]. The major exopolysaccharide of the biofilm matrix of S. aureus is a chain of poly-N-acetyl-β-(1–6)-glucosamine, which is termed polysaccharide intercellular adhesin (PIA). Notably, the genes that code for enzymes essential for PIA synthesis are part of ica operon. Honokiol disintegrates existing biofilms of S. aureus by reducing the expression of biofilm-related genes (such as sarA, cidA, and icaA), decreasing the extracellular DNA release, and suppressing the expression of PIA [28]. In another study, honokiol was found to inhibit the secretion of α-hemolysin (an exotoxin released by S. aureus which selectively induces hemolysis of RBCs) and prevent α-hemolysin-induced hemolysis of rabbit RBCs. Honokiol protected mice from S. aureus-induced liver damage by suppressing the activation of the NLRP3 inflammasome and the expression of proinflammatory cytokines [29].

2.1.4. Kaempferol

Kaempferol [3,5,7-trihydroxy-2-(4-hydroxyphenyl)-4H-1-benzopyran-4-one] is a flavonol abundantly present in tea (Camellia sinensis), broccoli (Brassica oleracea), apple (Malus domestica), and strawberry (Fragaria x ananassa) [30]. It has also been reported to be found in medicinal plants, including Sophora japonica, Equisetum spp., Ginkgo biloba, and Euphorbia pekinensis [19]. The antibacterial activity and mechanism of action of kaempferol have been demonstrated in various studies. Kaempferol inhibits the PriA helicase (an enzyme crucial for the restart of DNA replication and bacterial survival) activity of S. aureus [31] and displayed efflux pump inhibition in S. aureus with an IC50 value of 19 µg/mL [21]. In another study, kaempferol exhibited a synergistic effect with colistin against biofilm formation and growth of clinical isolates of colistin-resistant Gram-negative bacteria such as P. aeruginosa, Escherichia coli, K. pneumoniae, and A. baumannii [32]. The observed antibacterial activity was found to be mediated by disrupting the integrity of the cell membrane by kaempferol, which enables the increased interaction of colistin with the lipopolysaccharide of target bacteria.
The pretreatment of kaempferol-3-O-glucorhamnoside (a derivative of kaempferol isolated from Thesium chinense Turcz) in mice challenged with K. pneumoniae, effectively downregulated the expression of important inflammatory mediators such as TNF-α, IL-6, IL-1β and PGE2 with parallel amelioration of lung edema. In addition, kaempferol-3-O-glucorhamnoside rescued RAW cells from the deleterious effects of K. pneumoniae infection [33]. Biofilm formation is one of the important factors responsible for offering resistance against antibacterial agents; therefore, the development of antibiofilm agents can increase the drug-sensitivity of bacteria. Attachment, maturation, and detachment are the three crucial phases in the development of the bacterial biofilm. Ming and colleagues identified that kaempferol suppresses the primary attachment phase of biofilm formation in S. aureus by decreasing the activity of sortase A and the expression of adhesion-related genes. S. aureus surface proteins (such as ClfA and ClfB) attach to the cell wall by sortase A and have a prominent role in the formation of biofilms [34].

2.1.5. Naringin and Naringenin

Naringin (4′,5,7-trihydroxyflavanone-7-rhamnoglucoside) is a flavonoid glycoside that is excessively present in grapefruit and orange, whereas naringenin (5,7,4′-trihydroxyflavanone) is an aglycone form of naringin. Naringenin is effective against Enterococcus faecalis, a gram-positive bacterium present in the alimentary canal of humans and animals, which can cause life-threatening diseases in humans. Homology modeling and docking studies demonstrated that naringenin interacts with the active site of β-ketoacyl acyl carrier protein synthase Ⅲ, which is a key enzyme in the initiation of fatty acid synthesis in bacteria. The same study also demonstrated that naringenin displayed an antibacterial effect against E. faecalis (MIC: 256 µg/mL) [35]. The combination of naringenin with oxacillin and cloxacillin displayed synergistic antibacterial activity against MRSA [36]. The therapeutic application of naringenin is hampered due to its poor aqueous solubility and low bioavailability upon oral administration. Khan and coworkers prepared a naringenin-loaded, self-nanoemulsifying drug delivery system (NRG-SNEDDS) and examined the bioavailability upon oral administration. The total plasma concentration of NRG-SNEDDS was found to be significantly elevated compared to the naringenin control [37].
Several studies have demonstrated the antibacterial and antibiofilm effect of naringin in combination with standard antibiotics that are used in clinics. For instance, naringin potentiated the antibiofilm activity of ciprofloxacin and tetracycline on P. aeruginosa upon combinational treatment [38]. Naringin was reported to abrogate the biofilms of metallo-β-lactamases producing Pseudomonas species, which was evidenced by a remarkable reduction in the production of exopolysaccharides and alginate [39]. Zhou and colleagues reviewed the underlying antibacterial mechanisms of naringin and reported recently [40].

2.1.6. Nimbolide

Nimbolide (5,7,4′-trihydroxy-3′, 5′-diprenyl flavanone) is one of the vital limonoids present in the seeds, leaves, and flowers of Azadirachta indica, commonly known as neem [41]. It displayed bactericidal activity against H. pylori, a pathogen responsible for some diseases of the gastrointestinal tract, including peptic ulcers and gastric cancer. Merrell and colleagues demonstrated that neem oil extract possesses bactericidal activity [42]. Since A. indica has been reported to possess more than 300 phytochemicals, the same research group also demonstrated that nimbolide imparts bactericidal and antibiofilm activity against H. pylori [43]. In another study, nimbolide induced significant growth inhibitory activity against multi-drug-resistant (MDR) MRSA by damaging the membrane, lysis of bacterial cells, and disruption of biofilm [44].

2.1.7. Resveratrol

Resveratrol (3,4′,5,-trihydroxystilbene) is a naturally occurring phytoalexin and is present in red wine, grapes, peanuts, berries, etc. [45,46]. It has exhibited antibacterial properties against a variety of organisms including E. coli, vancomycin-intermediate Staphylococcus aureus (VISA), S. aureus, Campylobacter species, and Vibrio species [47]. In an interesting study, the effect of resveratrol in combination with polymyxin B was examined against 50 MDR bacterial strains (26 strains of K. pneumoniae and 24 strains of E. coli), and among them, 44 strains were resistant to polymyxin B. Interestingly, resveratrol potentiated the antibacterial activity of polymyxin B against K. pneumoniae and E. coli [48]. Resveratrol has been reported to inhibit the electron transport chain and F0F1-ATPase, which contributes to the decline of ATP production and subsequent suppression of the growth of microorganisms [47]. It can also impart antibacterial activity by forming a copper-peroxide complex, upon which it interacts with DNA to form a DNA-resveratrol-copper ternary complex, which in turn ultimately results in the induction of DNA damage [49]. Another study indicated that resveratrol imparted growth-inhibitory activity against E. coli via inhibition of Z-ring formation through abrogation of FtsZ expression. FtsZ serves a pivotal role by assembling into a contractile ring (called the Z-ring) at the midcell site of the future septum during the division of bacteria [50]. The other mechanisms involved in the induction of antimicrobial effects by resveratrol against clinically important bacterial pathogens are comprehensively reviewed in the previous reports [47,51].

2.1.8. Sanguinarine

Sanguinarine is a benzophenanthridine alkaloid obtained from the rhizomes of Sanguinaria canadensis L. (bloodroot), Chelidonium majus L. (Celandine), Fumaria officinalis L. (Fumitory), and Bocconia frutescens L. (Plume poppy). In P. aeruginosa, glucose enters the cell through the OprB and OprB2 porins and enters the periplasmic space, where it can directly enter the cytoplasm through an ABC transporter or it can be metabolized, which subsequently transported from the periplasm to the cytoplasm through the KguT (2-ketogluconate transporter). It has been shown that the P. aeruginosa mutant that lacks 2-ketogluconate transporter was relatively less pathogenic than wild-type P. aeruginosa [52]. Falchi and colleagues demonstrated that sanguinarine suppresses the 2-ketogluconate pathway of glucose utilization in P. aeruginosa by either targeting KguD or KguK [53]. In another report, sanguinarine induced an antibacterial effect against MRSA by triggering the release of membrane-bound cell wall autolytic enzymes, which eventually leads to bacterial lysis [54]. Interestingly, sanguinarine potentiated the antibacterial efficacy of standard clinically used antibiotics (such as ampicillin, oxacillin, norfloxacin, ciprofloxacin, and vancomycin) against MRSA [55].

2.1.9. Withaferin A

Withaferin A (4-β,27-dihydroxy-1-oxo-5β,6β-epoxywitha-2,24-dienolide) is a natural steroidal lactone present in Withania somnifera and other members of the Solanaceae family, such as Acnistus arborescens [56]. Withaferin A displayed effective antibacterial activity against P. aeruginosa with a MIC and MBC of 60 µM and 80 µM, respectively. The effect was mediated by damaging the bacterial cell membrane. In addition, a significant reduction in the level of ROS and lipid peroxidation was reported upon withaferin A administration in P. aeruginosa-infected zebra fish larvae [57]. Metallo-β-lactamases are the antibiotic inactivating enzymes that contribute to the resistance against carbapenems. New Delhi metallo-β-lactamase (NDM-1) is contributing greatly to the emergence of antibiotic resistance among ESKAPE pathogens. In silico and in vitro screening studies revealed that withaferin A reduced the enzyme activity of New Delhi metallo-β-lactamase (IC50: 24.03 ± 2.9 μM). Withaferin A also displayed good synergistic activity with imipenem against clinical isolates of NDM-1-producing carbapenem-resistant A. baumannii with an FIC index value of 0.3125.
Table 1. List of phytochemicals that have demonstrated antibacterial activity against clinically important organisms.
Table 1. List of phytochemicals that have demonstrated antibacterial activity against clinically important organisms.
Sl. No.PhytocompoundSourcesMicroorganisms Affected by the Title Compound and Dose Mechanism of ActionRef.
1AllicinAllium sativum,
Allium spp.
Streptococcus pneumoniae (MIC: 64 µg/mL),
Streptococcus pyogenes (MIC: 32 µg/mL)
ND [58]
2ConessineHolarrhena floribunda, Holarrhena antidysenterica, Funtumia elasticaP. aeruginosa (MIC: 20 mg/L)Inhibition of MexAB-OprM efflux pump [59]
3ThymolThymus vulgaris, Thymus capitatusK. pneumoniae (MIC: 128 µg/mL)Inhibition of biofilm formation[60,61]
S. aureus (MIC: 72 µg/mL)Reversal of efflux pump action
4CarvacrolOriganum vulgareS. aureus (MIC: 256 µg/mL)Reversal of efflux pump action [61]
5EugenolSyzygium aromaticum, Eugenia caryophyllusA. baumannii,
Salmonella enteritidis, Campylobacter Jejuni,
P. aeruginosa, E. coli
ND[62]
6BerberineBerberis vulgaris,
Berberis fremontii, Hydrastis Canadensis
H. pyloriIncreased the sensitivity of amoxicillin and tetracycline, and reduced the expression of hefA mRNA upon treatment with amoxicillin, tetracycline, and berberine[63]
7Curcumin ICurcuma longaP. aeruginosaDamage to the bacterial membrane[64]
H. pyloriInhibition of biofilm formation
8QuercetinCapparis spinosa,
Polymnia fruticose, Ginkgo biloba
Salmonella enterica serotype Typhimurium (MIC: 0.0072 µm/mL),
S. aureus (MIC: 0.0068 µm/mL),
P. aeruginosa (MIC: 0.0085 µm/mL)
Disruption of cell membrane integrity, thereby causing cell lysis[65]
9EpigallocatechinCamellia sinensisS. aureus (MIC: 62.5 µg/mL),
P. aeruginosa (MIC: 125 µg/mL)
Increased the sensitivity of gentamycin against S. aureus and P. aeruginosa[66]
10CatechinFructus CrataegiMRSA (MIC: 0.1 g/L)Inhibition of biofilm formation via suppression of fibronectin-binding protein A and B (fnbA and fnbB)[67]
11GenisteinGlycine maxAeromonas hydrophilaDisruption of QS, biofilm formation, and aerolysin production[68]
12Protocatechuic acidScrophularia frutescensYersinia enterocolitica (MIC: 2.5 mg/mL)Cell membrane depolarization, reduction of intracellular pH and adenosine triphosphate (ATP), leakage of potassium ions [69]
13Gallic acidVitis rotundifoliaP. aeruginosa (MIC: 500 µg/mL),
S. aureus (MIC: 1750 μg/mL),
Listeria monocytogenes (MIC: 2000 μg/mL)
Membrane permeabilization, the release of intracellular potassium ions, disruption of the physicochemical surface properties of the cell[70,71]
Shigella flexneri (MIC: 2 mg/mL)Inhibition of biofilm formation via regulation of mdoH gene expression and the OpgH protein
14HydroquinoneVaccinium myrtillusP. aeruginosa (MIC: 7.81 µg/mL),
S. aureus (MIC: 15.625 µg/mL)
Depolarization of the cell membrane potential, increase in cell permeability, and leakage of intracellular potassium ions[72]
15OstholeCnidium monnieri, Angelica archangelica, Angelica pubescensS. typhimurium (MIC: 1.67 ± 0.58 µg/mL),
K. pneumoniae (MIC: 3.33 ± 1.15 µg/mL),
A. baumannii (MIC: 1.68 ± 0.58 µg/mL)
ND[73]
16TaxifolinSilybum marianum, Allium cepa, Pseudotsuga taxifolia, Pinus pinasterE. faecalis (MIC: 128 µg/mL),
VREF (MIC: 512 µg/mL)
Based on docking data, taxifolin showed a good binding affinity for β-ketoacyl acyl carrier protein synthase III, which is an important enzyme for bacterial fatty acid biosynthesis[74]
17PiperinePiper nigrumMRSA (MIC: 100 µg/mL)
Liposomal formulation of piperine and gentamicin acts as an efflux pump inhibitor[75]
S. aureus (MIC: >16 µg/mL)Piperine, in combination with ciprofloxacin, causes inhibition of efflux pump
18Sophoraflavanone
B
Desmodium caudatumMRSA (MIC: 15.6–31.25 µg/mL)Disturbance of the cell membrane and leakage of cell contents[76]
19FarnesolVachellia farnesianaS. aureus
(MIC: 184 µg/mL),
L. monocytogenes
(MIC: 133 µg/mL)
ND[77]
20RosthorninRabdosia rosthorniiPropionibacterium acnes (MIC: 3.17–25 µg/mL)ND[78]
21Ellagic acidRosa rugosaH. pylori (MIC: 5–30 mg/L)ND [79]
22Chebulagic acidTerminalia chebulaA. baumanniiND[80]
23.Hexahydroxy diphenoyl
ester vescalagin
Lythrum salicariaS. aureus (MIC: 62 µg/mL),
P. mirabilis (MIC: 62 µg/mL)
ND[81]
24StigmasterolNeocarya macrophyllaMRSA (MIC: 6.25–25 µg/mL),
Streptococcus faecalis
(MIC: 6.25–25 µg/mL), S. aureus
(MIC: 6.25–25 µg/mL)
Broad spectrum antibacterial activity[82]
25Chlorogenic acidFruits, vegetables, and graminaceous plantsStreptococcus pneumoniae (MIC: 20 µg/mL),
Salmonella typhimurium (MIC: 20 µg/mL),
Shigella dysenteriae (MIC: 10 µg/mL)
An increase in cell membrane permeability binds to bacterial DNA and thereby inhibits cellular functions [83]
26ThymoquinoneNigella sativaS. flexneri
(MIC: 0.4 mg/mL)
Disruption of the cell membrane integrity [84]
27GuggulsteroneCommiphora wightii (Arn.) BhandariE. coli (MIC: 0.5 mg/mL),
K. pneumoniae (MIC: 2 mg/mL),
P. aeruginosa (MIC: >2 mg/mL),
Salmonella typhi (MIC: >2 mg/mL),
E. faecalis (MIC: 0.5 mg/mL),
S. aureus (MIC: 2 mg/mL)
ND[85]
28IsoliquiritigeninGlycyrrhiza uralensisStaphylococcus xylosus (MIC: 80 µg/mL)Downregulation of the IGPD gene [86]
29CelastrolTripterygium
Wilfordii
S. aureus (MIC:1.25 µg/mL),
E. faecalis (MIC: 1.25 µg/mL)
Disruption of DNA and protein synthesis [87]
30CryptotanshinoneSalvia miltiorrhiza BungeS. aureus (MIC:12.5 µg/mL)Dissipation of membrane
Potential. Respiratory chain inhibition probably by
targeting type II NADH:quinone dehydrogenase
[88]
31OridoninRabdosia rubescens,
Isodon rubescens
MRSA (MIC: 64 µg/mL)Permeability of cell membrane, disruption in protein and DNA metabolism[89]
32MagnololMagnolia officinalisS. aureus (MIC: 16 ppm)Based on simulation studies magnolol exhibited a high binding affinity for cell division Protein FtsZ [90,91]
MRSA (MIC: 10 µg/mL)Repression of mecA, mecI, and upregulation of mecR1
33Hesperidincitrus fruits,
Poncirus trifoliata
S. aureus (MIC: 1 mg/mL),
Bacillus cereus (MIC: 2 mg/mL),
E. coli (MIC: >2 mg/mL)
P. aeruginosa (MIC: 2 mg/mL)
ND[92]
35EvocarpineEvodiae fructusMycobacterium smegmatis
(MIC: 2–4 mg/mL),
Mycobacterium
tuberculosis
(MIC: 5 mg/mL)
ND[93]
36Ursolic acidMalus domesticaK. pneumoniae
(MIC: 400 µg/mL),
CRKP-1 (MIC: 800 µg/mL),
CRKP-2 (MIC: 800 µg/mL),
CRKP-8 (MIC: 800 µg/mL),
CRKP-10 (MIC: 800 µg/mL)
Increase in membrane integrity, reduction in membrane potential, and intracellular ATP[94]
37Ferulic aidAll plants E. coli (MIC: 100 µg/mL),
P. aeruginosa (MIC: 100 µg/mL),
S. aureus (MIC: 1100 µg/mL),
L. monocytogenes
(MIC: 1200 µg/mL)
Disruption of membrane integrity, cell surface hydrophobicity, and potassium ion leakage[71]
38MorusinMorus albaS. aureus (MIC:14.9 μmol/L)Disruption of membrane integrity,
Modulation of expression of phosphatidic acid biosynthesis-associated genes
[95]
39LonicerinLonicera japonicaP. aeruginosaInhibition of
alginate secretion protein (AlgE) and inhibition of biofilm formation
[96]
40GalanginAllium sativumVISA
(MIC: 32 μg/mL)
Inhibition of murein hydrolase activity and growth of VISA strain-Mu50[97]
41ArtemisininArtemisia annuaS. aureus
(MIC: 0.09 mg/mL)
ND[98]
42PunicalaginPunica granatumS. aureus
(MIC: 0.25 mg/mL)
Disruption of the cell membrane, leakage of potassium ions, Inhibition of biofilm formation[99]
43Aloe-emodinCassia occidentalis, Aloe vera, Polygonum multiflorum Thunb.S. aureus (MIC: 32 μg/mL),
MRSA
(MIC: 16 μg/mL),
Staphylococcus epidermidis
(MIC: 4 μg/mL),
P. aeruginosa
(MIC: 256 μg/mL)
Transcriptional profile studies have revealed alterations of genes involved in sulfur metabolism, L-lysine, peptidoglycan biosynthesis, and biofilm formation [100]
44Skullcapflavone IIScutellaria
baicalensis
M. smegmatis
(MIC99: 128 mg/L),
Mycobacterium aurum (MIC99: 7.8 mg/L),
Mycobacterium bovis
(MIC99: 31.25 mg/L)
Efflux pump inhibition in
M. aurum and M. smegmatis
[101]
45WogoninAgrimonia pilosaP. aeruginosaReduction of the quorum sensing-related genes. decreased production of virulence factors,
inhibition of biofilm formation
[102]
46SulforaphaneBrassica oleracea and other cruciferous plantsH. pylori
(MBC: 2.8–5.6 µg/mL)
Inhibition of bacterial urease[103]
47Arjunolic acidSyzygium guineense,
Syzygium cordatum
Shigella sonnei
(MIC: 30 µg/spot)
ND[104,105]
48Terminolic acidSyzygium guineenseS. sonnei
(MIC: 50 µg/spot)
ND[106]
49Asiatic acidCentella asiaticaClostridium difficile
(MIC: 10–20 μg/mL)
Disruption of membrane permeability, inhibition of cell motility[107]
50Cinnamic acidCinnamomum cassiaM. tuberculosis (MIC: 270 µM)
Neisseria gonorrhoeae (MIC: 6.75 mM)
ND[108,109]
51Caffeic acidAbundantly present in fruits and vegetables, such as olives, cinnamon, nutmeg, blueberries, apple, star aniseS. aureus
(MIC: 256 µg/mL)
ND[110]
52AndrographolideAndrographis paniculataBurkholderia pseudomallei (MIC: 0.5 µg/mL)Andrographolide-stabilized silver nanoparticle binding and charge neutralization at the membrane surface, and the production of Ag+ and ROS[111,112]
P. aeruginosaSuppression of QS regulators LasR and RhlR, which control
the expression of many genes in P. aeruginosa
53DiosgeninRhizoma polgonati, Smilax china, Trigonella foenumgraecumPorphyromonas gingivalis,
Prevotella intermedia
Inhibition of biofilm formation[113]
54RheinRheum palmatum, Reynoutria
japonica (Houtt.), Fallopia multiflora
Cutibacterium acnes (MIC: 6.25 µg/mL)Inhibition of C. acnes NADH dehydrogenase-2 activity[114,115]
MRSA (MIC: 62.5–250 μg/mL) Rhein in combination with oxacillin causes a reduction of mecA/mecI/mecR1 and blaZ/blaI/blaR1 gene expressions
55Riccardin C
derivatives
Reboulia hemisphaericaMRSA (MIC: 1 µg/mL),
E. faecalis (MIC: 4 µg/mL),
P. aeruginosa (MIC: >128 µg /mL),
Vibrio parahaemolyticus (MIC: >128 µg/mL)
Disruption of membrane permeability and cell morphology,
Alterations in intracellular Na+ and K+ concentrations,
Mutation in FabI (an enoyl-acyl carrier protein reductase) in the S. aureus mutants
[116]
56ArtesunateArtemisia annuaM. tuberculosis
(MIC: 75 µg/mL)
ND[117]
57Betulinic acidMikania cordataP. aeruginosa (MIC: 256 µg/mL),
S. aureus (MIC: 256 µg /mL)
Increased production of a superoxide anion radical and malondialdehyde, elevated NAD+/NADH ratio, reduced glutathione, and DNA fragmentation[118]
58SakuranetinPolymnia fruticosaH. pylori (MIC: 87.3 µM /mL)Inhibition of β-hydroxy acyl-acyl carrier protein dehydratase [119]
59ProtoanemoninRanunculus bulbosusS. aureus (MIC: 31.25 µg/mL),
P. aeruginosa (MIC: 62.5 µg/mL),
Serratia marcescens (MIC: 15.625 µg/mL),
K. pneumoniae (MIC: 31.25 µg /mL),
Providencia stuartii (MIC: 15.625 µg/mL),
P. acnes (MIC: 31.25 µg/mL),
Clostridium perfringens (MIC: 62.5 µg/mL)
Broad spectrum antibacterial activity[120]
60CapsaicinPiper nigrum,
Capsicum annuum
Streptococcus pyogenes
(MIC: 64–128 μg/mL)
Cell membrane damage, reduction of cell invasion and hemolytic activity, inhibition of biofilm formation[121,122]
61ThymoquinoneNigella sativaP. aeruginosa
(MIC: 1.56 µg/mL),
S. aureus
(MIC: 3.125 µg/mL)
Depolarization of the membrane, production of ROS, and inhibition of biofilm formation[84,123,124]
V. paraheamolyticus (MIC: 32µg/mL),
Vibrio alginolyticus (MIC: 256µg/mL),
Salmonella enterica serovar Typhimurium (MIC: >512 µg/mL),
Staphylococcus epidermidis (MIC: 8 µg /mL),
S. aureus
(MIC: 8 µg/mL)
Inhibition of biofilm formation
S. flexneri
(MIC: 0.4 mg/mL)
Disruption of cell membrane integrity, inhibition of biofilm formation
62PiceatannolGrapes, white tea, passion fruit, Japanese knotweedStreptococcus mutans,
Streptococcus sanguinis,
Streptococcus gordonii
Inhibition of Streptococcus glucosyl transferase-GtfC [125]
63CurcuminCurcuma longaMRSA
(MIC: 125–250 μg/mL),
E. faecalis,
P. aeruginosa
Membrane damage, inhibition of FtsZ proteins, inhibition of mecA gene transcription, reduced expression of PBP2α proteins[64,126]
64ReserpineRauvolfia serpentinaS. aureus (MIC:1200
µg/mL)
Inhibition of biofilm formation and virulence-regulatory proteins[127,128]
M. tuberculosisND
65TomatidineSolanum lycopersicumS. aureus, L. monocytogenes,
Bacillus species.
Inhibition of ATP synthase subunit C[129]
66IsoliquirtigeninDalbergia odorifera, Glycyrrhiza
uralensis
M. bovis
(MIC: 50 µg/mL),
MRSA (MIC: 50–100 µg/mL)
Inhibition of FAS I and FAS II[130]
672,2′,4-Trihydroxy chalconeDalbergia odoriferaM. bovis
(MIC: 55 µg/mL)
Inhibition of FAS I and FAS II[131]
68FisetinRhus cotinusM. bovis
(MIC: 63 µg/mL)
Inhibition of FAS II[131]
69ButeinRhus
verniciflua
M. bovis (MIC: 43 µg/mL)Inhibition of FAS II[131]
70CoumarinAll plants S. typhimurium (MIC: 2.5 mg/mL),
Enterobacter aerogenes (MIC: 0.625 mg/mL)
ND[132]
71PlumbaginPlumbago rosea, Plumbago zeylanicaS. aureus (MIC: 5
μg/mL),
MRSA (MIC: 4–8 μg/mL)
Inhibition of DNA gyrase[133]
72HibiscetinHibiscus sabdariffaK. pneumoniae (MIC: 1024 μg/mL),
E. aerogenes (MIC: 1024
μg/mL)
ND[134]
73TerchebulinTerminalia chebulaA. baumannii (MIC: 500 μg/mL)ND[135]
74NorwogoninScutellaria baicalensisA. baumannii (MIC: 128 µg/mL)ND[135]
Abbreviations: CRKP: carbepenem-resistant-Klebsiella pneumoniae; FAS: fatty acid synthase; IGPD: imidazole glycerol phosphate dehydratase; MBC: minimum bactericidal concentration; MIC: minimum inhibitory concentration; ND: not determined; QS: quorum sensing.

2.2. Antifungal Agents Derived from Plants

Fungal infections in humans can be considered one of the low-key maladies in the antimicrobial research and healthcare sectors. Medical interventions (such as the use of catheters, intravascular or intracranial devices, neurosurgical procedures, the usage of contaminated devices, and the overuse of broad-spectrum antibiotics), treatment-associated immunosuppression (organ transplantations or stem cell transplantations), disease-associated immunosuppression (HIV infection), and co-infections (tuberculosis) are the risk factors abetting the fungal infections in humans [136]. COVID-19-associated fungal infections, such as mucormycosis, aspergillosis, and candidaemia, are recent examples of co-infections. Fungal infections are annually causing around 1.6 million deaths, which is higher than the deaths caused by tuberculosis (1.5 million deaths/year) [137,138]. The number of deaths related to fungal infections is increasing every year, which is posing a serious global health concern. Until recently, the WHO did not have any action plan or guidelines for fungal infections. On 25th October 2022, the WHO released its first-ever fungal priority pathogens list (FPPL) to direct and drive the research efforts towards life-threatening fungal pathogens, to accelerate international coordination, and to attract investments in research and development in therapeutics and diagnostics against the fungal infections, in addition to many other goals throughout the world. The WHO categorized fungal pathogens into three priority groups, i.e., critical, high, and medium priority, based on their antibiotic resistance status and their public health impact [139,140]. In the FPPL report, Cryptococcus neoformans, Candida auris, Aspergillus fumigatus, and Candida albicans are categorized under critical priority; Nakaseomyces glabrata (Candida glabrata), Histoplasma spp., Eumycetoma causative agents, Mucorales. Fusarium spp., Candida tropicalis, and Candida parapsilosis are kept under high priority; and Scedosporium spp., Lomentospora prolificans, Coccidioides spp., Cryptococcus gattii, Pichia kudriavzeveii (Candida krusei), Talaromyces marneffei, Pneumocystis jirovecii, and Paracoccidioides spp. are listed under medium priority pathogens [139,140].
Cryptococcus neoformans stands as the top fungal pathogen as per multicriteria decision analysis (MCDA). It initially infects the lungs and later spreads to the central nervous system to cause lethal cryptococcal meningitis and cryptococcaemia with a mortality rate accounting for about 41% to 61% among the infected [141]. Unlike other fungal pathogens, this pathogen is not transferred from one person to another. Fluconazole, amphotericin B, and flucytosine are clinically approved drugs used for the treatment of C. neoformans infections. The mechanisms of antifungal resistance adapted by C. neoformans have not been precisely understood [139].
Candida auris is the next ranked life-threatening fungal pathogen that causes invasive candidiasis, which affects the heart, central nervous system, eyes, bones, and internal organs [142]. Echinocandins, azoles, pyrimidines, and polyenes are the only four classes of antifungal agents used in clinics today. Importantly, 90% of C. auris isolates confer resistance to at least one class of antifungal agents, and 30% of the isolates display resistance against at least two classes of antifungal drugs [138]. In addition, other fungal pathogens provided in the priority list pose serious health consequences by developing resistance against antimycotics. An increasing number of fungal infections, the availability of a limited number of antifungal agents, and emerging antibiotic resistance demand the discovery of new antifungal agents. In the following section, we have discussed the effects and mechanisms of action of some of the plant-derived secondary metabolites with promising antifungal activity. (Table 2). The structure of plant-derived compounds that are active against fungi is given in Figure 2.

2.2.1. Carvacrol

Carvacrol (5-isopropyl-2-methylphenol) is a major constituent of essential oils obtained from the Lamiaceae family of plants, such as thyme and oregano [143]. Investigations carried out by Ahmed et al., (2011) showed the fungicidal activity of carvacrol against the various strains of fluconazole-sensitive and -resistant candida species, such as C. albicans, C. tropicalis, C. parapsilosis, C. krusei, and C. glabrata (mean MIC values of 75–90 mg/L for fluconazole-sensitive candida species and 75–100 mg/L for fluconazole-sensitive candida species) [144]. The study also suggested that the fungicidal activity of carvacrol could be due to interference with ergosterol biosynthesis and disruption of membrane integrity. Rao et al., showed that carvacrol disrupts Ca2+ and H+ homeostasis in Saccharomyces cerevisiae. Transcriptional profiling post-exposure to carvacrol showed a robust transcriptional response closely resembling that of calcium stress. It was speculated that the antifungal activity of carvacrol could be due to the induction of Ca2+ stress and inhibition of the TOR signaling pathway [144]. Chaillot et al., (2015) demonstrated that carvacrol disrupts the integrity of the endoplasmic reticulum, which in turn leads to endoplasmic reticulum stress and unfolded protein response in C. albicans [145]. In another study, carvacrol imparted cell death in C. albicans by increasing ROS levels, disrupting the mitochondrial membrane potential, causing DNA fragmentation and metacaspase activation, increasing cytosolic and mitochondrial Ca2+ levels, and activating calcineurin [144].
Table 2. List of phytochemicals that have demonstrated antifungal activity against clinically important organisms.
Table 2. List of phytochemicals that have demonstrated antifungal activity against clinically important organisms.
Sl. No.PhytocompoundSourcesMicroorganisms Affected by the Title Compound and Dose Mechanism of ActionRef.
1CarvoneCarum carvi, Anethum graveolens, Mentha spicataC. albicansInhibition of the transition from yeast form to filamentous form[146]
2ThymolPresent in the plants belong to genera such as Thymus, Ocimum, Origanum, Satureja, Thymbra, MonardaCandida species (MIC: 100 µg/mL) Inhibition of H+ ATPase[147]
C. neoformansInterferes in intracellular Ca2+ homeostasis, reduction in ergosterol content through HOG-dependent pathway, reduction in protein glycosylation
3MentholMentha piperita, Mentha longifolia, etc.Aspergillus niger (MIC: 150 µg/mL),
Aspergillus fumigatus (MIC: 150 µg/mL),
Aspergillus flavus (MIC: 100 µg/mL), Aspergillus ochraceus (MIC: 100 µg/mL),
Alternaria alternate (MIC: 450 µg/mL), Botrytis cinerea (MIC: 400 µg/mL), Cladosporium spp. (MIC: 125 µg/mL),
Penicillium citrinum (MIC: 100 µg/mL), Penicillium chrysogenum (MIC: 300 µg/mL), Fusarium oxysporum (MIC: 200 µg/mL), and Rhizopus oryzae (MIC: 250 µg/mL)
Decreased the fungal growth dose-dependently [148]
4CinnamaldehydeCinnamomum
Cassia,
Cinnamomum burmannii
Geotrichum citri-aurantiiDisruption of cell wall permeability and integrity[149]
C. neoformans var. grubii (MIC90: 0.683 mg/mL)Damage to the cell wall, induction of cell gigantism
5CitronellalCymbopogon citratesC. albicans (MIC: 1 mg/mL)Disruption of membrane homeostasis,
inhibition of yeast to hyphal transition and biofilm formation
[150]
6WogoninScutellaria baicalensis Georgi T. rubrum (MIC50: 0.06 mM),
A. fumigatus
(MIC50: 0.23 mM)
Perturbance in cell wall synthesis, [151]
Trichophyton mentagrophytes (MIC50: 0.03 mM)Perturbance in cell wall synthesis and generation of reactive oxygen species
7Gallic acidPunica granatumT. rubrum
(MIC: 43.75 μg/mL)
Inhibition of ergosterol biosynthesis,
reduction in sterol 14α-demethylase P450 (CYP51) and squalene epoxidase activity
[152]
T.mentagrophytes, Trichophyton violaceum, Trichophyton verrucosum, Trichophyton schoenleinii
(Mean MIC: 54.17–83.33 μg/mL), C. albicans
(Mean MIC: 12.5 μg/mL)
ND
8α-pineneEucalyptus plantsC. parapsilosis (MFC: 128 μg/mL)Inhibition of pseudo-hyphae and promoting a marked reduction in blastoconidia[153]
9β-AsaroneAcorus calamusA. nigerReduces ergosterol content in the plasma membrane[154]
10QuercetinMorus alba, Camellia sinensis, Allium fistulosum, Calamus scipionum, Centella asiatica, Lactuca sativaC. albicansProgrammed cell death through mitochondrial dysfunction[155,156]
11OstholeCnidii Fructus,
Cnidium monnieri
Microsporum canis (MIC: 1.95 µg/mL)Decrease in 1,3-β-D-glucan and chitin contents[157]
12Plagiochin EMarchantia polymorphaC. albicansInduction of the metacaspase-dependent apoptotic pathway, inhibition of chitin biosynthesis[158,159]
13Riccardin D Dumortiera hirsuteC. albicansDown-regulation of hypha-specific genes, such as ALS1, ALS3, ECE1, EFG1, HWP1 and CDC35, leading to retardation of hypha formation[160,161]
Azole-resistant C. albicans strains (MIC80: 16 µg/mL)Interferes in sterol biosynthesis
14SilibininSilybum marianumC. albicansInhibition of biofilm development, disruption of cell membrane[162]
15Chlorogenic acidPresent in a wide variety of fruits, vegetables, olive oil, wine, and coffeeC. albicansInduction of apoptosis by mitochondrial depolarization, production of reactive oxygen species, DNA fragmentation, externalization of phosphatidyl serine. [163,164]
Abbreviations: HOG: high-osmolarity glycerol response; MFC: minimum fungicidal concentration; ND: not determined.

2.2.2. Emodin

Emodin (1,3,8-trihydroxy-6-methyl-anthraquinone) is a secondary metabolite produced by plants, such as Senna alata, Rumex abyssinicus, Odontites serotina, Reynoutria japonica, polygonum spp., and Rheum palmatum [165,166]. It possesses a therapeutic potential against many human ailments, such as hepatitis, cancer, arthritis, cholelithiasis, Alzheimer’s disease, ulcerative colitis, pancreatitis, asthma, and many bacterial and viral infections [167]. The antifungal activity of emodin against C. albicans (MIC: 12.5 μg/mL) was demonstrated [168]. Emodin showed antibiofilm activity and inhibition of hyphal formation in C. albicans cells [169]. Emodin also inhibited 50% total kinase activity of C. albicans at concentrations beginning from 0.5 µg/mL [169]. Additionally, emodin reduced the activity of CK2 (the most pleiotropic kinase in C. albicans cells) with an IC50 value of 2.7 μg/mL [169]. Molecular docking studies of emodin with CK2 showed that emodin binds to the ATP binding pocket of CK2 to impart its activity. In another study, emodin was found to reduce the activity of (1,3)-β-D-glucan synthase from C. albicans and increased cell wall damage [168]. (1,3)-β-D-glucan is the major polysaccharide found in the fungal cell wall and it is synthesized by (1,3)-β-D-glucan synthase. (1,3)-β-D-glucan synthase is regarded as the major drug target for the development of antifungal drugs, and echinocandins (known antifungal drugs) are known to target (1,3)-β-D-glucan synthase to impart antifungal activity.
Ma et al., (2020) isolated aloe-emodin (1,8-dihydroxy 3-(hydroxymethyl)-9,10-anthracenedione), a compound with a close structural resemblance with emodin, from the root and rhizome of Rheum palmatum and the leaves of Aloe vera [170]. Aloe-emodin was demonstrated to show antifungal activity through antimicrobial photodynamic therapy against C. albicans cells [170]. Antimicrobial photodynamic therapy is a novel, promising approach against drug-resistant microorganisms in which a photosensitizer compound is incubated with a microorganism and an appropriate wavelength of light is irradiated to it. Upon irradiation, the photosensitizer is excited and undergoes molecular collision with surrounding oxygen molecules, and generates ROS (such as superoxide anions, hydroxyl radicals, singlet oxygen, etc.). These ROS cause damage to cellular envelopes (such as the cell wall and cell membrane), ultimately leading to cell death. Ma and colleagues demonstrated that aloe-emodin can be used as a photosensitizer in antimicrobial photodynamic therapy against drug-resistant C. albicans cells [170].

2.2.3. Eucalyptol

Eucalyptol [1,8-Cineole (1,3,3-trimethyl-2-oxabicyclo [2.2.2]acetate)] is a major component of essential oils extracted from plants of Eucalyptus species, such as Eucalyptus smithii, Eucalyptus globules, etc. It is also obtained from the essential oils of other plants, such as tea trees, mugwort, rosemary, etc. [171]. Eucalyptol showed antifungal activity against C. albicans and C. glabrata (MIC90 value: 800 μg/mL) by increasing ROS generation, G1/S phase arrest, elevating membrane permeability, and disrupting mitochondrial membrane potential [172]. Gene expression analysis revealed that genes essential for hyphal cell wall protein (HWP1), secreted aspartyl proteinase (SAP1), and cell surface adhesion (ALS1) are downregulated [172]. Mishra and colleagues synthesized the eucalyptol/β-cyclodextrin inclusion complex-loaded gellan/PVA nanofibers (EPNF) and studied their antibiofilm activity against C. albicans and C. glabrata cells [173]. EPNF inhibited approximately 70% of biofilm formation in the aforementioned fungal cells. A time-kill assay showed that the antifungal activity of EPNF was prolonged compared to eucalyptol alone [173]. Mączka and colleagues comprehensively discussed the possibility of the replacement of antibiotics with eucalyptol in their article [171].

2.2.4. Eugenol

Eugenol (2-methoxy-4-[2-propenyl] phenol) belongs to the class of phenylpropanoids and is present in the essential oils obtained from Cinnamomum and clove [174]. Pereira et al., (2013) studied the antifungal activity of eugenol against Trichophyton rubrum, which is responsible for causing dermatophytosis [175]. Eugenol inhibited the growth of different strains of T. rubrum with MIC values ranging from 64–512 µg/mL. The inhibitory growth activity was found to be mediated by causing membrane abnormalities, which include short, twisted hyphae and a reduction in conidia formation [175]. In another study, eugenol was reported to impart antifungal activity against C. albicans by inhibiting the synthesis of ergosterol, inducing oxidative stress, promoting lipid peroxidation, and increasing membrane permeability [176]. Similarly, eugenol displayed antifungal activity against clinical isolates of C. glabrata (MIC value: 128 μg/mL) by the inhibition of biofilm formation [177]. In addition, eugenol also caused an increase in ROS generation, cell lysis, and ergosterol content in the plasma membrane and reduced the enzyme activities of catalase, phospholipase, and proteinase [177]. The gene expression analysis using qRT-PCR revealed that exposure of eugenol to C. glabrata differentially modulated the levels of ergosterol synthesis genes (ERG2, ERG3, ERG4, ERG10, and ERG11), sterol importer (AUS1), GPI-anchored cell wall protein (KRE1), 1,3-β-glucan synthase (FKS1), and multidrug transporter (CDR1). The expression of AUS1, KRE1, and FKS1 was reduced, whereas ERG2, ERG3, ERG10, ERG11, and CDR1 were increased upon eugenol treatment in C. glabrata [177]. The reduction in membrane potential and release of cytochrome c was also observed in C. glabrata cells treated with eugenol, which indicates the activation of apoptosis [177].
Similarly, eugenol was reported to have antifungal activity against C. gattii and C. neoformans with GMIC values of 200 and 187 mg/L, respectively. Eugenol altered cellular morphology, increased oxidative burst, and promoted lipid peroxidation in C. gattii and C. neoformans cells [178]. Eugenol, along with cinnamaldehyde, showed an additive effect and inhibited the growth of candida species such as C. albicans, C. glabrata, and Candida lusitaniae [179]. Many investigations reported the antibiofilm activity of eugenol against Candida species. Eugenol inhibited the single and mixed biofilms of fluconazole-resistant C. albicans [180]. El-Baz et al., (2021) performed a molecular docking analysis of eugenol against Als3, one of the adhesive proteins responsible for the adhesion of Candida cells to host or surfaces of medical devices, which subsequently results in biofilm formation [181]. Computational studies demonstrated that eugenol showed the highest binding capacity to Als3 compared to 1,8-cineole, 2-phenylthiolane, and cinnamaldehyde, implying that eugenol may interfere with the adhesion of Candida cells. Based on all of these investigations, eugenol can be investigated as a therapeutic agent against fungal infections in the clinical setting.

2.2.5. Geraniol

Geraniol (3,7-dimethylocta-trans-2,6-dien-1-ol) is a monoterpene alcohol and a major constituent of essential oils extracted from wild bergamot, rose, lavender, palmarosa, etc. [182]. Geraniol is commercially used as a fragrance material in deodorants and cosmetic products. It is reported to have anticancer activity against murine leukemia, hepatoma, and melanoma cells. Miron et al., (2014) studied the antifungal activity of geraniol against many dermatophytes (Trichophyton mentagrophytes, T. rubrum, Microsporum canis, and Microsporum gypseum) and yeasts (C. albicans, C. krusei, C. glabrata, C. tropicalis, C. parapsilosis, C. neoformans, Trichosporon asahii) [183]. Geraniol demonstrated potent antifungal activity against Microsporum strains and other dermatophytes with GMIC values of 19.5 and 25.4 µg/mL, respectively [183]. It also displayed moderate antifungal activity against yeasts compared to dermatophytes. The investigation of the mechanism of action of the antifungal properties of geraniol against T. asahii revealed the ability of binding of geraniol to ergosterol and subsequent membrane destabilization [183]. Sharma et al., (2016) reported the antifungal activity of geraniol against three Candida species, such as C. albicans, C. tropicalis, and C. glabrata with MIC values of 130 µg/mL, 80 µg/mL, and 130 µg/mL respectively [184]. Geraniol did not show significant toxicity as evidenced by a hemolytic assay, compared to fluconazole and amphotericin B [184]. Geraniol was found to be involved in the inhibition of H+-ATPase and cell disruption of membrane integrity by interfering in ergosterol biosynthesis. Similarly, Pereira et al., (2015) also studied the antifungal activity of geraniol against T. rubrum (MIC value: 16–256 µg/mL) and reported that geraniol causes damage to the cell wall and cell membrane through the inhibition of ergosterol biosynthesis [185]. However, Leite et al.,(2015) studied the antifungal activity of geraniol against C. albicans and reported that geraniol neither interacts with ergosterol nor the cell wall, and they demonstrated the inhibition of pseudo-hyphae and chlamydoconidia formation by geraniol [186]. The pseudo-hyphae formation provides a survival benefit to the fungus by evading the host’s phagocytic system and acts as one of the contributing factors for the virulence of Candida species. Dalleau et al.,(2008) showed that geraniol possesses antibiofilm activity in C. albicans, which inhibited more than 80% of biofilm formation [187].

2.2.6. Hibiscuslide C

Hibiscuslide C (1-formyl-2, 8-dihydroxy-7-methoxy-6-methylnaphthalene) is a phytochemical reported to be present in plants, such as Hibiscus taiwanensis and Abutilon theophrasti [188]. Hibiscuslide C showed antifungal activity against C. albicans, C. parapsilosis, Trichosporon beigelii, and Malassezia furfur, with MIC values of 5, 5–10, 10, and 5 µg/mL, respectively [188]. The mechanism of antifungal property of hibiscuslide C against C. albicans was found to be due to its involvement in membrane disruptive mechanisms, such as membrane depolarization and pore formation [188]. The same study also demonstrated that hibiscuslide C induces apoptosis in C. albicans via increased ROS generation, an increase in intracellular Ca2+, metacaspase activation, mitochondrial dysfunctions such as membrane depolarization, and the release of cytochrome c [189].

2.2.7. Magnoflorine

Magnoflorine is a phytochemical present in medicinal plants, such as Phellodendron amurense, Sinomenium acutum, Thalictrum isopyroides, Magnolia officinalis, and Berberis kansuensis. It is reported to possess many pharmacological properties, such as antidiabetic, anti-inflammatory, immunomodulatory, antioxidant, and antifungal activities [190]. Magnoflorine displayed antifungal activity against various Candida strains, such as C. albicans C. tropicalis, C. parapsilosis, and C. glabrata [191]. Magnoflorine also presented alpha-glucosidase inhibitory activity and antibiofilm activity at a concentration of 150 μM. In another study, magnoflorine demonstrated antifungal activity against dermatophytes, such as T. rubrum and T. mentagrophyte (MIC: 62.5 μg/mL) [192], and inhibited conidia formation, abrogated hyphal growth, and altered mycelia morphology (deformed growth, cytoplasmic contraction, and surface peeling) in T. rubrum [192]. In addition, magnoflorine also caused cell membrane damage, nuclear content leakage, decreased ergosterol content, and reduced the activities of squalene epoxidase and CYP51 [192].

2.2.8. Tea Saponin

Tea saponin is a phytochemical that belongs to an oleanane-type pentacyclic triterpene that is distributed in plants, such as Camellia oleifera and Camellia sinensis [193]. Tea saponin is present in the seed cake, which is obtained as the byproduct during the extraction of oil from tea or camellia seeds. Tea saponin is a natural surfactant used extensively in the food, chemical, pesticide, and cosmetic industries. Tea saponin is endowed with many pharmacological properties, such as antimicrobial, anti-inflammatory, antioxidant, and antiallergic properties [194]. Li et al., (2020) demonstrated the antifungal activity of tea saponin against different strains of C. albicans, on which it showed a moderate growth inhibition (MIC: 64 µg/mL) compared to fluconazole (MIC: 0.5-128 µg/mL) [195]. They also investigated the effect of tea saponin on the process of filamentation in C. albicans. The yeast to hyphal form transition is known as the filamentation process, which plays a critical role in the pathogenicity of C. albicans [195]. Li et al., (2020) found that tea saponin and fluconazole arrested the filamentation process in C. albicans until 12 h, at 64 µg/mL and 2 µg/mL concentrations, respectively, whereas the extensive filamentation process was observed at 9 h in the control and 16 µg/mL in tea saponin groups [195]. Biofilm formation is one of the unique mechanisms adapted by pathogens to acquire resistance against the host’s immune system and antimicrobial agents. Tea saponin inhibited 80% of biofilm formation at a concentration of 64 µg/mL in C. albicans. The same study also demonstrated that inhibition of filamentation and biofilm formation by tea saponin is due to a reduction in the level of cAMP in C. albicans [195]. The investigations by Yu et al., (2022) showed that tea saponin isolated from Camellia oleifera seed cake inhibited the growth of C. albicans, S. cerevisiae, and Penicillium with MIC values of 0.078, 0.156, and 0.156 mg/mL, respectively [196]. The antifungal activity of tea saponin is attributed to its involvement in cell membrane damage, a reduction in cell adhesion and aggregation, and antibiofilm activity in C. albicans. Transcriptomics analysis also revealed that hyphae- and biofilm-related genes, such as ALS3, ECE1, HWP1, EFG1, and UME6, are downregulated in the presence of tea saponin [196]. These studies suggest that tea saponin can be further studied and developed as a potential antifungal agent.

2.3. Antiviral Agents Derived from Plants

The knowledge of viral diseases in humans dates back to 1796, when Edward Jenner developed a vaccine against smallpox using the cowpox virus. In 1885, Louis Pasteur developed a vaccine against rabies [197]. In those days, the concept and existence of viruses were not known. In one of the early studies on the discovery of viruses, Dimitri Ivanovsky and Martinus Beijerinck identified an agent responsible for causing mosaic disease in tobacco plants. Studies by Ivanovsky and Beijerinck revealed that this infectious agent can pass through Chamberland ultrafilters. Beijerinck hypothesized that the ultrafilterable infectious agent is an infectious liquid that he called “contagium vivum fluidum” [198]. Subsequently, many ultrafilterable infectious agents (such as foot-and-mouth disease virus and myxoma virus in 1898, yellow fever virus in 1901, poliovirus in 1909, and many more) that are responsible for causing many diseases were identified. Ernst Ruska and Max Knol invented the electron microscope in 1933. Bodo von Borries, Helmut Ruska, and Ernst Ruska published the electron microscopic images of the mousepox virus and vaccinia virus in 1938, which laid the foundation for understanding the structure of viruses [199]. Thereafter, about 26 virus families that are known to infect humans have been identified, and about three to four new virus species infecting humans are being identified every year [200]. Human viruses implicate health impacts ranging from mild to life-threatening illnesses. As per the WHO statistics, the recent pandemic caused by the outbreak of the COVID-19 virus resulted in a mortality rate of 6,887,000 (as of 3 April 2023) [201]. The outbreak of Ebola viral infection in 2014 in West Africa resulted in a fatality rate of approximately 55% among the infected [202]. It is important to note that two-thirds of human pathogens are viruses. Viruses are highly adaptable biological entities that contribute to the emergence/reemergence of virus disease outbreaks. The study of the intricate “host-pathogen-environment” is crucial to understand these disease outbreaks [203]. It is important to discover novel antiviral compounds due to the high adaptability and rapid evolution of viral pathogens. In the following section, the antiviral efficacy of some of the selected plant-derived antiviral compounds has been discussed (Table 3). The structures of plant-derived compounds that are active against viruses are given as Figure 3.

2.3.1. Betulinic Acid

Betulinic acid is a pentacyclic lupane-type triterpenoid widely present in different plant species [204]. It is generally isolated from the Birch tree (Betula sp., Betulaceae), which has well-known medicinal applications. Betulinic acid is also present in plants belonging to the genera Ziziphus, Syzygium, Diospyros, and Paeonia [205]. Many investigations revealed the antiviral potency of betulinic acid against different viruses. The antiviral function of betulinic acid against influenza A/PR/8 virus-infected A549 (human lung cancer) cells was examined [206]. Betulinic acid (50 µM) displayed good antiviral activity (98%) against the influenza A/PR/8/34 virus in A549 cells without significant cytotoxicity towards host A549 cell lines. Betulinic acid (10 mg/kg/dose for seven days) administration attenuated pulmonary pathological symptoms, including necrosis, number of inflammatory cells, and pulmonary edema in influenza A/PR/8/34 virus-infected C57BL/6 mice [206]. In general, the influenza A/PR/8/34 virus infection triggers the upsurge of proinflammatory cytokines (IFN-γ, IL-1β, and TNF-α) in the host, which leads to severe pulmonary inflammation. Interestingly, betulinic acid reduced the levels of IFN-γ in influenza A/PR/8/34 virus-infected C57BL/6 mice, indicating that betulinic acid may assist the recovery of infected mice by reducing severe pulmonary inflammation [206].
In another study, betulinic acid was reported to impart antiviral activity against the dengue virus type 2 (DENV-2). Betulinic acid (5 and 10 μM) reduced the viral titer 1.4 log10 fold in DENV-2-infected Huh7 (human liver cancer) cells. In vitro studies indicated that a 50% cytotoxic concentration (CC50), a 50% inhibitory concentration (IC50), and selectivity index values of betulinic acid against DENV-2 were found to be 28.24 µM, 0.9463 µM, and 29.843, respectively. The antiviral activity of betulinic acid was found to be due to its involvement in the inhibition of the post-entry stage of the DENV-2 replication cycle, viral RNA synthesis, and viral protein production [207].

2.3.2. Guggulsterone

Guggulsterone is a phytosteroid present in the plant Commiphora gileadensis (L.), which is generally known as the “balsam of Mecca” [208]. C. gileadensis is known for its usage in the traditional Arabian medicinal system to treat urinary retention, jaundice, constipation, inflammatory disorders, and liver disorders [208]. This compound is also reported to be present in the plant Guggul tree (Commiphora mukul), and its medicinal values are well-documented in Ayurveda, a traditional Indian medicinal system [209]. Bouslama et al., (2019) studied the antiviral effect of methanolic extract of C. gileadensis leaves on two enveloped viruses (herpes simplex virus type 2 and respiratory syncytial virus type B) and two nonenveloped viruses (coxsackie virus B type 3 and adenovirus type 5). Methanolic extract of C. gileadensis leaves showed antiviral activity against enveloped viruses with an IC50 and a selectivity index of approximately 20 µg/mL and >10, respectively [208]. Subsequent bio-guided assays revealed that the leaf extract contains guggulsterone as the active compound. Chen et al., (2021) investigated the antiviral activity of guggulsterone against DENV and found that guggulsterone inhibited protein synthesis and RNA replication in DENV in a dose-dependent manner [210]. In vivo analysis in a ICR suckling mouse model demonstrated that guggulsterone stimulates the Nrf2-driven expression of heme oxygenase-1 to increase antiviral interferon response [210]. Hemeoxygenase-1 is a host antioxidant enzyme that breakdowns the heme ring into biliverdin. As per the previous reports, biliverdin inhibits DENV NS2B/NS3 protease activity, which is known to positively contribute to antiviral interferon response.

2.3.3. Salvianolic Acids

Salvianolic acids are the class of phytochemicals present in Salvia miltiorrhiza (Danshen). The medicinal properties of S. miltiorrhiza have been recorded in traditional Chinese medicine and it has been known to promote blood circulation. S. miltiorrhiza contains about 10 different salvianolic acids and all of them have a common core chemical structure known as Danshensu [(R)-3-(3,4-Dihydroxyphenyl)-2-hydroxypropanoic acid] [211]. Out of these types, salvianolic acids A, B, and C are reported to have antiviral activity against SARS-CoV-2. Salvianolic acids demonstrate antiviral activity by binding to the SARS-CoV-2 spike (S) protein [212]. S protein is present on the surface of SARS-CoV-2 and interacts with angiotensin-converting enzyme 2 (ACE2), which is present in the host cells, to mediate the viral entry into the cells. Structurally, the S protein has two subunits, namely, S1 and S2, which are structurally distinct. S1 has a receptor binding domain that is involved in establishing an interaction with ACE2 on the host cell membrane. The binding of SARS-CoV-2 to ACE2 induces a conformational change in the S1 subunit, leading to the exposure of the S2′ cleavage site in the S2 subunit. Subsequently, the S2′ site is cleaved either by transmembrane serine protease 2 (TMPRSS2) present in the cell membrane (cell surface entry), or by cathepsins in the endosomes (endosomal entry pathway), which mark the two distinct SARS-CoV-2 entry pathways. The cleavage of the S2′ site in either pathway leads to shedding of the S1 subunit and the exposure of the fusion peptide (FP) domain in the S2 subunit, which subsequently leads to the insertion of the FP domain into the host cell membrane to facilitate membrane fusion. Additionally, the HR2 domain of the S2 subunit folds back and interacts with the HR1 domain, resulting in the formation of a six-helix bundle structure that brings the two membranes in close proximity and leads to the membrane pore formation through which the viral genome is injected into the host cell [213,214]. Yang et al., (2020) developed the pseudovirus model system using the SARS-Cov-2 S protein and studied the effect of salvianolic acid C (Sal-C) on the viral entry process in the host cells [215]. Sal-C inhibited the viral entry into the ACE2-expressing HEK293T and Vero-E6 cells with IC50 values of 3.85 and 0.47 μM, respectively. It was also shown using the plaque reduction assay that Sal-C reduced the number of plaques in the ongoing infection model (rather than the post-infection model), in which authentic SARS-CoV-2 was used [215]. The formation of the six-helix bundle core by the HR1 and HR2 domains of the S protein is a crucial event in the fusion of SARS-CoV-2 to the host cells. To understand the anti-SARS-CoV-2 activity of Sal-C, the authors synthesized HR1P and HR2P peptides, which contain the interacting regions of the HR1 and HR2 fusion core. Circular-dichroism spectroscopic analysis was performed to understand the biophysical change in the mixture of HR1P and HR2P peptides and HR1P or HR2P peptides alone. HR1P and HR2P peptides formed a HR1P/HR2P complex and showed a typical α-helical conformation of the six-helix bundle. Interestingly, the dose-dependent addition of Sal-C disrupted the characteristic α-helical conformation of the six-helix bundle in the HR1P and HR2P mixture. In addition, the dose-dependent treatment of Sal-C decreased the concentration of the six-helix bundle, as evidenced by the native-PAGE analysis [215]. These data concretely presented that the antiviral activity of Sal-C is due to its involvement in the disruption of the six-helix bundle conformation and thereby the abrogation of viral entry. The antiviral efficacy of salvianolic acids (Sal-A, Sal-B, and Sal-C) against SARS-COV-2 was studied in a pseudovirus system. For this, ACE2-overexpressing HEK293T cells were infected with 2019-nCoV spike pseudovirus. Sal-B displayed superior inhibitory activity over Sal-A and Sal-C towards the 2019-nCoV spike pseudovirus entry ratio, with an EC50 value of 6.22, 11.31, and 10.14 μM, respectively [212]. The mechanistic analysis also revealed that Sal-A, Sal-B, and Sal-C can suppress the entry of 2019-nCoV spike pseudovirus into ACE2-overexpressing HEK293T cells by interacting with the RBD of the spike protein and ACE2 [212].

2.3.4. Silvestrol

Silvestrol (cyclopenta[b]benzofuran flavagline) is a secondary metabolite present in the species belonging to the Aglaia genus, and it is reported to have broad-spectrum antiviral activity against different viruses. Silvestrol imparts an antiviral function against the Ebola virus by inhibiting viral replication [216]. Silvestrol induces antitumor activity by binding to the eIF4A subunit of the eIF4F complex and thereby attenuates the translation of oncoproteins, such as c-MYC and PIM-1. The eIF4F complex contributes to the scanning of the 5′ untranslated region (UTR) of mRNA and the recognition of start codons by the ribosome to initiate translation. During Ebola infection, the virus delivers its RNA into the host cells, where viral transcription is initiated using it as a template. Additionally, the Ebola virus utilizes host cell machinery for the synthesis of viral proteins through a cap-dependent translation process. Biedenkopf and colleagues hypothesized that silvestrol can abrogate the viral translation as the viral mRNAs also contain 5′ cap and UTR regions (similar to eukaryotic cells) [216]. To determine the antiviral activity of silvestrol, Huh-7 cells were preincubated with silvestrol (10 nM), and then the cells were infected with the Ebola virus. The virus titers in the supernatant of infected cells were determined by performing TCID50 (50% tissue culture infectious dose) analysis by using Vero E6 cells, which demonstrated a significant reduction in the viral infection and the dose-dependent reduction in EBOV viral titer in the post-infection model. Silvestrol also inhibited the expression of viral proteins, such as VP40, NP, and GP proteins which could be due to the targeting of eIF4A by silvestrol [216]. In another study, 5′ UTRs of Ebola virus mRNAs were fused to a dual luciferase reporter plasmid (pFR_HCV_xb) containing the HSV-TK promoter and the firefly luciferase gene [217]. In this plasmid, hepatitis C virus internal ribosome entry site elements were placed downstream to the firefly luciferase gene, which helps in the translation of the Renilla luciferase gene, through the eIF4A-independent mechanism. The plasmid is transfected to HepG2 cells, and a dual-luciferase reporter assay was carried out. A decrease in the luciferase activity was observed for all Ebola viral 5’ UTR constructs upon treatment with silvestrol (10 nM), which emphasizes the significance of 5′ cap and UTR regions in the translation of Ebola viral proteins [217].
Table 3. List of phytochemicals that have demonstrated antiviral activity against human viruses.
Table 3. List of phytochemicals that have demonstrated antiviral activity against human viruses.
Sl. No.PhytocompoundSourcesMicroorganisms Affected by the Title Compound and Dose Mechanism of ActionRef.
1BerberineBerberis vulgaris, Berberis fremontii,
Hydrastis Canadensis
Chikungunya virus (EC50: 37.6–50.9 µM)Reduction in viral RNA and protein synthesis[218]
2BaicaleinPolygonatum sibiricum,
Scutellaria baicalensis
Japanese encephalitis virus (IC50: 14.28 µg/mL,
CC50: 115.2 ± 0.2 µg/mL)
ND[219,220]
3Rosmarinic acidSalvia miltiorrhizaEV-A71
(CC50: 327.68 ± 14.43 µM,
EC50: 31.57 ± 4.14–114 ± 4.10 µM,
SI: 2.87–10.36)
Interferes with virus-host receptor interaction [221]
4Raoulic acidRaoulia australisHRV 2
(CC50: 201.78 µg/mL,
IC50: <0.1 µg/mL, TI: 2017.8),

HRV 3 (CC50: 201.78 µg/mL,
IC50: 0.197 ± 0.11 µg/mL, TI: 1090.7),

CV B3
(CC50: 65.86 µg/mL
IC50: 0.337 ± 0.02, TI; 199.58),

CV B4
(CC50: 65.86
IC50: 0.40 ± 0.05, TI: 164.65),

EV 71
(CC50: 65.86 µg/mL,
IC50: <0.1, TI: >658.6)
Broad spectrum antiviral activity[222,223]
5Tetra-O-galloyl-β-d-glucose (TGG)Galla chinensisSARS-CoV (CC50: 1.08 mM, EC50: 4.5 μM, SI: 240)Interferes with viral entry into host cells[224]
6Saikosaponin B2Bupleurum spp., Heteromorpha spp., Scrophularia scorodoniaHCoV-229E
(IC50: 1.7 ± 0.1 mmol/L,
CC50: 383.3 ± 0.2 μmol/L, SI: 221.9)
Interferes in virus absorption and penetration into host cells[225]
7Patentiflorin AJustica gendarussaHIV (IC50: 24–37 nM, CC50: 75 μM)Inhibition of reverse transcriptase[226]
8OligonolLitchi chinensisInfluenza virus (H3N2)Inhibition of the proliferation of the influenza virus by blocking ROS-dependent
ERK phosphorylation
[227]
9PunicalaginPunica granatumInfluenza virus (H3N2)Inhibition of agglutination of RBCs[228]
103-hydroxy caruilignan CSwietenia macrophyllaHCV
(EC50: 10.5 ± 1.2 μM)
Inhibition of viral RNA and protein synthesis[229]
11LycorineLycoris radiate,
Narcissus pseudonarcissus
Zika virus
(CC50: 4.29–21 μM, EC50: 0.22–0.39 μM, SI: 19.5–54)
Inhibition of viral RNA synthesis and protein synthesis, inhibition of viral RDRP activity[230]
12QuercetinHouttuynia cordataHSV-1
(CC50: 485.69 μg/mL, EC50: 52.9 μg/mL, SI: 9.18)
Inhibition of viral entry and NF-κB activation[231]
HSV-2
(CC50: 485.69 μg/mL, EC50: 70.01 μg/mL, SI: 6.94)
ND
13ShikoninRadix LithospermiADV-3Inhibition of hexon protein expression[232]
14NaringeninCitrus paradisi, Citrus aurantium, Prunus cerasus, Solanum lycopersicum, Origanum vulgareHCVReduction in HCV secretion in infected cells[233]
15Ursolic acidOcimum basilicumCV B1 (CC50: 100.5 mg/L, EC50: 0.4 ± 0.1 mg/L, SI: 251.3),

EV 71 (CC50: 100.5 mg/L, EC50: 0.5 ± 0.2 mg/L, SI: 201)
Interferes in the viral replication phase [234]
HSV-1 (CC50: 100.5 mg/L, EC50: 6.6 ± 1.8 mg/L, SI: 15.2),

ADV-8 (CC50: 100.5 mg/L, EC50: 4.2 ± 0.3 mg/L, SI: 23.8)
ND
16MyricetinAbundant in fruit, vegetables, tea, berriesSARS-CoV-2Inhibition of SARS-CoV-2 Mpro activity[235]
17EmetineCephaelisipecacuanhaSARS-CoV-2 (CC50: 1603.8 nM, EC50: 0.147 nM, SI: 10,910.4)Inhibition of SARS-CoV-2 mRNA/eIF4E interaction[236]
18LadaneinMarrubium PeregrinumHCV (EC50: 2.54 μmol/L, toxic dose 50 %: 98.04 μmol/L)Interferes with virus entry into host cells[237]
19Samarangenin BLimonium sinenseHSV-1 (IC50: 11.4 ± 0.9 μM)Inhibition of HSV-1 α gene expression, inhibition of HSV-1 DNA synthesis, and structural protein expression[238]
20Pterocarnin APterocarya stenopteraHSV-2 (IC50: 5.4 ± 0.3 μM, CC50: 31.7 ± 1.6 μM, SI: 5.9)Inhibition of virus attachment and penetration into host cells and inhibition of virus replication [239]
Abbreviations: ADV: adenovirus; CV B: coxsackie virus B; ERK: extracellular signal-regulated kinase; EV: enterovirus; HCoV: human coronavirus; HCV: hepatitis C virus; HIV: human immunodeficiency virus; HRV: human rhinovirus; HSV: herpes simplex virus; ND: not determined; SARS-CoV-2: severe acute respiratory syndrome corona virus 2.

3. Synergistic Antimicrobial Effects of Plant Metabolites with Standard Antibiotics

The percentage of FDA-approved plant-derived antimicrobials is very insignificant (around 3%) compared to the abundance of plant metabolites [240]. Many traditional plant extract-based therapies involve the administration of a complex mixture of different phytochemicals that work in unison and may contribute to the synergistic effect to combat the growth of infectious microorganisms. Some researchers strongly believe that the synergistic potential of plant extract-based therapy might be a promising approach to address the rising antibiotic resistance [241]. In support of this, several plant-derived compounds have been demonstrated to potentiate the effect of antibiotics that are in clinical practice [242]. For instance, piperine, present in the Piper nigrum and Piper longum, inhibits bacterial efflux pumps to impart antibacterial activity. The nanoliposomes co-loaded with gentamicin and piperine showed synergistic antibacterial activity against MRSA and also reduced the MIC value of gentamicin about 32-fold [243]. Similarly, chanoclavine isolated from Ipomoea muricata also displayed bacterial efflux inhibition and presented a synergistic activity with tetracycline against MDR E. coli with a 16-fold reduction in the MIC of tetracycline [244]. Tomatidine, a secondary metabolite derived from the plants of tomato, potato, and eggplant, also demonstrated a synergistic effect with several aminoglycoside antibiotics against the MDR of S. aureus [245]. Thymol, a component of essential oil obtained from Thymus vulgaris and Origanum vulgare, displayed synergism with fluconazole against clinical isolates of Candida species such as C. albicans, C. glabrate, and C. krusei [246]. Epigallocatechin gallate (EGCG), a polyphenol present in tea leaves, showed synergistic antifungal activity with antimycotics such as miconazole, fluconazole, and amphotericin B against many Candida species [247]. These reports suggest that natural compounds obtained from plants can be used as potentiating agents of antimicrobial activity, and this fact can be considered in clinical trials.

4. Plant-Derived Drugs That Are in Clinical Practice for the Treatment of Human Ailments

Plants serve as an arsenal of secondary metabolites and their therapeutic applications against many infectious diseases are well-documented in ancient medical texts and paleobotanical findings at archeological sites [248,249]. Approximately 3% of natural products obtained from plants are approved by the FDA as antimicrobial agents, and an extensive portion of FDA-approved natural antimicrobial agents are obtained from microbes [240]. However, these reports may not reflect the true potential of phytochemicals as antimicrobial agents. According to the WHO, around 80% of the developing world is dependent on traditional medicine derived from medicinal plants [250]. In support of this, a huge number of drugs obtained from plants are in today’s clinical practice. Artemisinin, a phytochemical isolated from Artemisia annua, is widely used for the treatment of malaria, a life-threatening disease caused by P. falciparum. The discovery of artemisinin from plants is a breakthrough event in the research arena of plant-derived antimicrobial compounds. Apart from antimicrobials, many plant-derived compounds have also been developed as drugs against many human diseases. Approximately 25–28% of modern medicines are derived from plant sources [251]. For instance, galantamine, an isoquinoline alkaloid present in Galanthus nivalis and Galanthus woronowii, acts as an acetylcholinesterase inhibitor and is used in the treatment of Alzheimer’s disease [252,253]. Nitisinone is a chemical derivative of leptospermone, a phytochemical present in the plant Callistemon citrinus. Nitisinone is an inhibitor of 4-hydroxyphenylpyruvate dioxygenase and is used in the treatment of hereditary tyrosinemia type 1 [254]. Taxol, a blockbuster anticancer drug, was initially isolated from the bark of Taxus brevifolia, and was subsequently demonstrated to be produced by endophytes. Camptothecin is an approved drug that imparts an anticancer effect by inhibiting topoisomerase I and was initially identified to be produced by Camptotheca acuminata. Similarly, curcumin is a polyphenol present in the Curcuma longa and its medicinal applications are mentioned in ancient texts such as traditional Indian medicine and traditional Chinese medicine. It is considered a promising chemo-preventive agent against skin diseases such as psoriasis, vitiligo, and melanoma [255]. Curcumin is also reported to possess good antibacterial activity against different pathogenic microorganisms [256]. These examples provide a glimpse of the diverse therapeutic potential of phytochemicals. The logical drug repurposing approach also serves as an alternative approach for determining the antimicrobial activity of plant-derived drugs that are used against other diseases.

5. Conclusions and Future Perspectives

Antimicrobial resistance is one of the most serious health concerns across the globe, as many pathogens are rapidly developing resistance against existing antimicrobials. In the current scenario, there is no effective therapeutic agent with the potential to reverse antimicrobial resistance, and many leading laboratories are extensively working to discover new antimicrobials. Plant-based natural compounds are relatively less studied in the context of developing antimicrobial drugs. Natural compounds have been of great interest in the drug discovery process due to their structural diversity, chemical novelty, abundance, and bioactivity. Natural compounds have been isolated from various organisms, including bacteria, fungi, invertebrates, marine creatures, and plants. All of them have enormously contributed to the development of drugs against various human ailments. For instance, doxorubicin, bleomycins, epothilones, paclitaxel, camptothecin, podophyllotoxins, and vinca alkaloids are some of the well-known drugs derived either from bacteria, fungi, or plants [8,257,258]. In 2000, it was estimated that 57% of compounds that were undergoing clinical trials for cancer treatment were natural compounds [257]. It may be noted that many plant-derived metabolites have displayed antimicrobial activity against drug-resistant microorganisms, as discussed in the present article. A comprehensive investigation of the antimicrobial functions of plant metabolites needs to be carried out to explore their therapeutic potential. The plant metabolites can also be considered as scaffolds or template structures to chemically derivatize them to obtain compounds with improved antimicrobial efficacy. Additionally, the role of phytogenous compounds needs to be examined along with standard antibiotics to explore the possible synergistic effects. Overall, some plant metabolites have demonstrated good antimicrobial effects on clinically important microbes, and they could serve as future drugs against MDR pathogens.

Author Contributions

All the authors were involved in drafting and finalizing the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Council of Scientific and Industrial Research and the Indian Science Congress Association (ISCA) for providing an emeritus scientist fellowship and an Asutosh Mookerjee fellowship, respectively.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

K.S.R. and C.D.M. thank the Institution of Excellence, DST-PURSE, at the University of Mysore for providing infrastructure and other research facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Youssef, F.S.; Ashour, M.L.; Singab, A.N.B.; Wink, M. A Comprehensive Review of Bioactive Peptides from Marine Fungi and Their Biological Significance. Mar. Drugs 2019, 17, 559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Abdel-Razek, A.S.; El-Naggar, M.E.; Allam, A.; Morsy, O.M.; Othman, S.I. Microbial Natural Products in Drug Discovery. Processes 2020, 8, 470. [Google Scholar] [CrossRef] [Green Version]
  3. Hutchings, M.I.; Truman, A.W.; Wilkinson, B. Antibiotics: Past, present and future. Curr. Opin. Microbiol. 2019, 51, 72–80. [Google Scholar] [CrossRef]
  4. Prestinaci, F.; Pezzotti, P.; Pantosti, A. Antimicrobial resistance: A global multifaceted phenomenon. Pathog. Glob. Health 2015, 109, 309–318. [Google Scholar] [CrossRef] [Green Version]
  5. García, J.; García-Galán, M.J.; Day, J.W.; Boopathy, R.; White, J.R.; Wallace, S.; Hunter, R.G. A review of emerging organic contaminants (EOCs), antibiotic resistant bacteria (ARB), and antibiotic resistance genes (ARGs) in the environment: Increasing removal with wetlands and reducing environmental impacts. Bioresour. Technol. 2020, 307, 123228. [Google Scholar] [CrossRef]
  6. Rather, I.A.; Kim, B.-C.; Bajpai, V.K.; Park, Y.-H. Self-medication and antibiotic resistance: Crisis, current challenges, and prevention. Saudi J. Biol. Sci. 2017, 24, 808–812. [Google Scholar] [CrossRef]
  7. Subramaniam, G.; Girish, M. Antibiotic resistance—A cause for reemergence of infections. Indian J. Pediatr. 2020, 87, 937–944. [Google Scholar] [CrossRef] [PubMed]
  8. Jadimurthy, R.; Mayegowda, S.B.; Nayak, S.C.; Mohan, C.D.; Rangappa, K.S. Escaping mechanisms of ESKAPE pathogens from antibiotics and their targeting by natural compounds. Biotechnol. Rep. 2022, 34, e00728. [Google Scholar] [CrossRef] [PubMed]
  9. Mohan, C.D.; Rangappa, S.; Preetham, H.D.; Chandra Nayaka, S.; Gupta, V.K.; Basappa, S.; Sethi, G.; Rangappa, K.S. Targeting STAT3 signaling pathway in cancer by agents derived from Mother Nature. Semin. Cancer Biol. 2022, 80, 157–182. [Google Scholar] [CrossRef] [PubMed]
  10. Mohan, C.D.; Hari, S.; Preetham, H.D.; Rangappa, S.; Barash, U.; Ilan, N.; Nayak, S.C.; Gupta, V.K.; Basappa; Vlodavsky, I.; et al. Targeting Heparanase in Cancer: Inhibition by Synthetic, Chemically Modified, and Natural Compounds. iScience 2019, 15, 360–390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Mohan, C.D.; Rangappa, S.; Nayak, S.C.; Sethi, G.; Rangappa, K.S. Paradoxical functions of long noncoding RNAs in modulating STAT3 signaling pathway in hepatocellular carcinoma. Biochim. Biophys. Acta BBA—Rev. Cancer 2021, 1876, 188574. [Google Scholar] [CrossRef]
  12. Hegde, M.; Girisa, S.; Naliyadhara, N.; Kumar, A.; Alqahtani, M.S.; Abbas, M.; Mohan, C.D.; Warrier, S.; Hui, K.M.; Rangappa, K.S. Natural compounds targeting nuclear receptors for effective cancer therapy. Cancer Metastasis Rev. 2022, 1–58. [Google Scholar] [CrossRef]
  13. Mohan, C.D.; Yang, M.H.; Rangappa, S.; Chinnathambi, A.; Alharbi, S.A.; Alahmadi, T.A.; Deivasigamani, A.; Hui, K.M.; Sethi, G.; Rangappa, K.S.; et al. 3-Formylchromone Counteracts STAT3 Signaling Pathway by Elevating SHP-2 Expression in Hepatocellular Carcinoma. Biology 2022, 11, 29. [Google Scholar] [CrossRef]
  14. Mohan, C.D.; Kim, C.; Siveen, K.S.; Manu, K.A.; Rangappa, S.; Chinnathambi, A.; Alharbi, S.A.; Rangappa, K.S.; Kumar, A.P.; Ahn, K.S. Crocetin imparts antiproliferative activity via inhibiting STAT3 signaling in hepatocellular carcinoma. IUBMB Life 2021, 73, 1348–1362. [Google Scholar] [CrossRef]
  15. Mohan, C.D.; Liew, Y.Y.; Jung, Y.Y.; Rangappa, S.; Preetham, H.D.; Chinnathambi, A.; Alahmadi, T.A.; Alharbi, S.A.; Lin, Z.-X.; Rangappa, K.S.; et al. Brucein D modulates MAPK signaling cascade to exert multi-faceted anti-neoplastic actions against breast cancer cells. Biochimie 2021, 182, 140–151. [Google Scholar] [CrossRef]
  16. Kim, N.Y.; Mohan, C.D.; Chinnathambi, A.; Alharbi, S.A.; Sethi, G.; Rangappa, K.S.; Ahn, K.S. Euphorbiasteroid Abrogates EGFR and Wnt/beta-Catenin Signaling in Non-Small-Cell Lung Cancer Cells to Impart Anticancer Activity. Molecules 2022, 27, 3824. [Google Scholar] [CrossRef]
  17. Sin, Z.W.; Mohan, C.D.; Chinnathambi, A.; Govindasamy, C.; Rangappa, S.; Rangappa, K.S.; Jung, Y.Y.; Ahn, K.S. Leelamine Exerts Antineoplastic Effects in Association with Modulating Mitogen-Activated Protein Kinase Signaling Cascade. Nutr. Cancer 2022, 74, 3375–3387. [Google Scholar] [CrossRef] [PubMed]
  18. Tacconelli, E.; Carrara, E.; Savoldi, A.; Harbarth, S.; Mendelson, M.; Monnet, D.L.; Pulcini, C.; Kahlmeter, G.; Kluytmans, J.; Carmeli, Y.; et al. Discovery, research, and development of new antibiotics: The WHO priority list of antibiotic-resistant bacteria and tuberculosis. Lancet Infect. Dis. 2018, 18, 318–327. [Google Scholar] [CrossRef]
  19. Kashyap, D.; Sharma, A.; Tuli, H.S.; Sak, K.; Garg, V.K.; Buttar, H.S.; Setzer, W.N.; Sethi, G. Apigenin: A natural bioactive flavone-type molecule with promising therapeutic function. J. Funct. Foods 2018, 48, 457–471. [Google Scholar] [CrossRef]
  20. Wu, D.; Kong, Y.; Han, C.; Chen, J.; Hu, L.; Jiang, H.; Shen, X. D-Alanine:D-alanine ligase as a new target for the flavonoids quercetin and apigenin. Int. J. Antimicrob. Agents 2008, 32, 421–426. [Google Scholar] [CrossRef] [PubMed]
  21. Brown, A.R.; Ettefagh, K.A.; Todd, D.; Cole, P.S.; Egan, J.M.; Foil, D.H.; Graf, T.N.; Schindler, B.D.; Kaatz, G.W.; Cech, N.B. A mass spectrometry-based assay for improved quantitative measurements of efflux pump inhibition. PLoS ONE 2015, 10, e0124814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Tuli, H.S.; Garg, V.K.; Mehta, J.K.; Kaur, G.; Mohapatra, R.K.; Dhama, K.; Sak, K.; Kumar, A.; Varol, M.; Aggarwal, D.; et al. Licorice (Glycyrrhiza glabra L.)-Derived Phytochemicals Target Multiple Signaling Pathways to Confer Oncopreventive and Oncotherapeutic Effects. OncoTargets Ther. 2022, 15, 1419–1448. [Google Scholar] [CrossRef] [PubMed]
  23. Long, D.R.; Mead, J.; Hendricks, J.M.; Hardy, M.E.; Voyich, J.M. 18β-Glycyrrhetinic acid inhibits methicillin-resistant Staphylococcus aureus survival and attenuates virulence gene expression. Antimicrob. Agents Chemother. 2013, 57, 241–247. [Google Scholar] [CrossRef] [Green Version]
  24. Banik, K.; Ranaware, A.M.; Deshpande, V.; Nalawade, S.P.; Padmavathi, G.; Bordoloi, D.; Sailo, B.L.; Shanmugam, M.K.; Fan, L.; Arfuso, F.; et al. Honokiol for cancer therapeutics: A traditional medicine that can modulate multiple oncogenic targets. Pharmacol. Res. 2019, 144, 192–209. [Google Scholar] [CrossRef] [PubMed]
  25. Li, B.; Yin, F.; Zhao, X.; Guo, Y.; Wang, W.; Wang, P.; Zhu, H.; Yin, Y.; Wang, X. Colistin Resistance Gene mcr-1 Mediates Cell Permeability and Resistance to Hydrophobic Antibiotics. Front. Microbiol. 2020, 10, 3015. [Google Scholar] [CrossRef] [Green Version]
  26. Guo, Y.; Lv, X.; Wang, Y.; Zhou, Y.; Lu, N.; Deng, X.; Wang, J. Honokiol restores polymyxin susceptibility to MCR-1-positive pathogens both in vitro and in vivo. Appl. Environ. Microbiol. 2020, 86, e02346-19. [Google Scholar] [CrossRef]
  27. Guo, Y.; Hou, E.; Wen, T.; Yan, X.; Han, M.; Bai, L.-P.; Fu, X.; Liu, J.; Qin, S. Development of membrane-active honokiol/magnolol amphiphiles as potent antibacterial agents against methicillin-resistant Staphylococcus aureus (MRSA). J. Med. Chem. 2021, 64, 12903–12916. [Google Scholar] [CrossRef]
  28. Li, W.-L.; Zhao, X.-C.; Zhao, Z.-W.; Huang, Y.-J.; Zhu, X.-Z.; Meng, R.-Z.; Shi, C.; Yu, L.; Guo, N. In vitro antimicrobial activity of honokiol against Staphylococcus aureus in biofilm mode. J. Asian Nat. Prod. Res. 2016, 18, 1178–1185. [Google Scholar] [CrossRef]
  29. Guo, N.; Liu, Z.; Yan, Z.; Liu, Z.; Hao, K.; Liu, C.; Wang, J. Subinhibitory concentrations of Honokiol reduce α-Hemolysin (Hla) secretion by Staphylococcus aureus and the Hla-induced inflammatory response by inactivating the NLRP3 inflammasome. Emerg. Microbes Infect. 2019, 8, 707–716. [Google Scholar] [CrossRef] [Green Version]
  30. Chen, A.Y.; Chen, Y.C. A review of the dietary flavonoid, kaempferol on human health and cancer chemoprevention. Food Chem. 2013, 138, 2099–2107. [Google Scholar] [CrossRef] [Green Version]
  31. Huang, Y.-H.; Huang, C.-C.; Chen, C.-C.; Yang, K., Jr.; Huang, C.-Y. Inhibition of Staphylococcus aureus PriA helicase by flavonol kaempferol. Protein J. 2015, 34, 169–172. [Google Scholar] [CrossRef]
  32. Zhou, H.; Xu, M.; Guo, W.; Yao, Z.; Du, X.; Chen, L.; Sun, Y.; Shi, S.; Cao, J.; Zhou, T. The Antibacterial Activity of Kaempferol Combined with Colistin against Colistin-Resistant Gram-Negative Bacteria. Microbiol. Spectr. 2022, 10, e02265-22. [Google Scholar] [CrossRef]
  33. Sun, Z.; Li, Q.; Hou, R.; Sun, H.; Tang, Q.; Wang, H.; Hao, Z.; Kang, S.; Xu, T.; Wu, S. Kaempferol-3-O-glucorhamnoside inhibits inflammatory responses via MAPK and NF-κB pathways in vitro and in vivo. Toxicol. Appl. Pharmacol. 2019, 364, 22–28. [Google Scholar] [CrossRef] [PubMed]
  34. Ming, D.; Wang, D.; Cao, F.; Xiang, H.; Mu, D.; Cao, J.; Li, B.; Zhong, L.; Dong, X.; Zhong, X. Kaempferol inhibits the primary attachment phase of biofilm formation in Staphylococcus aureus. Front. Microbiol. 2017, 8, 2263. [Google Scholar] [CrossRef] [Green Version]
  35. Jeong, K.-W.; Lee, J.-Y.; Kim, Y.-M. Homology Modeling and Docking Study of β-Ketoacyl Acyl Carrier Protein Synthase Ⅲ from Enterococcus faecalis. Bull. Korean Chem. Soc. 2007, 28, 1335–1340. [Google Scholar]
  36. Tomar, A.; Broor, S.; Kaushik, S.; Bharara, T.; Arya, D. Synergistic effect of naringenin with conventional antibiotics against methicillin resistant Staphylococcus aureus. Eur. J. Mol. Clin. Med. 2021, 7, 2020. [Google Scholar]
  37. Khan, A.W.; Kotta, S.; Ansari, S.H.; Sharma, R.K.; Ali, J. Self-nanoemulsifying drug delivery system (SNEDDS) of the poorly water-soluble grapefruit flavonoid Naringenin: Design, characterization, in vitro and in vivo evaluation. Drug Deliv. 2015, 22, 552–561. [Google Scholar] [CrossRef]
  38. Dey, P.; Parai, D.; Banerjee, M.; Hossain, S.T.; Mukherjee, S.K. Naringin sensitizes the antibiofilm effect of ciprofloxacin and tetracycline against Pseudomonas aeruginosa biofilm. Int. J. Med. Microbiol. 2020, 310, 151410. [Google Scholar] [CrossRef] [PubMed]
  39. Husain, F.M.; Perveen, K.; Qais, F.A.; Ahmad, I.; Alfarhan, A.H.; El-Sheikh, M.A. Naringin inhibits the biofilms of metallo-β-lactamases (MβLs) producing Pseudomonas species isolated from camel meat. Saudi J. Biol. Sci. 2021, 28, 333–341. [Google Scholar] [CrossRef] [PubMed]
  40. Zhao, Y.; Liu, S. Bioactivity of naringin and related mechanisms. Die Pharm.-Int. J. Pharm. Sci. 2021, 76, 359–363. [Google Scholar]
  41. Zhang, J.; Jung, Y.Y.; Mohan, C.D.; Deivasigamani, A.; Chinnathambi, A.; Alharbi, S.A.; Rangappa, K.S.; Hui, K.M.; Sethi, G.; Ahn, K.S. Nimbolide enhances the antitumor effect of docetaxel via abrogation of the NF-κB signaling pathway in prostate cancer preclinical models. Biochim. Biophys. Acta BBA—Mol. Cell Res. 2022, 1869, 119344. [Google Scholar] [CrossRef]
  42. Blum, F.C.; Singh, J.; Merrell, D.S. In vitro activity of neem (Azadirachta indica) oil extract against Helicobacter pylori. J. Ethnopharmacol. 2019, 232, 236–243. [Google Scholar] [CrossRef] [PubMed]
  43. Wylie, M.R.; Windham, I.H.; Blum, F.C.; Wu, H.; Merrell, D.S. In vitro antibacterial activity of nimbolide against Helicobacter pylori. J. Ethnopharmacol. 2022, 285, 114828. [Google Scholar] [CrossRef] [PubMed]
  44. Sarkar, P.; Acharyya, S.; Banerjee, A.; Patra, A.; Thankamani, K.; Koley, H.; Bag, P.K. Intracellular, biofilm-inhibitory and membrane-damaging activities of nimbolide isolated from Azadirachta indica A. Juss (Meliaceae) against meticillin-resistant Staphylococcus aureus. J. Med. Microbiol. 2016, 65, 1205–1214. [Google Scholar] [CrossRef] [PubMed]
  45. Baek, S.H.; Ko, J.-H.; Lee, H.; Jung, J.; Kong, M.; Lee, J.-W.; Lee, J.; Chinnathambi, A.; Zayed, M.E.; Alharbi, S.A.; et al. Resveratrol inhibits STAT3 signaling pathway through the induction of SOCS-1: Role in apoptosis induction and radiosensitization in head and neck tumor cells. Phytomedicine 2016, 23, 566–577. [Google Scholar] [CrossRef] [PubMed]
  46. Gupta, D.S.; Gadi, V.; Kaur, G.; Chintamaneni, M.; Tuli, H.S.; Ramniwas, S.; Sethi, G. Resveratrol and Its Role in the Management of B-Cell Malignancies—A Recent Update. Biomedicines 2023, 11, 221. [Google Scholar] [CrossRef] [PubMed]
  47. Ma, D.S.; Tan, L.T.-H.; Chan, K.-G.; Yap, W.H.; Pusparajah, P.; Chuah, L.-H.; Ming, L.C.; Khan, T.M.; Lee, L.-H.; Goh, B.-H. Resveratrol—Potential antibacterial agent against foodborne pathogens. Front. Pharmacol. 2018, 9, 102. [Google Scholar] [CrossRef] [Green Version]
  48. Liu, L.; Yu, J.; Shen, X.; Cao, X.; Zhan, Q.; Guo, Y.; Yu, F. Resveratrol enhances the antimicrobial effect of polymyxin B on Klebsiella pneumoniae and Escherichia coli isolates with polymyxin B resistance. BMC Microbiol. 2020, 20, 306. [Google Scholar] [CrossRef]
  49. Subramanian, M.; Soundar, S.; Mangoli, S. DNA damage is a late event in resveratrol-mediated inhibition of Escherichia coli. Free Radic. Res. 2016, 50, 708–719. [Google Scholar] [CrossRef]
  50. Hwang, D.; Lim, Y.-H. Resveratrol antibacterial activity against Escherichia coli is mediated by Z-ring formation inhibition via suppression of FtsZ expression. Sci. Rep. 2015, 5, 10029. [Google Scholar] [CrossRef] [Green Version]
  51. Vestergaard, M.; Ingmer, H. Antibacterial and antifungal properties of resveratrol. Int. J. Antimicrob. Agents 2019, 53, 716–723. [Google Scholar] [CrossRef] [PubMed]
  52. Raneri, M.; Pinatel, E.; Peano, C.; Rampioni, G.; Leoni, L.; Bianconi, I.; Jousson, O.; Dalmasio, C.; Ferrante, P.; Briani, F. Pseudomonas aeruginosa mutants defective in glucose uptake have pleiotropic phenotype and altered virulence in non-mammal infection models. Sci. Rep. 2018, 8, 16912. [Google Scholar] [CrossRef] [PubMed]
  53. Falchi, F.A.; Borlotti, G.; Ferretti, F.; Pellegrino, G.; Raneri, M.; Schiavoni, M.; Caselli, A.; Briani, F. Sanguinarine Inhibits the 2-Ketogluconate Pathway of Glucose Utilization in Pseudomonas aeruginosa. Front. Microbiol. 2021, 12, 2552. [Google Scholar] [CrossRef] [PubMed]
  54. Obiang-Obounou, B.W.; Kang, O.-H.; Choi, J.-G.; Keum, J.-H.; Kim, S.-B.; Mun, S.-H.; Shin, D.-W.; Kim, K.W.; Park, C.-B.; Kim, Y.-G. The mechanism of action of sanguinarine against methicillin-resistant Staphylococcus aureus. J. Toxicol. Sci. 2011, 36, 277–283. [Google Scholar] [CrossRef] [Green Version]
  55. Obiang-Obounou, B.W.; Kang, O.-H.; Choi, J.-G.; Keum, J.-H.; Kim, S.-B.; Mun, S.-H.; Shin, D.-W.; Park, C.-B.; Kim, Y.-G.; Han, S.-H. In vitro potentiation of ampicillin, oxacillin, norfloxacin, ciprofloxacin, and vancomycin by sanguinarine against methicillin-resistant Staphylococcus aureus. Foodborne Pathog. Dis. 2011, 8, 869–874. [Google Scholar] [CrossRef]
  56. Kumar, S.; Mathew, S.O.; Aharwal, R.P.; Tulli, H.S.; Mohan, C.D.; Sethi, G.; Ahn, K.S.; Webber, K.; Sandhu, S.S.; Bishayee, A. Withaferin A: A Pleiotropic Anticancer Agent from the Indian Medicinal Plant Withania somnifera (L.) Dunal. Pharmaceuticals 2023, 16, 160. [Google Scholar] [CrossRef]
  57. Murugan, R.; Rajesh, R.; Seenivasan, B.; Haridevamuthu, B.; Sudhakaran, G.; Guru, A.; Rajagopal, R.; Kuppusamy, P.; Juliet, A.; Gopinath, P. Withaferin A targets the membrane of Pseudomonas aeruginosa and mitigates the inflammation in zebrafish larvae; an in vitro and in vivo approach. Microb. Pathog. 2022, 172, 105778. [Google Scholar] [CrossRef]
  58. Reiter, J.; Levina, N.; Van der Linden, M.; Gruhlke, M.; Martin, C.; Slusarenko, A.J. Diallylthiosulfinate (Allicin), a volatile antimicrobial from garlic (Allium sativum), kills human lung pathogenic bacteria, including MDR strains, as a vapor. Molecules 2017, 22, 1711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Siriyong, T.; Srimanote, P.; Chusri, S.; Yingyongnarongkul, B.-E.; Suaisom, C.; Tipmanee, V.; Voravuthikunchai, S.P. Conessine as a novel inhibitor of multidrug efflux pump systems in Pseudomonas aeruginosa. BMC Complement. Altern. Med. 2017, 17, 405. [Google Scholar] [CrossRef] [Green Version]
  60. Bisso Ndezo, B.; Tokam Kuaté, C.R.; Dzoyem, J.P. Synergistic antibiofilm efficacy of thymol and piperine in combination with three aminoglycoside antibiotics against Klebsiella pneumoniae biofilms. Can. J. Infect. Dis. Med. Microbiol. 2021, 2021, 7029944. [Google Scholar] [CrossRef]
  61. Sousa Silveira, Z.D.; Macêdo, N.S.; Sampaio dos Santos, J.F.; Sampaio de Freitas, T.; Rodrigues dos Santos Barbosa, C.; Júnior, D.L.D.S.; Muniz, D.F.; Castro de Oliveira, L.C.; Júnior, J.P.S.; Cunha, F.A.B.D. Evaluation of the antibacterial activity and efflux pump reversal of thymol and carvacrol against Staphylococcus aureus and their toxicity in Drosophila melanogaster. Molecules 2020, 25, 2103. [Google Scholar] [CrossRef] [PubMed]
  62. Mak, K.-K.; Kamal, M.; Ayuba, S.; Sakirolla, R.; Kang, Y.-B.; Mohandas, K.; Balijepalli, M.; Ahmad, S.; Pichika, M. A comprehensive review on eugenol’s antimicrobial properties and industry applications: A transformation from ethnomedicine to industry. Pharmacogn. Rev. 2019, 13, 1–9. [Google Scholar]
  63. Huang, Y.-Q.; Huang, G.-R.; Wu, M.-H.; Tang, H.-Y.; Huang, Z.-S.; Zhou, X.-H.; Yu, W.-Q.; Su, J.-W.; Mo, X.-Q.; Chen, B.-P. Inhibitory effects of emodin, baicalin, schizandrin and berberine on hefA gene: Treatment of Helicobacter pylori-induced multidrug resistance. World J. Gastroenterol. WJG 2015, 21, 4225. [Google Scholar] [CrossRef]
  64. Tyagi, P.; Singh, M.; Kumari, H.; Kumari, A.; Mukhopadhyay, K. Bactericidal activity of curcumin I is associated with damaging of bacterial membrane. PLoS ONE 2015, 10, e0121313. [Google Scholar] [CrossRef] [Green Version]
  65. Wang, S.; Yao, J.; Zhou, B.; Yang, J.; Chaudry, M.T.; Wang, M.; Xiao, F.; Li, Y.; Yin, W. Bacteriostatic effect of quercetin as an antibiotic alternative in vivo and its antibacterial mechanism in vitro. J. Food Prot. 2018, 81, 68–78. [Google Scholar] [CrossRef]
  66. Bazzaz, B.S.F.; Sarabandi, S.; Khameneh, B.; Hosseinzadeh, H. Effect of catechins, green tea extract and methylxanthines in combination with gentamicin against Staphylococcus aureus and Pseudomonas aeruginosa-combination therapy against resistant bacteria. J. Pharmacopunct. 2016, 19, 312–318. [Google Scholar] [CrossRef] [Green Version]
  67. Zhao, Y.; Qu, Y.; Tang, J.; Chen, J.; Liu, J. Tea Catechin Inhibits Biofilm Formation of Methicillin-Resistant S. aureus. J. Food Qual. 2021, 2021, 8873091. [Google Scholar] [CrossRef]
  68. Dong, J.; Zhang, D.; Li, J.; Liu, Y.; Zhou, S.; Yang, Y.; Xu, N.; Yang, Q.; Ai, X. Genistein Inhibits the Pathogenesis of Aeromonas hydrophila by Disrupting Quorum Sensing Mediated Biofilm Formation and Aerolysin Production. Front. Pharmacol. 2021, 12, 753581. [Google Scholar] [CrossRef] [PubMed]
  69. Wu, M.; Tian, L.; Fu, J.; Liao, S.; Li, H.; Gai, Z.; Gong, G. Antibacterial mechanism of Protocatechuic acid against Yersinia enterocolitica and its application in pork. Food Control 2022, 133, 108573. [Google Scholar] [CrossRef]
  70. Kang, J.; Liu, L.; Liu, M.; Wu, X.; Li, J. Antibacterial activity of gallic acid against Shigella flexneri and its effect on biofilm formation by repressing mdoH gene expression. Food Control 2018, 94, 147–154. [Google Scholar] [CrossRef]
  71. Borges, A.; Ferreira, C.; Saavedra, M.J.; Simões, M. Antibacterial activity and mode of action of ferulic and gallic acids against pathogenic bacteria. Microb. Drug Resist. 2013, 19, 256–265. [Google Scholar] [CrossRef]
  72. Jeyanthi, V.; Velusamy, P.; Kumar, G.V.; Kiruba, K. Effect of naturally isolated hydroquinone in disturbing the cell membrane integrity of Pseudomonas aeruginosa MTCC 741 and Staphylococcus aureus MTCC 740. Heliyon 2021, 7, e07021. [Google Scholar] [CrossRef] [PubMed]
  73. Zhou, Y.; Wang, J.; Guo, Y.; Liu, X.; Liu, S.; Niu, X.; Wang, Y.; Deng, X. Discovery of a potential MCR-1 inhibitor that reverses polymyxin activity against clinical mcr-1-positive Enterobacteriaceae. J. Infect. 2019, 78, 364–372. [Google Scholar] [CrossRef] [PubMed]
  74. Jeong, K.-W.; Lee, J.-Y.; Kang, D.-I.; Lee, J.-U.; Shin, S.Y.; Kim, Y. Screening of flavonoids as candidate antibiotics against Enterococcus faecalis. J. Nat. Prod. 2009, 72, 719–724. [Google Scholar] [CrossRef]
  75. Alves, F.S.; Cruz, J.N.; de Farias Ramos, I.N.; do Nascimento Brandão, D.L.; Queiroz, R.N.; da Silva, G.V.; da Silva, G.V.; Dolabela, M.F.; da Costa, M.L.; Khayat, A.S. Evaluation of Antimicrobial Activity and Cytotoxicity Effects of Extracts of Piper nigrum L. and Piperine. Separations 2023, 10, 21. [Google Scholar] [CrossRef]
  76. Mun, S.-H.; Joung, D.-K.; Kim, S.-B.; Park, S.-J.; Seo, Y.-S.; Gong, R.; Choi, J.-G.; Shin, D.-W.; Rho, J.-R.; Kang, O.-H. The mechanism of antimicrobial activity of sophoraflavanone B against methicillin-resistant Staphylococcus aureus. Foodborne Pathog. Dis. 2014, 11, 234–239. [Google Scholar] [CrossRef]
  77. Togashi, N.; Inoue, Y.; Hamashima, H.; Takano, A. Effects of two terpene alcohols on the antibacterial activity and the mode of action of farnesol against Staphylococcus aureus. Molecules 2008, 13, 3069–3076. [Google Scholar] [CrossRef] [Green Version]
  78. Tiwari, V.; Roy, R.; Tiwari, M. Antimicrobial active herbal compounds against Acinetobacter baumannii and other pathogens. Front. Microbiol. 2015, 6, 618. [Google Scholar] [CrossRef] [Green Version]
  79. De, R.; Sarkar, A.; Ghosh, P.; Ganguly, M.; Karmakar, B.C.; Saha, D.R.; Halder, A.; Chowdhury, A.; Mukhopadhyay, A.K. Antimicrobial activity of ellagic acid against Helicobacter pylori isolates from India and during infections in mice. J. Antimicrob. Chemother. 2018, 73, 1595–1603. [Google Scholar] [CrossRef] [Green Version]
  80. Manso, T.; Lores, M.; de Miguel, T. Antimicrobial Activity of Polyphenols and Natural Polyphenolic Extracts on Clinical Isolates. Antibiotics 2021, 11, 46. [Google Scholar] [CrossRef]
  81. Becker, H.; Scher, J.M.; Speakman, J.-B.; Zapp, J. Bioactivity guided isolation of antimicrobial compounds from Lythrum salicaria. Fitoterapia 2005, 76, 580–584. [Google Scholar] [CrossRef]
  82. Mailafiya, M.M.; Yusuf, A.J.; Abdullahi, M.I.; Aleku, G.A.; Ibrahim, I.A.; Yahaya, M.; Abubakar, H.; Sanusi, A.; Adamu, H.W.; Alebiosu, C.O. Antimicrobial activity of stigmasterol from the stem bark of Neocarya macrophylla. J. Med. Plants Econ. Dev. 2018, 2, 1–5. [Google Scholar]
  83. Tan, S.; Yan, F.; Li, Q.; Liang, Y.; Yu, J.; Li, Z.; He, F.; Li, R.; Li, M. Chlorogenic acid promotes autophagy and alleviates Salmonella Typhimurium infection through the lncRNAGAS5/miR-23a/PTEN axis and the p38 MAPK pathway. Front. Cell Dev. Biol. 2020, 8, 552020. [Google Scholar] [CrossRef]
  84. Fan, Q.; Yuan, Y.; Jia, H.; Zeng, X.; Wang, Z.; Hu, Z.; Gao, Z.; Yue, T. Antimicrobial and anti-biofilm activity of thymoquinone against Shigella flexneri. Appl. Microbiol. Biotechnol. 2021, 105, 4709–4718. [Google Scholar] [CrossRef] [PubMed]
  85. Jain, N.; Nadgauda, R.S. Commiphora wightii (Arnott) Bhandari—A natural source of guggulsterone: Facing a high risk of extinction in its natural habitat. Am. J. Plant Sci. 2013, 4, 33323. [Google Scholar] [CrossRef] [Green Version]
  86. Qu, Q.; Wang, J.; Cui, W.; Zhou, Y.; Xing, X.; Che, R.; Liu, X.; Chen, X.; Bello-Onaghise, G.S.; Dong, C. In vitro activity and In vivo efficacy of Isoliquiritigenin against Staphylococcus xylosus ATCC 700404 by IGPD target. PLoS ONE 2019, 14, e0226260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Padilla-Montaño, N.; de León Guerra, L.; Moujir, L. Antimicrobial Activity and Mode of Action of Celastrol, a Nortriterpen Quinone Isolated from Natural Sources. Foods 2021, 10, 591. [Google Scholar] [CrossRef]
  88. Chen, B.-C.; Ding, Z.-S.; Dai, J.-S.; Chen, N.-P.; Gong, X.-W.; Ma, L.-F.; Qian, C.-D. New insights into the antibacterial mechanism of Cryptotanshinone, a representative diterpenoid quinone from Salvia miltiorrhiza bunge. Front. Microbiol. 2021, 12, 647289. [Google Scholar] [CrossRef]
  89. Yuan, Z.; Ouyang, P.; Gu, K.; Rehman, T.; Zhang, T.; Yin, Z.; Fu, H.; Lin, J.; He, C.; Shu, G. The antibacterial mechanism of oridonin against methicillin-resistant Staphylococcus aureus (MRSA). Pharm. Biol. 2019, 57, 710–716. [Google Scholar] [CrossRef] [Green Version]
  90. Liu, T.; Pan, Y.; Lai, R. New mechanism of magnolol and honokiol from Magnolia officinalis against Staphylococcus aureus. Nat. Prod. Commun. 2014, 9, 1934578X1400900922. [Google Scholar] [CrossRef] [Green Version]
  91. Chiu, K.-C.; Shih, Y.-H.; Wang, T.-H.; Lan, W.-C.; Li, P.-J.; Jhuang, H.-S.; Hsia, S.-M.; Shen, Y.-W.; Chen, M.Y.-C.; Shieh, T.-M. In vitro antimicrobial and antipro-inflammation potential of honokiol and magnolol against oral pathogens and macrophages. J. Formos. Med. Assoc. 2021, 120, 827–837. [Google Scholar] [CrossRef] [PubMed]
  92. Choi, S.-S.; Lee, S.-H.; Lee, K.-A. A Comparative Study of Hesperetin, Hesperidin and Hesperidin Glucoside: Antioxidant, Anti-Inflammatory, and Antibacterial Activities In Vitro. Antioxidants 2022, 11, 1618. [Google Scholar] [CrossRef]
  93. Hochfellner, C.; Evangelopoulos, D.; Zloh, M.; Wube, A.; Guzman, J.; McHugh, T.; Kunert, O.; Bhakta, S.; Bucar, F. Antagonistic effects of indoloquinazoline alkaloids on antimycobacterial activity of evocarpine. J. Appl. Microbiol. 2015, 118, 864–872. [Google Scholar] [CrossRef] [PubMed]
  94. Qian, W.; Wang, W.; Zhang, J.; Wang, T.; Liu, M.; Yang, M.; Sun, Z.; Li, X.; Li, Y. Antimicrobial and antibiofilm activities of ursolic acid against carbapenem-resistant Klebsiella pneumoniae. J. Antibiot. 2020, 73, 382–391. [Google Scholar] [CrossRef] [PubMed]
  95. Pang, D.; Liao, S.; Wang, W.; Mu, L.; Li, E.; Shen, W.; Liu, F.; Zou, Y. Destruction of the cell membrane and inhibition of cell phosphatidic acid biosynthesis in Staphylococcus aureus: An explanation for the antibacterial mechanism of morusin. Food Funct. 2019, 10, 6438–6446. [Google Scholar] [CrossRef]
  96. Xu, Z.; Li, K.; Pan, T.; Liu, J.; Li, B.; Li, C.; Wang, S.; Diao, Y.; Liu, X. Lonicerin, an anti-algE flavonoid against Pseudomonas aeruginosa virulence screened from Shuanghuanglian formula by molecule docking based strategy. J. Ethnopharmacol. 2019, 239, 111909. [Google Scholar] [CrossRef]
  97. Ouyang, J.; Sun, F.; Feng, W.; Xie, Y.; Ren, L.; Chen, Y. Antimicrobial activity of galangin and its effects on murein hydrolases of vancomycin-intermediate Staphylococcus aureus (VISA) strain Mu50. Chemotherapy 2018, 63, 20–28. [Google Scholar] [CrossRef]
  98. Appalasamy, S.; Lo, K.Y.; Ch’ng, S.J.; Nornadia, K.; Othman, A.S.; Chan, L.-K. Antimicrobial activity of artemisinin and precursor derived from in vitro plantlets of Artemisia annua L. BioMed Res. Int. 2014, 2014, 215872. [Google Scholar] [CrossRef] [Green Version]
  99. Xu, Y.; Shi, C.; Wu, Q.; Zheng, Z.; Liu, P.; Li, G.; Peng, X.; Xia, X. Antimicrobial activity of punicalagin against Staphylococcus aureus and its effect on biofilm formation. Foodborne Pathog. Dis. 2017, 14, 282–287. [Google Scholar] [CrossRef]
  100. Li, T.; Lu, Y.; Zhang, H.; Wang, L.; Beier, R.C.; Jin, Y.; Wang, W.; Li, H.; Hou, X. Antibacterial Activity and Membrane-Targeting Mechanism of Aloe-Emodin against Staphylococcus epidermidis. Front. Microbiol. 2021, 12, 621866. [Google Scholar] [CrossRef]
  101. Solnier, J.; Martin, L.; Bhakta, S.; Bucar, F. Flavonoids as novel efflux pump inhibitors and antimicrobials against both environmental and pathogenic intracellular mycobacterial species. Molecules 2020, 25, 734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Wang, S.; Feng, Y.; Han, X.; Cai, X.; Yang, L.; Liu, C.; Shen, L. Inhibition of virulence factors and biofilm formation by Wogonin attenuates pathogenicity of Pseudomonas aeruginosa PAO1 via targeting pqs quorum-sensing system. Int. J. Mol. Sci. 2021, 22, 12699. [Google Scholar] [CrossRef] [PubMed]
  103. Fahey, J.W.; Stephenson, K.K.; Wade, K.L.; Talalay, P. Urease from Helicobacter pylori is inactivated by sulforaphane and other isothiocyanates. Biochem. Biophys. Res. Commun. 2013, 435, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Ghosh, J.; Sil, P.C. Arjunolic acid: A new multifunctional therapeutic promise of alternative medicine. Biochimie 2013, 95, 1098–1109. [Google Scholar] [CrossRef] [PubMed]
  105. Djoukeng, J.; Abou-Mansour, E.; Tabacchi, R.; Tapondjou, A.; Bouda, H.; Lontsi, D. Antibacterial triterpenes from Syzygium guineense (Myrtaceae). J. Ethnopharmacol. 2005, 101, 283–286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Ramos, G.F.; Amponsah, I.K.; Harley, B.K.; Jibira, Y.; Baah, M.K.; Adjei, S.; Armah, F.A.; Mensah, A.Y. Triterpenoids mediate the antimicrobial, antioxidant, and anti-inflammatory activities of the stem bark of Reissantia indica. J. Appl. Pharm. Sci. 2021, 11, 039–048. [Google Scholar]
  107. Harnvoravongchai, P.; Chankhamhaengdecha, S.; Ounjai, P.; Singhakaew, S.; Boonthaworn, K.; Janvilisri, T. Antimicrobial effect of asiatic acid against Clostridium difficile is associated with disruption of membrane permeability. Front. Microbiol. 2018, 9, 2125. [Google Scholar] [CrossRef]
  108. Guzman, J.D. Natural cinnamic acids, synthetic derivatives and hybrids with antimicrobial activity. Molecules 2014, 19, 19292–19349. [Google Scholar] [CrossRef]
  109. Alves, M.J.; Ferreira, I.C.; Froufe, H.J.; Abreu, R.; Martins, A.; Pintado, M. Antimicrobial activity of phenolic compounds identified in wild mushrooms, SAR analysis and docking studies. J. Appl. Microbiol. 2013, 115, 346–357. [Google Scholar] [CrossRef]
  110. Kępa, M.; Miklasińska-Majdanik, M.; Wojtyczka, R.D.; Idzik, D.; Korzeniowski, K.; Smoleń-Dzirba, J.; Wąsik, T.J. Antimicrobial potential of caffeic acid against Staphylococcus aureus clinical strains. BioMed Res. Int. 2018, 2018, 7413504. [Google Scholar] [CrossRef] [Green Version]
  111. Zhang, L.; Bao, M.; Liu, B.; Zhao, H.; Zhang, Y.; Ji, X.; Zhao, N.; Zhang, C.; He, X.; Yi, J. Effect of andrographolide and its analogs on bacterial infection: A review. Pharmacology 2020, 105, 123–134. [Google Scholar] [CrossRef] [PubMed]
  112. Thammawithan, S.; Talodthaisong, C.; Srichaiyapol, O.; Patramanon, R.; Hutchison, J.A.; Kulchat, S. Andrographolide stabilized-silver nanoparticles overcome ceftazidime-resistant Burkholderia pseudomallei: Study of antimicrobial activity and mode of action. Sci. Rep. 2022, 12, 10701. [Google Scholar] [CrossRef] [PubMed]
  113. Cong, S.; Tong, Q.; Peng, Q.; Shen, T.; Zhu, X.; Xu, Y.; Qi, S. In vitro anti-bacterial activity of diosgenin on Porphyromonas gingivalis and Prevotella intermedia. Mol. Med. Rep. 2020, 22, 5392–5398. [Google Scholar] [CrossRef] [PubMed]
  114. Nguyen, A.T.; Kim, K.-Y. Rhein inhibits the growth of Propionibacterium acnes by blocking NADH dehydrogenase-2 activity. J. Med. Microbiol. 2020, 69, 689–696. [Google Scholar] [CrossRef]
  115. Gong, R.; Lee, D.Y.; Lee, J.W.; Choi, D.J.; Kim, G.-S.; Lee, S.H.; Lee, Y.-S. Potentiating activity of rhein in targeting of resistance genes in methicillin-resistant Staphylococcus aureus. Asian Pac. J. Trop. Med. 2019, 12, 14. [Google Scholar]
  116. Morita, D.; Sawada, H.; Ogawa, W.; Miyachi, H.; Kuroda, T. Riccardin C derivatives cause cell leakage in Staphylococcus aureus. Biochim. Biophys. Acta BBA-Biomembr. 2015, 1848, 2057–2064. [Google Scholar] [CrossRef] [Green Version]
  117. Qian, Y.; Xia, L.; Wei, L.; Li, D.; Jiang, W. Artesunate inhibits Staphylococcus aureus biofilm formation by reducing alpha-toxin synthesis. Arch. Microbiol. 2021, 203, 707–717. [Google Scholar] [CrossRef]
  118. Oloyede, H.; Ajiboye, H.; Salawu, M.; Ajiboye, T. Influence of oxidative stress on the antibacterial activity of betulin, betulinic acid and ursolic acid. Microb. Pathog. 2017, 111, 338–344. [Google Scholar] [CrossRef]
  119. Zhang, L.; Kong, Y.; Wu, D.; Zhang, H.; Wu, J.; Chen, J.; Ding, J.; Hu, L.; Jiang, H.; Shen, X. Three flavonoids targeting the β-hydroxyacyl-acyl carrier protein dehydratase from Helicobacter pylori: Crystal structure characterization with enzymatic inhibition assay. Protein Sci. 2008, 17, 1971–1978. [Google Scholar] [CrossRef] [Green Version]
  120. Didry, N.; Dubreuil, L.; Pinkas, M. Microbiological properties of protoanemonin isolated from Ranunculus bulbosus. Phytother. Res. 1993, 7, 21–24. [Google Scholar] [CrossRef]
  121. Marini, E.; Magi, G.; Mingoia, M.; Pugnaloni, A.; Facinelli, B. Antimicrobial and anti-virulence activity of capsaicin against erythromycin-resistant, cell-invasive group a streptococci. Front. Microbiol. 2015, 6, 1281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Agarwal, P.; Das, C.; Dias, O.; Shanbhag, T. Antimicrobial property of capsaicin. Int. Res. J. Biol. Sci. 2017, 6, 7–11. [Google Scholar]
  123. Goel, S.; Mishra, P. Thymoquinone inhibits biofilm formation and has selective antibacterial activity due to ROS generation. Appl. Microbiol. Biotechnol. 2018, 102, 1955–1967. [Google Scholar] [CrossRef] [PubMed]
  124. Chaieb, K.; Kouidhi, B.; Jrah, H.; Mahdouani, K.; Bakhrouf, A. Antibacterial activity of Thymoquinone, an active principle of Nigella sativa and its potency to prevent bacterial biofilm formation. BMC Complement. Altern. Med. 2011, 11, 29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Nijampatnam, B.; Zhang, H.; Cai, X.; Michalek, S.M.; Wu, H.; Velu, S.E. Inhibition of Streptococcus mutans biofilms by the natural stilbene piceatannol through the inhibition of glucosyltransferases. ACS Omega 2018, 3, 8378–8385. [Google Scholar] [CrossRef] [Green Version]
  126. Teow, S.-Y.; Liew, K.; Ali, S.A.; Khoo, A.S.-B.; Peh, S.-C. Antibacterial action of curcumin against Staphylococcus aureus: A brief review. J. Trop. Med. 2016, 2016, 2853045. [Google Scholar] [CrossRef] [Green Version]
  127. Parai, D.; Banerjee, M.; Dey, P.; Mukherjee, S.K. Reserpine attenuates biofilm formation and virulence of Staphylococcus aureus. Microb. Pathog. 2020, 138, 103790. [Google Scholar] [CrossRef]
  128. Begum, S.; Naqvi, S.Q.Z.; Ahmed, A.; Tauseef, S.; Siddiqui, B.S. Antimycobacterial and antioxidant activities of reserpine and its derivatives. Nat. Prod. Res. 2012, 26, 2084–2088. [Google Scholar] [CrossRef]
  129. Lamontagne Boulet, M.; Isabelle, C.; Guay, I.; Brouillette, E.; Langlois, J.-P.; Jacques, P.-É.; Rodrigue, S.; Brzezinski, R.; Beauregard, P.B.; Bouarab, K. Tomatidine is a lead antibiotic molecule that targets Staphylococcus aureus ATP synthase subunit C. Antimicrob. Agents Chemother. 2018, 62, e02197-17. [Google Scholar] [CrossRef] [Green Version]
  130. Gaur, R.; Gupta, V.K.; Singh, P.; Pal, A.; Darokar, M.P.; Bhakuni, R.S. Drug Resistance Reversal Potential of Isoliquiritigenin and Liquiritigenin Isolated from Glycyrrhiza glabra Against Methicillin-Resistant Staphylococcus aureus (MRSA). Phytother. Res. 2016, 30, 1708–1715. [Google Scholar] [CrossRef]
  131. Brown, A.K.; Papaemmanouil, A.; Bhowruth, V.; Bhatt, A.; Dover, L.G.; Besra, G.S. Flavonoid inhibitors as novel antimycobacterial agents targeting Rv0636, a putative dehydratase enzyme involved in Mycobacterium tuberculosis fatty acid synthase II. Microbiology 2007, 153, 3314–3322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Nitiema, L.W.; Savadogo, A.; Simpore, J.; Dianou, D.; Traore, A.S. In vitro antimicrobial activity of some phenolic compounds (coumarin and quercetin) against gastroenteritis bacterial strains. Int. J. Microbiol. Res. 2012, 3, 183–187. [Google Scholar]
  133. Periasamy, H.; Iswarya, S.; Pavithra, N.; Senthilnathan, S.; Gnanamani, A. In vitro antibacterial activity of plumbagin isolated from Plumbago zeylanica L. against methicillin-resistant Staphylococcus aureus. Lett. Appl. Microbiol. 2019, 69, 41–49. [Google Scholar] [PubMed]
  134. Djeussi, D.E.; Noumedem, J.A.; Seukep, J.A.; Fankam, A.G.; Voukeng, I.K.; Tankeo, S.B.; Nkuete, A.H.; Kuete, V. Antibacterial activities of selected edible plants extracts against multidrug-resistant Gram-negative bacteria. BMC Complement. Altern. Med. 2013, 13, 164. [Google Scholar] [CrossRef] [Green Version]
  135. Miyasaki, Y.; Rabenstein, J.D.; Rhea, J.; Crouch, M.-L.; Mocek, U.M.; Kittell, P.E.; Morgan, M.A.; Nichols, W.S.; Van Benschoten, M.; Hardy, W.D. Isolation and characterization of antimicrobial compounds in plant extracts against multidrug-resistant Acinetobacter baumannii. PLoS ONE 2013, 8, e61594. [Google Scholar] [CrossRef] [Green Version]
  136. Firacative, C. Invasive fungal disease in humans: Are we aware of the real impact? Memórias Inst. Oswaldo Cruz 2020, 115, e200430. [Google Scholar] [CrossRef]
  137. Bongomin, F.; Gago, S.; Oladele, R.O.; Denning, D.W. Global and Multi-National Prevalence of Fungal Diseases—Estimate Precision. J. Fungi 2017, 3, 57. [Google Scholar] [CrossRef]
  138. Kainz, K.; Bauer, M.A.; Madeo, F.; Carmona-Gutierrez, D. Fungal infections in humans: The silent crisis. Microb. Cell 2020, 7, 143. [Google Scholar] [CrossRef]
  139. Alastruey-Izquierdo, A.; World Health Organization. WHO Fungal Priority Pathogens List to Guide Research, Development and Public Health Action; World Health Organization: Geneva, Switzerland, 2022. [Google Scholar]
  140. World Health Organization. First Meeting of the WHO Antifungal Expert Group on Identifying Priority Fungal Pathogens: Meeting Report; World Health Organization: Geneva, Switzerland, 2020. [Google Scholar]
  141. Normile, T.G.; Bryan, A.M.; Del Poeta, M. Animal models of Cryptococcus neoformans in identifying immune parameters associated with primary infection and reactivation of latent infection. Front. Immunol. 2020, 11, 581750. [Google Scholar] [CrossRef]
  142. Okoye, C.A.; Nweze, E.; Ibe, C. Invasive candidiasis in Africa, what is the current picture? Pathog. Dis. 2022, 80, ftac012. [Google Scholar] [CrossRef]
  143. Zotti, M.; Colaianna, M.; Morgese, M.G.; Tucci, P.; Schiavone, S.; Avato, P.; Trabace, L. Carvacrol: From ancient flavoring to neuromodulatory agent. Molecules 2013, 18, 6161–6172. [Google Scholar] [CrossRef] [PubMed]
  144. Niu, C.; Wang, C.; Yang, Y.; Chen, R.; Zhang, J.; Chen, H.; Zhuge, Y.; Li, J.; Cheng, J.; Xu, K. Carvacrol induces Candida albicans apoptosis associated with Ca2+/calcineurin pathway. Front. Cell. Infect. Microbiol. 2020, 10, 192. [Google Scholar] [CrossRef] [PubMed]
  145. Chaillot, J.; Tebbji, F.; Remmal, A.; Boone, C.; Brown, G.W.; Bellaoui, M.; Sellam, A. The monoterpene carvacrol generates endoplasmic reticulum stress in the pathogenic fungus Candida albicans. Antimicrob. Agents Chemother. 2015, 59, 4584–4592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Morcia, C.; Tumino, G.; Ghizzoni, R.; Terzi, V. Carvone (Mentha spicata L.) oils. In Essential Oils in Food Preservation, Flavor and Safety; Elsevier: Amsterdam, The Netherlands, 2016; pp. 309–316. [Google Scholar]
  147. Jung, K.-W.; Chung, M.-S.; Bai, H.-W.; Chung, B.-Y.; Lee, S. Investigation of Antifungal Mechanisms of Thymol in the Human Fungal Pathogen, Cryptococcus Neoformans. Molecules 2021, 26, 3476. [Google Scholar] [CrossRef]
  148. Abbaszadeh, S.; Sharifzadeh, A.; Shokri, H.; Khosravi, A.; Abbaszadeh, A. Antifungal efficacy of thymol, carvacrol, eugenol and menthol as alternative agents to control the growth of food-relevant fungi. J. Mycol. Med. 2014, 24, e51–e56. [Google Scholar] [CrossRef]
  149. OuYang, Q.; Duan, X.; Li, L.; Tao, N. Cinnamaldehyde exerts its antifungal activity by disrupting the cell wall integrity of Geotrichum citri-aurantii. Front. Microbiol. 2019, 10, 55. [Google Scholar] [CrossRef] [Green Version]
  150. Singh, S.; Fatima, Z.; Hameed, S. Citronellal-induced disruption of membrane homeostasis in Candida albicans and attenuation of its virulence attributes. Rev. Soc. Bras. Med. Trop. 2016, 49, 465–472. [Google Scholar] [CrossRef] [Green Version]
  151. Da, X.; Nishiyama, Y.; Tie, D.; Hein, K.Z.; Yamamoto, O.; Morita, E. Antifungal activity and mechanism of action of Ou-gon (Scutellaria root extract) components against pathogenic fungi. Sci. Rep. 2019, 9, 1683. [Google Scholar] [CrossRef]
  152. Li, Z.J.; Liu, M.; Dawuti, G.; Dou, Q.; Ma, Y.; Liu, H.G.; Aibai, S. Antifungal activity of gallic acid in vitro and in vivo. Phytother. Res. 2017, 31, 1039–1045. [Google Scholar] [CrossRef]
  153. Nóbrega, J.R.; Silva, D.D.F.; Andrade Júnior, F.P.D.; Sousa, P.M.S.; Figueiredo, P.T.R.D.; Cordeiro, L.V.; Lima, E.D.O. Antifungal action of α-pinene against Candida spp. isolated from patients with otomycosis and effects of its association with boric acid. Nat. Prod. Res. 2021, 35, 6190–6193. [Google Scholar] [CrossRef]
  154. Venkatesan, R.; Karuppiah, P.S.; Arumugam, G.; Balamuthu, K. β-Asarone exhibits antifungal activity by inhibiting ergosterol biosynthesis in Aspergillus niger ATCC 16888. Proc. Natl. Acad. Sci. India Sect. B Biol. Sci. 2019, 89, 173–184. [Google Scholar] [CrossRef]
  155. David, A.V.A.; Arulmoli, R.; Parasuraman, S. Overviews of biological importance of quercetin: A bioactive flavonoid. Pharmacogn. Rev. 2016, 10, 84. [Google Scholar]
  156. Kwun, M.S.; Lee, D.G. Quercetin-induced yeast apoptosis through mitochondrial dysfunction under the accumulation of magnesium in Candida albicans. Fungal Biol. 2020, 124, 83–90. [Google Scholar] [CrossRef]
  157. Liu, C.; Yang, R.; Jiang, P.; Sun, T.; Zhang, T.; Han, C. Antifungal activity of Osthole on Microsporum canis through interfering with biosynthesis of fungal cell wall. Indian J. Pharm. Sci. 2018, 80, 852–857. [Google Scholar] [CrossRef] [Green Version]
  158. Wu, X.-Z.; Chang, W.-Q.; Cheng, A.-X.; Sun, L.-M.; Lou, H.-X. Plagiochin E, an antifungal active macrocyclic bis (bibenzyl), induced apoptosis in Candida albicans through a metacaspase-dependent apoptotic pathway. Biochim. Biophys. Acta BBA-Gen. Subj. 2010, 1800, 439–447. [Google Scholar] [CrossRef]
  159. Wu, X.-Z.; Cheng, A.-X.; Sun, L.-M.; Lou, H.-X. Effect of plagiochin E, an antifungal macrocyclic bis (bibenzyl), on cell wall chitin synthesis in Candida albicans. Acta Pharmacol. Sin. 2008, 29, 1478–1485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Li, Y.; Ma, Y.; Zhang, L.; Guo, F.; Ren, L.; Yang, R.; Li, Y.; Lou, H. In vivo inhibitory effect on the biofilm formation of Candida albicans by liverwort derived riccardin D. PLoS ONE 2012, 7, e35543. [Google Scholar] [CrossRef]
  161. Hong-Zhuo, S.; Chang, W.-Q.; Zhang, M.; Hong-Xiang, L. Two natural molecules preferentially inhibit azole-resistant Candida albicans with MDR1 hyperactivation. Chin. J. Nat. Med. 2019, 17, 209–217. [Google Scholar]
  162. Yun, D.G.; Lee, D.G. Assessment of silibinin as a potential antifungal agent and investigation of its mechanism of action. IUBMB Life 2017, 69, 631–637. [Google Scholar] [CrossRef] [Green Version]
  163. Rocha da Silva, C.; Sá, L.G.D.A.V.; Dos Santos, E.V.; Ferreira, T.L.; Coutinho, T.D.N.P.; Moreira, L.E.A.; de Sousa Campos, R.; de Andrade, C.R.; Barbosa da Silva, W.M.; de Sá Carneiro, I. Evaluation of the antifungal effect of chlorogenic acid against strains of Candida spp. resistant to fluconazole: Apoptosis induction and in silico analysis of the possible mechanisms of action. J. Med. Microbiol. 2022, 71, 001526. [Google Scholar] [CrossRef]
  164. Veljkovic, E.; Xia, W.; Phillips, B.; Wong, E.T.; Ho, J.; Oviedo, A.; Hoeng, J.; Peitsch, M. Nicotine and Other Tobacco Compounds in Neurodegenerative and Psychiatric Diseases: Overview of Epidemiological Data on Smoking and Preclinical and Clinical Data on Nicotine; Academic Press: Cambridge, MA, USA, 2018. [Google Scholar]
  165. Subramaniam, A.; Shanmugam, M.K.; Ong, T.H.; Li, F.; Perumal, E.; Chen, L.; Vali, S.; Abbasi, T.; Kapoor, S.; Ahn, K.S.; et al. Emodin inhibits growth and induces apoptosis in an orthotopic hepatocellular carcinoma model by blocking activation of STAT3. Br. J. Pharmacol. 2013, 170, 807–821. [Google Scholar] [CrossRef] [Green Version]
  166. Manu, K.A.; Shanmugam, M.K.; Ong, T.H.; Subramaniam, A.; Siveen, K.S.; Perumal, E.; Samy, R.P.; Bist, P.; Lim, L.H.K.; Kumar, A.P.; et al. Emodin Suppresses Migration and Invasion through the Modulation of CXCR4 Expression in an Orthotopic Model of Human Hepatocellular Carcinoma. PLoS ONE 2013, 8, e57015. [Google Scholar] [CrossRef] [Green Version]
  167. Stompor-Gorący, M. The health benefits of Emodin, a natural anthraquinone derived from rhubarb—A summary update. Int. J. Mol. Sci. 2021, 22, 9522. [Google Scholar] [CrossRef]
  168. Janeczko, M. Emodin Reduces the Activity of (1, 3)-D-Glucan Synthase from and Does Not Interact with Caspofungin. Pol. J. Microbiol. 2018, 67, 463–470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Janeczko, M.; Masłyk, M.; Kubiński, K.; Golczyk, H. Emodin, a natural inhibitor of protein kinase CK2, suppresses growth, hyphal development, and biofilm formation of Candida albicans. Yeast 2017, 34, 253–265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Ma, W.; Liu, C.; Li, J.; Hao, M.; Ji, Y.; Zeng, X. The effects of aloe emodin-mediated antimicrobial photodynamic therapy on drug-sensitive and resistant Candida albicans. Photochem. Photobiol. Sci. 2020, 19, 485–494. [Google Scholar] [CrossRef] [PubMed]
  171. Mączka, W.; Duda-Madej, A.; Górny, A.; Grabarczyk, M.; Wińska, K. Can eucalyptol replace antibiotics? Molecules 2021, 26, 4933. [Google Scholar] [CrossRef] [PubMed]
  172. Gupta, P.; Pruthi, V.; Poluri, K.M. Mechanistic insights into Candida biofilm eradication potential of eucalyptol. J. Appl. Microbiol. 2021, 131, 105–123. [Google Scholar] [CrossRef]
  173. Mishra, P.; Gupta, P.; Srivastava, A.K.; Poluri, K.M.; Prasad, R. Eucalyptol/β-cyclodextrin inclusion complex loaded gellan/PVA nanofibers as antifungal drug delivery system. Int. J. Pharm. 2021, 609, 121163. [Google Scholar] [CrossRef]
  174. Bendre, R.S.; Rajput, J.D.; Bagul, S.D.; Karandikar, P. Outlooks on medicinal properties of eugenol and its synthetic derivatives. Nat. Prod. Chem. Res. 2016, 4, 100021. [Google Scholar] [CrossRef]
  175. de Oliveira Pereira, F.; Mendes, J.M.; de Oliveira Lima, E. Investigation on mechanism of antifungal activity of eugenol against Trichophyton rubrum. Med. Mycol. 2013, 51, 507–513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Didehdar, M.; Chegini, Z.; Shariati, A. Eugenol: A novel therapeutic agent for the inhibition of Candida species infection. Front. Pharmacol. 2022, 13, 872127. [Google Scholar] [CrossRef]
  177. Gupta, P.; Gupta, S.; Sharma, M.; Kumar, N.; Pruthi, V.; Poluri, K.M. Effectiveness of phytoactive molecules on transcriptional expression, biofilm matrix, and cell wall components of Candida glabrata and its clinical isolates. ACS Omega 2018, 3, 12201–12214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Alves, J.C.; Ferreira, G.F.; Santos, J.R.; Silva, L.C.; Rodrigues, J.F.; Neto, W.R.; Farah, E.I.; Santos, Á.R.; Mendes, B.S.; Sousa, L.V. Eugenol induces phenotypic alterations and increases the oxidative burst in Cryptococcus. Front. Microbiol. 2017, 8, 2419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Saracino, I.M.; Foschi, C.; Pavoni, M.; Spigarelli, R.; Valerii, M.C.; Spisni, E. Antifungal Activity of Natural Compounds vs. Candida spp.: A Mixture of Cinnamaldehyde and Eugenol Shows Promising In Vitro Results. Antibiotics 2022, 11, 73. [Google Scholar] [CrossRef]
  180. Jafri, H.; Khan, M.S.A.; Ahmad, I. In vitro efficacy of eugenol in inhibiting single and mixed-biofilms of drug-resistant strains of Candida albicans and Streptococcus mutans. Phytomedicine 2019, 54, 206–213. [Google Scholar] [CrossRef] [PubMed]
  181. El-Baz, A.M.; Mosbah, R.A.; Goda, R.M.; Mansour, B.; Sultana, T.; Dahms, T.E.; El-Ganiny, A.M. Back to nature: Combating Candida albicans biofilm, phospholipase and hemolysin using plant essential oils. Antibiotics 2021, 10, 81. [Google Scholar] [CrossRef]
  182. Chen, W.; Viljoen, A.M. Geraniol—A review of a commercially important fragrance material. S. Afr. J. Bot. 2010, 76, 643–651. [Google Scholar] [CrossRef] [Green Version]
  183. Miron, D.; Battisti, F.; Silva, F.K.; Lana, A.D.; Pippi, B.; Casanova, B.; Gnoatto, S.; Fuentefria, A.; Mayorga, P.; Schapoval, E.E. Antifungal activity and mechanism of action of monoterpenes against dermatophytes and yeasts. Rev. Bras. Farmacogn. 2014, 24, 660–667. [Google Scholar] [CrossRef]
  184. Sharma, Y.; Khan, L.; Manzoor, N. Anti-Candida activity of geraniol involves disruption of cell membrane integrity and function. J. Mycol. Med. 2016, 26, 244–254. [Google Scholar] [CrossRef]
  185. Pereira, F.D.O.; Mendes, J.M.; Lima, I.O.; Mota, K.S.D.L.; Oliveira, W.A.D.; Lima, E.D.O. Antifungal activity of geraniol and citronellol, two monoterpenes alcohols, against Trichophyton rubrum involves inhibition of ergosterol biosynthesis. Pharm. Biol. 2015, 53, 228–234. [Google Scholar] [CrossRef] [Green Version]
  186. Leite, M.C.A.; de Brito Bezerra, A.P.; de Sousa, J.P.; de Oliveira Lima, E. Investigating the antifungal activity and mechanism (s) of geraniol against Candida albicans strains. Med. Mycol. 2015, 53, 275–284. [Google Scholar] [CrossRef] [Green Version]
  187. Dalleau, S.; Cateau, E.; Bergès, T.; Berjeaud, J.-M.; Imbert, C. In vitro activity of terpenes against Candida biofilms. Int. J. Antimicrob. Agents 2008, 31, 572–576. [Google Scholar] [CrossRef]
  188. Hwang, J.H.; Jin, Q.; Woo, E.R.; Lee, D.G. Antifungal property of hibicuslide C and its membrane-active mechanism in Candida albicans. Biochimie 2013, 95, 1917–1922. [Google Scholar] [CrossRef]
  189. Hwang, J.H.; Choi, H.; Kim, A.R.; Yun, J.W.; Yu, R.; Woo, E.-R.; Lee, D.G. Hibicuslide C-induced cell death in Candida albicans involves apoptosis mechanism. J. Appl. Microbiol. 2014, 117, 1400–1411. [Google Scholar] [CrossRef]
  190. Xu, T.; Kuang, T.; Du, H.; Li, Q.; Feng, T.; Zhang, Y.; Fan, G. Magnoflorine: A review of its pharmacology, pharmacokinetics and toxicity. Pharmacol. Res. 2020, 152, 104632. [Google Scholar] [CrossRef]
  191. Kim, J.; Ha Quang Bao, T.; Shin, Y.-K.; Kim, K.-Y. Antifungal activity of magnoflorine against Candida strains. World J. Microbiol. Biotechnol. 2018, 34, 167. [Google Scholar] [CrossRef]
  192. Luo, N.; Jin, L.; Yang, C.; Zhu, Y.; Ye, X.; Li, X.; Zhang, B. Antifungal activity and potential mechanism of magnoflorine against Trichophyton rubrum. J. Antibiot. 2021, 74, 206–214. [Google Scholar] [CrossRef] [PubMed]
  193. Fan, Z.; Deng, Z.; Wang, T.; Yang, W.; Wang, K. Synthesis of Natural Tea-Saponin-Based Succinic Acid Sulfonate as Anionic Foaming Agent. J. Surfactants Deterg. 2018, 21, 303–312. [Google Scholar] [CrossRef]
  194. Yan, J.; Wu, Z.; Zhao, Y.; Jiang, C. Separation of tea saponin by two-stage foam fractionation. Sep. Purif. Technol. 2011, 80, 300–305. [Google Scholar] [CrossRef]
  195. Li, Y.; Shan, M.; Li, S.; Wang, Y.; Yang, H.; Chen, Y.; Gu, B.; Zhu, Z. Teasaponin suppresses Candida albicans filamentation by reducing the level of intracellular cAMP. Ann. Transl. Med. 2020, 8, 175. [Google Scholar] [CrossRef]
  196. Yu, Z.; Wu, X.; He, J. Study on the antifungal activity and mechanism of tea saponin from Camellia oleifera cake. Eur. Food Res. Technol. 2022, 248, 783–795. [Google Scholar] [CrossRef]
  197. Weiss, R.A.; Esparza, J. The prevention and eradication of smallpox: A commentary on Sloane (1755)‘An account of inoculation’. Philos. Trans. R. Soc. B Biol. Sci. 2015, 370, 20140378. [Google Scholar] [CrossRef] [Green Version]
  198. Zerbini, F.M.; Kitajima, E.W. From Contagium vivum fluidum to Riboviria: A Tobacco Mosaic Virus-Centric History of Virus Taxonomy. Biomolecules 2022, 12, 1363. [Google Scholar] [CrossRef]
  199. Burrell, C.J.; Howard, C.R.; Murphy, F.A. History and impact of virology. Fenner White’s Med. Virol. 2017, 3–14. [Google Scholar]
  200. Siegel, R.D. Classification of human viruses. Princ. Pract. Pediatr. Infect. Dis. 2018, 1044–1048.e1. [Google Scholar]
  201. Available online: https://covid19.who.int/?mapFilter=deaths (accessed on 28 March 2023).
  202. Meyers, L.; Frawley, T.; Goss, S.; Kang, C. Ebola virus outbreak 2014: Clinical review for emergency physicians. Ann. Emerg. Med. 2015, 65, 101–108. [Google Scholar] [CrossRef]
  203. Sudhan, S.S.; Sharma, P. Human Viruses: Emergence and Evolution. In Emerging and Reemerging Viral Pathogens; Elsevier: Amsterdam, The Netherlands, 2020; pp. 53–68. [Google Scholar]
  204. Lou, H.; Li, H.; Zhang, S.; Lu, H.; Chen, Q. A review on preparation of betulinic acid and its biological activities. Molecules 2021, 26, 5583. [Google Scholar] [CrossRef] [PubMed]
  205. Ríos, J.L.; Máñez, S. New pharmacological opportunities for betulinic acid. Planta Med. 2018, 84, 8–19. [Google Scholar] [CrossRef] [Green Version]
  206. Hong, E.-H.; Song, J.H.; Kang, K.B.; Sung, S.H.; Ko, H.-J.; Yang, H. Anti-influenza activity of betulinic acid from Zizyphus jujuba on influenza A/PR/8 virus. Biomol. Ther. 2015, 23, 345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Loe, M.W.C.; Hao, E.; Chen, M.; Li, C.; Lee, R.C.H.; Zhu, I.X.Y.; Teo, Z.Y.; Chin, W.-X.; Hou, X.; Deng, J. Betulinic acid exhibits antiviral effects against dengue virus infection. Antivir. Res. 2020, 184, 104954. [Google Scholar] [CrossRef]
  208. Bouslama, L.; Kouidhi, B.; Alqurashi, Y.M.; Chaieb, K.; Papetti, A. Virucidal Effect of guggulsterone isolated from Commiphora gileadensis. Planta Med. 2019, 85, 1225–1232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Baliga, M.; Nandhini, J.; Emma, F.; Venkataranganna, M.; Venkatesh, P.; Fayad, R. Indian medicinal plants and spices in the prevention and treatment of ulcerative colitis. In Bioactive Food as Dietary Interventions for Liver and Gastrointestinal Disease: Bioactive Foods in Chronic Disease States; Academic Press: Cambridge, MA, USA, 2012; p. 173. [Google Scholar]
  210. Chen, W.-C.; Wei, C.-K.; Hossen, M.; Hsu, Y.-C.; Lee, J.-C. (E)-Guggulsterone Inhibits Dengue Virus Replication by Upregulating Antiviral Interferon Responses through the Induction of Heme Oxygenase-1 Expression. Viruses 2021, 13, 712. [Google Scholar] [CrossRef] [PubMed]
  211. Ma, L.; Tang, L.; Yi, Q. Salvianolic acids: Potential source of natural drugs for the treatment of fibrosis disease and cancer. Front. Pharmacol. 2019, 10, 97. [Google Scholar] [CrossRef] [Green Version]
  212. Hu, S.; Wang, J.; Zhang, Y.; Bai, H.; Wang, C.; Wang, N.; He, L. Three salvianolic acids inhibit 2019-nCoV spike pseudovirus viropexis by binding to both its RBD and receptor ACE2. J. Med. Virol. 2021, 93, 3143–3151. [Google Scholar] [CrossRef]
  213. Huang, Y.; Yang, C.; Xu, X.-F.; Xu, W.; Liu, S.-W. Structural and functional properties of SARS-CoV-2 spike protein: Potential antivirus drug development for COVID-19. Acta Pharmacol. Sin. 2020, 41, 1141–1149. [Google Scholar] [CrossRef] [PubMed]
  214. Jackson, C.B.; Farzan, M.; Chen, B.; Choe, H. Mechanisms of SARS-CoV-2 entry into cells. Nat. Rev. Mol. Cell Biol. 2022, 23, 3–20. [Google Scholar] [CrossRef]
  215. Yang, C.; Pan, X.; Xu, X.; Cheng, C.; Huang, Y.; Li, L.; Jiang, S.; Xu, W.; Xiao, G.; Liu, S. Salvianolic acid C potently inhibits SARS-CoV-2 infection by blocking the formation of six-helix bundle core of spike protein. Signal Transduct. Target. Ther. 2020, 5, 220. [Google Scholar] [CrossRef] [PubMed]
  216. Biedenkopf, N.; Lange-Grünweller, K.; Schulte, F.W.; Weißer, A.; Müller, C.; Becker, D.; Becker, S.; Hartmann, R.K.; Grünweller, A. The natural compound silvestrol is a potent inhibitor of Ebola virus replication. Antivir. Res. 2017, 137, 76–81. [Google Scholar] [CrossRef]
  217. Müller, C.; Schulte, F.W.; Lange-Grünweller, K.; Obermann, W.; Madhugiri, R.; Pleschka, S.; Ziebuhr, J.; Hartmann, R.K.; Grünweller, A. Broad-spectrum antiviral activity of the eIF4A inhibitor silvestrol against corona-and picornaviruses. Antivir. Res. 2018, 150, 123–129. [Google Scholar] [CrossRef]
  218. Varghese, F.S.; Thaa, B.; Amrun, S.N.; Simarmata, D.; Rausalu, K.; Nyman, T.A.; Merits, A.; McInerney, G.M.; Ng, L.F.; Ahola, T. The antiviral alkaloid berberine reduces chikungunya virus-induced mitogen-activated protein kinase signaling. J. Virol. 2016, 90, 9743–9757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Johari, J.; Kianmehr, A.; Mustafa, M.R.; Abubakar, S.; Zandi, K. Antiviral activity of baicalein and quercetin against the Japanese encephalitis virus. Int. J. Mol. Sci. 2012, 13, 16785–16795. [Google Scholar] [CrossRef]
  220. Singla, R.K.; He, X.; Chopra, H.; Tsagkaris, C.; Shen, L.; Kamal, M.A.; Shen, B. Natural Products for the Prevention and Control of the COVID-19 Pandemic: Sustainable Bioresources. Front. Pharmacol. 2021, 12, 758159. [Google Scholar] [CrossRef]
  221. Hsieh, C.-F.; Jheng, J.-R.; Lin, G.-H.; Chen, Y.-L.; Ho, J.-Y.; Liu, C.-J.; Hsu, K.-Y.; Chen, Y.-S.; Chan, Y.F.; Yu, H.-M. Rosmarinic acid exhibits broad anti-enterovirus A71 activity by inhibiting the interaction between the five-fold axis of capsid VP1 and cognate sulfated receptors. Emerg. Microbes Infect. 2020, 9, 1194–1205. [Google Scholar] [CrossRef] [PubMed]
  222. Choi, H.J.; Song, J.-H.; Lim, C.-H.; Baek, S.-H.; Kwon, D.-H. Anti-human rhinovirus activity of raoulic acid from Raoulia australis. J. Med. Food 2010, 13, 326–328. [Google Scholar] [CrossRef] [PubMed]
  223. Choi, H.-J.; Lim, C.; Song, J.; Baek, S.; Kwon, D.H. Antiviral activity of raoulic acid from Raoulia australis against Picornaviruses. Phytomedicine 2009, 16, 35–39. [Google Scholar] [CrossRef]
  224. Yi, L.; Li, Z.; Yuan, K.; Qu, X.; Chen, J.; Wang, G.; Zhang, H.; Luo, H.; Zhu, L.; Jiang, P. Small molecules blocking the entry of severe acute respiratory syndrome coronavirus into host cells. J. Virol. 2004, 78, 11334–11339. [Google Scholar] [CrossRef] [Green Version]
  225. Cheng, P.W.; Ng, L.T.; Chiang, L.C.; Lin, C.C. Antiviral effects of saikosaponins on human coronavirus 229E in vitro. Clin. Exp. Pharmacol. Physiol. 2006, 33, 612–616. [Google Scholar] [CrossRef]
  226. Zhang, H.-J.; Rumschlag-Booms, E.; Guan, Y.-F.; Wang, D.-Y.; Liu, K.-L.; Li, W.-F.; Nguyen, V.H.; Cuong, N.M.; Soejarto, D.D.; Fong, H.H. Potent inhibitor of drug-resistant HIV-1 strains identified from the medicinal plant Justicia gendarussa. J. Nat. Prod. 2017, 80, 1798–1807. [Google Scholar] [CrossRef] [Green Version]
  227. Gangehei, L.; Ali, M.; Zhang, W.; Chen, Z.; Wakame, K.; Haidari, M. Oligonol a low molecular weight polyphenol of lychee fruit extract inhibits proliferation of influenza virus by blocking reactive oxygen species-dependent ERK phosphorylation. Phytomedicine 2010, 17, 1047–1056. [Google Scholar] [CrossRef]
  228. Haidari, M.; Ali, M.; Casscells, S.W., III; Madjid, M. Pomegranate (Punica granatum) purified polyphenol extract inhibits influenza virus and has a synergistic effect with oseltamivir. Phytomedicine 2009, 16, 1127–1136. [Google Scholar] [CrossRef]
  229. Wu, S.F.; Lin, C.K.; Chuang, Y.S.; Chang, F.R.; Tseng, C.K.; Wu, Y.C.; Lee, J.C. Anti-hepatitis C virus activity of 3-hydroxy caruilignan C from Swietenia macrophylla stems. J. Viral Hepat. 2012, 19, 364–370. [Google Scholar] [CrossRef]
  230. Chen, H.; Lao, Z.; Xu, J.; Li, Z.; Long, H.; Li, D.; Lin, L.; Liu, X.; Yu, L.; Liu, W. Antiviral activity of lycorine against Zika virus in vivo and in vitro. Virology 2020, 546, 88–97. [Google Scholar] [CrossRef]
  231. Hung, P.-Y.; Ho, B.-C.; Lee, S.-Y.; Chang, S.-Y.; Kao, C.-L.; Lee, S.-S.; Lee, C.-N. Houttuynia cordata targets the beginning stage of herpes simplex virus infection. PLoS ONE 2015, 10, e0115475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Andújar, I.; Ríos, J.L.; Giner, R.M.; Recio, M.C. Pharmacological properties of shikonin–a review of literature since 2002. Planta Med. 2013, 79, 1685–1697. [Google Scholar] [CrossRef] [Green Version]
  233. Nahmias, Y.; Goldwasser, J.; Casali, M.; Van Poll, D.; Wakita, T.; Chung, R.T.; Yarmush, M.L. Apolipoprotein B–dependent hepatitis C virus secretion is inhibited by the grapefruit flavonoid naringenin. Hepatology 2008, 47, 1437–1445. [Google Scholar] [CrossRef] [Green Version]
  234. Chiang, L.C.; Ng, L.T.; Cheng, P.W.; Chiang, W.; Lin, C.C. Antiviral activities of extracts and selected pure constituents of Ocimum basilicum. Clin. Exp. Pharmacol. Physiol. 2005, 32, 811–816. [Google Scholar] [CrossRef]
  235. Xiao, T.; Cui, M.; Zheng, C.; Wang, M.; Sun, R.; Gao, D.; Bao, J.; Ren, S.; Yang, B.; Lin, J. Myricetin inhibits SARS-CoV-2 viral replication by targeting Mpro and ameliorates pulmonary inflammation. Front. Pharmacol. 2021, 12, 1012. [Google Scholar] [CrossRef] [PubMed]
  236. Bleasel, M.D.; Peterson, G.M. Emetine, ipecac, ipecac alkaloids and analogues as potential antiviral agents for coronaviruses. Pharmaceuticals 2020, 13, 51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Haid, S.; Novodomská, A.; Gentzsch, J.; Grethe, C.; Geuenich, S.; Bankwitz, D.; Chhatwal, P.; Jannack, B.; Hennebelle, T.; Bailleul, F. A plant-derived flavonoid inhibits entry of all HCV genotypes into human hepatocytes. Gastroenterology 2012, 143, 213–222. [Google Scholar] [CrossRef]
  238. Kuo, Y.-C.; Lin, L.-C.; Tsai, W.-J.; Chou, C.-J.; Kung, S.-H.; Ho, Y.-H. Samarangenin B from Limonium sinense suppresses herpes simplex virus type 1 replication in Vero cells by regulation of viral macromolecular synthesis. Antimicrob. Agents Chemother. 2002, 46, 2854–2864. [Google Scholar] [CrossRef] [Green Version]
  239. Cheng, H.-Y.; Lin, T.-C.; Yang, C.-M.; Wang, K.-C.; Lin, C.-C. Mechanism of action of the suppression of herpes simplex virus type 2 replication by pterocarnin A. Microbes Infect. 2004, 6, 738–744. [Google Scholar] [CrossRef]
  240. Patridge, E.; Gareiss, P.; Kinch, M.S.; Hoyer, D. An analysis of FDA-approved drugs: Natural products and their derivatives. Drug Discov. Today 2016, 21, 204–207. [Google Scholar] [CrossRef] [PubMed]
  241. Dilbato, T.; Begna, F.; Joshi, R.K. Reviews on challenges, opportunities and future prospects of antimicrobial activities of medicinal plants: Alternative solutions to combat antimicrobial resistance. Int. J. Herb. Med. 2019, 7, 10–18. [Google Scholar]
  242. Khameneh, B.; Iranshahy, M.; Soheili, V.; Fazly Bazzaz, B.S. Review on plant antimicrobials: A mechanistic viewpoint. Antimicrob. Resist. Infect. Control 2019, 8, 118. [Google Scholar] [CrossRef] [Green Version]
  243. Khameneh, B.; Iranshahy, M.; Ghandadi, M.; Ghoochi Atashbeyk, D.; Fazly Bazzaz, B.S.; Iranshahi, M. Investigation of the antibacterial activity and efflux pump inhibitory effect of co-loaded piperine and gentamicin nanoliposomes in methicillin-resistant Staphylococcus aureus. Drug Dev. Ind. Pharm. 2015, 41, 989–994. [Google Scholar] [CrossRef]
  244. Dwivedi, G.R.; Maurya, A.; Yadav, D.K.; Singh, V.; Khan, F.; Gupta, M.K.; Singh, M.; Darokar, M.P.; Srivastava, S.K. Synergy of clavine alkaloid ‘chanoclavine’with tetracycline against multi-drug-resistant E. coli. J. Biomol. Struct. Dyn. 2019, 37, 1307–1325. [Google Scholar] [CrossRef] [PubMed]
  245. Mitchell, G.; Lafrance, M.; Boulanger, S.; Séguin, D.L.; Guay, I.; Gattuso, M.; Marsault, E.; Bouarab, K.; Malouin, F. Tomatidine acts in synergy with aminoglycoside antibiotics against multiresistant Staphylococcus aureus and prevents virulence gene expression. J. Antimicrob. Chemother. 2012, 67, 559–568. [Google Scholar] [CrossRef] [Green Version]
  246. Sharifzadeh, A.; Khosravi, A.R.; Shokri, H.; Shirzadi, H. Potential effect of 2-isopropyl-5-methylphenol (thymol) alone and in combination with fluconazole against clinical isolates of Candida albicans, C. glabrata and C. krusei. J. Mycol. Med. 2018, 28, 294–299. [Google Scholar] [CrossRef]
  247. Behbehani, J.; Irshad, M.; Shreaz, S.; Karched, M. Synergistic effects of tea polyphenol epigallocatechin 3-O-gallate and azole drugs against oral Candida isolates. J. Mycol. Med. 2019, 29, 158–167. [Google Scholar] [CrossRef]
  248. Quave, C.L. Antibiotics from nature: Traditional medicine as a source of new solutions for combating antimicrobial resistance. AMR Control 2016, 98–102. [Google Scholar]
  249. Day, J. Botany meets archaeology: People and plants in the past. J. Exp. Bot. 2013, 64, 5805–5816. [Google Scholar] [CrossRef] [Green Version]
  250. Vaou, N.; Stavropoulou, E.; Voidarou, C.; Tsigalou, C.; Bezirtzoglou, E. Towards advances in medicinal plant antimicrobial activity: A review study on challenges and future perspectives. Microorganisms 2021, 9, 2041. [Google Scholar] [CrossRef]
  251. Samuel, S.M.; Kubatka, P.; Büsselberg, D. Treating Cancers Using Nature’s Medicine: Significance and Challenges. Biomolecules 2021, 11, 1698. [Google Scholar] [CrossRef]
  252. Li, D.-D.; Zhang, Y.-H.; Zhang, W.; Zhao, P. Meta-analysis of randomized controlled trials on the efficacy and safety of donepezil, galantamine, rivastigmine, and memantine for the treatment of Alzheimer’s disease. Front. Neurosci. 2019, 13, 472. [Google Scholar] [CrossRef] [PubMed]
  253. Ng, Y.P.; Or, T.C.T.; Ip, N.Y. Plant alkaloids as drug leads for Alzheimer’s disease. Neurochem. Int. 2015, 89, 260–270. [Google Scholar] [CrossRef] [PubMed]
  254. Das, A.M. Clinical utility of nitisinone for the treatment of hereditary tyrosinemia type-1 (HT-1). Appl. Clin. Genet. 2017, 10, 43–48. [Google Scholar] [CrossRef] [Green Version]
  255. Thangapazham, R.L.; Sharad, S.; Maheshwari, R.K. Skin regenerative potentials of curcumin. Biofactors 2013, 39, 141–149. [Google Scholar] [CrossRef]
  256. Duarte, N.B.A.; Takahashi, J.A. Plant spices as a source of antimicrobial synergic molecules to treat bacterial and viral co-infections. Molecules 2022, 27, 8210. [Google Scholar] [CrossRef] [PubMed]
  257. Uzma, F.; Mohan, C.D.; Hashem, A.; Konappa, N.M.; Rangappa, S.; Kamath, P.V.; Singh, B.P.; Mudili, V.; Gupta, V.K.; Siddaiah, C.N.; et al. Endophytic Fungi—Alternative Sources of Cytotoxic Compounds: A Review. Front. Pharmacol. 2018, 9, 309. [Google Scholar] [CrossRef]
  258. Mohan, C.D.; Rangappa, S.; Nayak, S.C.; Jadimurthy, R.; Wang, L.; Sethi, G.; Garg, M.; Rangappa, K.S. Bacteria as a treasure house of secondary metabolites with anticancer potential. Semin. Cancer Biol. 2022, 86, 998–1013. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structure of phytocompounds with antibacterial activity.
Figure 1. Chemical structure of phytocompounds with antibacterial activity.
Life 13 00948 g001
Figure 2. Chemical structure of phytocompounds with good antifungal activity.
Figure 2. Chemical structure of phytocompounds with good antifungal activity.
Life 13 00948 g002
Figure 3. Chemical structure of phytocompounds with antiviral activity.
Figure 3. Chemical structure of phytocompounds with antiviral activity.
Life 13 00948 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jadimurthy, R.; Jagadish, S.; Nayak, S.C.; Kumar, S.; Mohan, C.D.; Rangappa, K.S. Phytochemicals as Invaluable Sources of Potent Antimicrobial Agents to Combat Antibiotic Resistance. Life 2023, 13, 948. https://doi.org/10.3390/life13040948

AMA Style

Jadimurthy R, Jagadish S, Nayak SC, Kumar S, Mohan CD, Rangappa KS. Phytochemicals as Invaluable Sources of Potent Antimicrobial Agents to Combat Antibiotic Resistance. Life. 2023; 13(4):948. https://doi.org/10.3390/life13040948

Chicago/Turabian Style

Jadimurthy, Ragi, Swamy Jagadish, Siddaiah Chandra Nayak, Sumana Kumar, Chakrabhavi Dhananjaya Mohan, and Kanchugarakoppal S. Rangappa. 2023. "Phytochemicals as Invaluable Sources of Potent Antimicrobial Agents to Combat Antibiotic Resistance" Life 13, no. 4: 948. https://doi.org/10.3390/life13040948

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

Article Metrics

Back to TopTop