Next Article in Journal / Special Issue
Effect of Etanercept on Plasmodium yoelii MDR-Induced Liver Lipid Infiltration
Previous Article in Journal / Special Issue
Macrophage Delivered HSV1716 Is Active against Triple Negative Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Preventing Microbial Infections with Natural Phenolic Compounds

1
3B’s Research Group, I3Bs–Research Institute on Biomaterials, Biodegradables and Biomimetics, University of Minho, Headquarters of the European Institute of Excellence on Tissue Engineering and Regenerative Medicine, Avepark, Parque de Ciência e Tecnologia, Zona Industrial da Gandra, 4805-017 Guimarães, Portugal
2
ICVS/3B’s PT Government Associated Laboratory, 4805-017 Guimarães, Portugal
*
Author to whom correspondence should be addressed.
Future Pharmacol. 2022, 2(4), 460-498; https://doi.org/10.3390/futurepharmacol2040030
Submission received: 23 September 2022 / Revised: 22 October 2022 / Accepted: 31 October 2022 / Published: 2 November 2022
(This article belongs to the Special Issue Feature Papers in Future Pharmacology)

Abstract

:
The struggle between humans and pathogens has taken and is continuing to take countless lives every year. As the misusage of conventional antibiotics increases, the complexity associated with the resistance mechanisms of pathogens has been evolving into gradually more clever mechanisms, diminishing the effectiveness of antibiotics. Hence, there is a growing interest in discovering novel and reliable therapeutics able to struggle with the infection, circumvent the resistance and defend the natural microbiome. In this regard, nature-derived phenolic compounds are gaining considerable attention due to their potential safety and therapeutic effect. Phenolic compounds comprise numerous and widely distributed groups with different biological activities attributed mainly to their structure. Investigations have revealed that phenolic compounds from natural sources exhibit potent antimicrobial activity against various clinically relevant pathogens associated with microbial infection and sensitize multi-drug resistance strains to bactericidal or bacteriostatic antibiotics. This review outlines the current knowledge about the antimicrobial activity of phenolic compounds from various natural sources, with a particular focus on the structure-activity relationship and mechanisms of actions of each class of natural phenolic compounds, including simple phenols, phenolic acids, coumarin, flavonoids, tannins, stilbenes, lignans, quinones, and curcuminoids.

Graphical Abstract

1. Introduction

The enthusiasm of antimicrobial discovery has sustained a defeat by the growing resistance of bacterial strains by virtue of the over usage and maladministration of antibiotics for decades [1,2]. Antibiotic resistance stands as an increasing public health concern that causes nearly 50,000 deaths annually across Europe and the US [3]. Additionally, according to United Nations Foundation, in the upcoming years, the scenario may pose dramatic consequences to human health worldwide. Thus, considerable efforts have been devoted to creating novel antimicrobial agents able to combat microbial infections and also overcome the antibiotic resistance reported for nearly every antibiotic used in clinical practice [4]. Besides the fabrication of new antibiotics, which is time-consuming, novel alternative therapeutic strategies are required to turn the tide in this battle as new resistance will arise, and there are no treatments on the horizon for particular infections. Therefore, antibiotic-resistance breakers are urgently needed to fill the void in the development of novel antibiotics [5]. To achieve this, priority antibiotics and bacteria must be clearly identified, along with an extensive survey to recognize and categorize potential resistance breakers for diverse bacterial species and strains. Nevertheless, it is undoubtful that the key feature to this challenge must surely be beyond the simple development of innovative drugs and also include a multidisciplinary culture of change [3].
Antibiotics and/or antimicrobial agents treat infections by affecting the growth or viability of microbial cells. Bactericidal compounds induce bacterial cell death via inhibiting cell wall synthesis, cell membrane function, or protein/enzyme synthesis, whereas bacteriostatic compounds suppress bacterial cellular activity and growth [6]. The concept of bactericidal and bacteriostatic outlines the indications, mechanisms, and contraindications of antimicrobial agents [7]. Nevertheless, over time, microbial species such as bacteria and fungi developed the ability to defeat the mechanisms of conventional antibiotics being designed to kill them. The mechanisms by which microbial pathogens acquire resistance to antibiotics can be summarized as follows: (i) changing the cellular permeability to inhibit the entrance of antibiotics into the microbial cells, (ii) changing the molecular targets of antibiotics so that they are no longer active (iii) enzymatic modification of antibiotics to make them nonfunctioning, and (iv) expression of efflux pumps to pump out antibiotics from the cell [8]. At this point, natural-based compounds have the ability to interact with the microbial cell through multiple antimicrobial mechanisms, making them a top-interest candidate in combating microbial infections and preventing the emergence of drug-resistant strains [9,10]. In the scope of this view, this review will be concentrated on the potential efficiency of natural phenolic compounds to prevent microbial infections owing to their remarkable features.
This review provides an in-depth survey of the up-to-date knowledge of a broad assortment of natural phenolic compounds as possible alternatives for antibiotics. For each class, the antimicrobial mechanism and structure-activity relationship of potential antimicrobial agent candidates from natural sources will be highlighted with a particular focus. In addition, a detailed list and description of the prominent and studied natural phenolic compounds with potential against clinically relevant pathogens will be specified, which may serve different research dedicated to discovering and resupplying these natural compounds with active antimicrobial properties. Lastly, we will also address the main hurdles, future prospects, and issues to overcome.

2. Natural Phenolic Compounds against Microbes

Phenolic compounds are molecules with at least one phenol unit that can be obtained from bacteria, fungi, and marine organisms, but mostly from plants [11,12]. Based on their chemical structure, phenolic compounds are subdivided into diverse subcategories, including simple phenols, phenolic acids, coumarins, flavonoids, tannins, stilbenes, lignans, quinones, and curcuminoids [13,14]. Natural phenolic compounds exhibit broad-range biological activities, including antibacterial, antifungal, anti-inflammatory, antiviral, hepatoprotective, antithrombotic, anticarcinogenic, antiallergic, and antioxidant actions [12,15,16,17,18,19,20,21,22]. Therefore, phenolics are considered potential therapeutic agents against diabetes, cancer, cardiovascular dysfunctions, neurodegenerative diseases, inflammatory diseases, and anti-aging [23,24]. A summarized catalogue of phenolics and polyphenolics compounds with antimicrobial properties against a wide panel of microorganisms is disclosed in Table 1.

2.1. Simple Phenols and Phenolic Acids

Simple phenols are described as compounds presenting an aromatic ring with one or more hydroxyl groups attached [25]. Representative examples of simple phenols include catechol, hydroquinone, resorcinol, and phloroglucinol, as illustrated in Figure 1 [26]. Although these compounds are hardly found alone in plants, they usually appear joint with either cinnamic acids or benzoic acid [25,27]. The main benefit of phenolic acids is their metabolizing capability by natural microbes [27]. Phenolic acids encompass a carboxylic acid linked to an aromatic compound, a phenol [28]. According to their structure, phenolic acids comprise two key classes, including benzoic acid and cinnamic acid derivatives [29]. Hydroxycinnamic acids are common phenolic acids within plant species and are generally derived from cinnamic acid [30]. These natural compounds can be found as their esters, glycosides, and/or conjugated with proteins [31,32]. The utmost common hydroxycinnamic acids derivatives are caffeic acid, ferulic acid, and coumaric acid, as shown in their chemical structure in Figure 1 [25]. Hydroxybenzoic acid (HBA) derivatives are phenolic compounds with a basic structure of C6-C1. Although HBAs can be identified as free acids, they mainly occur in conjugated form, generally as esters [33]. The basic skeleton structure of benzoic acid and the chemical structure of some representative examples of benzoic acid derivatives isolated from natural sources are shown in Figure 1.
As previously mentioned, even though phenolic acids and their derivatives are widely isolated from the plant kingdom, recent investigations indicated their presence in other sources, including marine-derived microorganisms, plant-derived endophytic fungus, marine organisms, and bacterium species [34,35,36,37]. Several studies have been conducted to emphasize the importance of phenolic compounds in the ability to develop resistance in multidrug-resistant bacteria. As a representative example, two phenolic acids (vanillic acid as a hydroxybenzoic acid derivative and 2-Hydroxycinnamic acid as a hydroxycinnamic acid derivative) and an antibiotic (Vancomycin) were exposed continuously to Methicillin-resistant Staphylococcus aureus (MRSA) and methicillin-susceptible Staphylococcus aureus (MSSA) bacteria. The resistance ability was assessed by ascertaining the MIC values of the tested compounds before, during, and after exposure of MRSA and MSSA bacteria to sub-inhibitory concentrations of these compounds. These data demonstrated that MRSA and MSSA did not acquire resistance to both vanillic acid and 2-hydroxycinnamic acid; by contrast, vancomycin caused the acquisition of resistance of both strains [38]. In another study, benzoic acid, purified from the endophytic fungus strain Neurospora crassa, displayed prominent antimicrobial activity against six different multidrug-resistant (MDR) clinical pathogens (see Table 1) and also confirmed its nontoxicity [39]. These findings indicate the promising aspect of phenolic acids in preventing the emergence of new resistant bacterial strains and combating with the MDR pathogens. On the other hand, a clinical study on Staphylococcus aureus (S. aureus) strains documented that caffeic acid exhibited a promising antibacterial effect with the MICs ranging between 256 μg/mL and 1024 μg/mL against reference strains and clinical isolates of MRSA and MSSA from infected wounds. In addition, the study verified that the combination of caffeic acid with the antibiotics (erythromycin, clindamycin, and cefoxitin) tested caused synergistic activity by sensitizing the bactericidal and bacteriostatic action of antibiotics. It should be noted that the combination of caffeic acid with vancomycin did not show a prominent difference [40].
Research towards comprehending antibacterial mechanisms functioning at the molecular level with the purpose of exploiting these bioactive compounds in clinical settings has also advanced. To date, many studies have documented that phenolic acids and their derivatives show antimicrobial effects through bactericidal actions. This can be explained mainly by the fact that phenolic acids are weak organic acids whose lipophilicity differs from each other. The dissociation constant and lipophilicity affect the solubility of the compounds in microbial membranes and, thus, their antimicrobial activity [41]. Usually, the undissociated forms of the phenolic acids cross the cell membrane through passive diffusion, acidifying the cytoplasm by disrupting the cell membrane, causing the outflow of essential intracellular constituents and resulting in microbial cell death [41,42,43]. Moreover, Lou et al. [44] also assessed the antibacterial activity of p-coumaric acid, where it was discovered that p-coumaric acid triggered the death of bacterial cells by two main mechanisms, disruption of the microbial cell membrane and/or binding to the microbial genomic DNA. Another study considering the relationship between pH and bacterial growth of phenolic acids, including chlorogenic acid and the hydroxycinnamic acids, caffeic acid, p-coumaric acid, and ferulic acid, reported that all tested hydroxycinnamic acids had a bactericidal effect at pH 4.5 and bacteriostatic effect at higher pH against Listeria Monocytogenes (L. monocytogenes) [45]. However, further studies on their action mechanisms are needed to prove the existence of possible bacteriostatic effects.
The antimicrobial actions of phenolic acids depend on the chain length, and number and position of substituents on the core benzene ring [46]. The antibacterial behavior of caffeic acid alkyl esters tends to increase with the increase in alkyl chain length. However, in the presence of a long alkyl chain, their antimicrobial activity diminishes due to the possible steric hindrance [47,48]. As mentioned earlier, due to their partly lipophilic character, the ability of these compounds to acidify the cytoplasm by passing through the cell membrane by passive diffusion depends on the number of hydroxyls (–OH), methoxy (–OCH3), carboxyl (−COOH), functional groups and saturation of alkyl side chain [27,43].
Table 1. Minimum inhibitory concentration (µg/mL) of phenolic and polyphenolic compounds against different pathogenic microorganisms.
Table 1. Minimum inhibitory concentration (µg/mL) of phenolic and polyphenolic compounds against different pathogenic microorganisms.
Secondary Metabolite ClassSubclassesCompoundSourceMicroorganismPositive ControlRef.
Gram-Positive BacteriaGram-Negative BacteriaFungiGram-Positive BacteriaGram- Negative BacteriaFungi
Phenolic acidsBenzoic acid derivativesBenzoic acidNeurospora crassa (Microorganism)SAa (587)ECa (274), PA a (302)CAa (347),
AN a (570)
Streptomycin
SA a (44),
EC a (210), PA a (210)
Ketoconazole
CA a (200),
AN a (200)
[39]
Gallic acidCaesalpinia mimosoides Lamk
(Plant)
SA (1250) Streptomycin
SA (0.16)
[49]
Diospyros virginiana L.
(Plant)
LM (40),
SA (10), BC c (25)
EC (40),
ST (10),
PA (25)
AN (30),
AV (10),
AF (25)
Streptomycin
LM (150),
SA (250),
BC c (50),
Streptomycin
EC (100),
ST (50),
PA (50)
Ketoconazole
AN (200),
AV (200),
AF (200)
[50]
Mezoneuron benthamianum
(Plant)
SAc (100)ECc (25), PA c (100) -- [51]
4- Hydroxybenzoic acidGanoderma lucidum
(Plant)
LM (30),
SA (3),
BC (3)
EC (30),
ST (3),
PA (3)
AN (30),
AV (3),
AF c (120)
Streptomycin
LM (170),
SA (40),
BC (90)
Streptomycin
EC (170),
ST (170),
PA (170)
Ketoconazole
AN (200),
AV (200),
AF c (200)
[52]
4-(2′R, 4′-dihydroxybutoxy) benzoic acidPenicillium sp. of Nerium indicum
(Microorganism)
EC (125), PA (125) Streptomycin sulfate
EC (7.81),
PA (7.81)
[53]
Vanillic acidStenoloma chusanum (Plant) CA (50),
AN (100),
TR (50)
-[54]
Cinnamic acid derivativesCinnamic acidGanoderma lucidum (Plant)LM (7),
SA (1.5),
BC (1.5)
EC (7),
ST (1.5),
PA (0.7)
AN (30),
AV (7),
AF (7)
Streptomycin
LM (170),
SA (40),
BC (90)
Streptomycin
EC (170),
ST (170),
PA (170)
Ketoconazole
AN (200),
AV (200),
AF (200)
[52]
Caffeic acidNauclea latifolia leaf
(Plant)
SA (5000)EC (625), PA (2500) Streptomycin
SA (125)
Streptomycin
EC (125),
PA (500)
[55]
p-Coumaric acidStereospermum zenkeri
(Plant)
SAa (37.50) Ampicillin
SA a (0.80)
[56]
trans-o-coumaric acidDistichochlamys benenica
(Plant)
SA (249.5)EC (1001.4), PA (1001.4) Ciprofloxacin
SA (0.215)
Ciprofloxacin
EC (0.013),
PA (0.013)
[57]
CoumarinsSimple coumarinsUmbelliferoneLoeselia Mexicana (Plant) CA (50),
TR (25)
Nystatin
CA (8),
TR (-)
Miconazole
CA (-), TR (4)
[58]
Ferulago Species
(Plant)
SA (250)EC (500), PA (250)CA (125)Streptomycin
SA (6.25)
Streptomycin
EC (25),
PA (25)
Ketoconazole
CA (25)
Miconazole
CA (3)
[59]
OstholMagydaris tomentosa
(Plant)
SA (64),
SE (32)
EC (256), PA (128) Cefotaxime
SA (2),
SE (0.1)
Cefotaxime
EC (0.1),
PA (1.6)
[60]
Prangos hulusii
(Plant)
SA (125), MRSA c (16) Cefotaxime
SA (2), MRSA c (16)
[61]
Prangos pabularia
(Plant)
MRSA (31.25)PA (31.25) -- [62]
Ferulago Species
(Plant)
SA (500)EC (500), PA (250)CA (500)Streptomycin
SA (6.25)
Streptomycin
EC (25), PA (25)
Ketoconazole
CA (25)
Miconazole
CA (3)
[59]
NovobiocinNocardiopsis gilva (Microorganism)SA (64) Kanamycin
SA (4)
[63]
Streptomyces strain (Microorganism)MRSA (0.25) - [64]
UlopterolToddalia asiatica (L.) Lam.
(Plant)
SA (125), MRSA c (250),
SE (15.6)
ECc (62.5–250),
ST (125),
SF (62.5),
PA (125)
CA (250),
AF (15.6),
TR (250)
Streptomycin
SA (6.25), MRSA c (6.25), SE (25)
Streptomycin
EC c (25),
ST (30),
SF (6.25),
PA (25)
Ketoconazole
CA (25),
AF (<12.5),
TR (<12.5)
[65]
Ferulago Species
(Plant)
SA (500)EC (500), PA (500)CA (250)Streptomycin
SA (6.25)
Streptomycin
EC (25),
PA (25)
Ketoconazole
CA (25)
Miconazole
CA (3)
[59]
FuranocoumarinsPeucedaninPeucedanum luxurians
(Plant)
SA (1500), SE (1750)EC (2750), PA (1400) Netilmicin
SA (4), SE (4)
Netilmicin
EC (10),
PA (88)
[66]
Oxypeucedanin hydrateAngelica pancicii Vandas (Apiaceae) (Plant)LM (1000), SA (1000)EC (1000), ST (1000), PA (1000) Streptomycin
LM (170),
SA (40)
Streptomycin
EC (170),
ST (170),
PA (170)
[67]
Angelica lucida
(Plant)
SA (650),
SE (600)
EC (650), PA (810) Netilmicin
SA (4), SE (4)
Netilmicin
EC (10),
PA (3)
[68]
(R)-(+) oxypeucedanin hydrateFicus exasperata
(Plant)
MRSAc (78.12), BC c (9.76)ECb,c (39.06), PA b,c (156.25)CAc (39.06)Gentamicin
MRSA c (4.88), BC c (4.88)
Gentamicin
EC b,c (4.88), PA b,c (9.76)
Nystatin
CA c (19.53)
[69]
ImperatorinHeracleum mantegazzianum Sommier and Levier (Apiaceae)
(Plant)
SA (250–1000),
BC (500),
SE (1000)
EC (1000), ST (1000), PA (1000)CA (250)---[70]
Magydaris tomentosa
(Plant)
SA (32),
SE (32)
EC (32),
PA (64)
Cefotaxime
SA (2),
SE (0.1)
Cefotaxime
EC (0.1),
PA (1.6)
[60]
Angelica lucida
(Plant)
SA (45),
SE (35)
EC (25),
PA (70)
Netilmicin
SA (4), SE (4)
Netilmicin
EC (10),
PA (3)
[68]
Prangos pabularia
(Plant)
MRSA (62.5)PA (65.5) -- [62]
5-methoxy-3-(3-methyl-2,3-
Dihydroxybutyl) psoralen
Dorstenia turbinata (Plant)MRSAc (39.06)EC b,c (78.12),
PA b,c (39.06)
CA c (19.53), CG c (39.06), TR c (9.76)Gentamycin
MRSA c (9.76)
Gentamycin
EC b,c (4.88), PA b,c (9.76)
Nystatin
CA c (19.53)
[71]
Pyrano coumarinsAgasyllinFerulago campestris (Plant)SAa,c (64)PA a,c (125)-Cefotaxime
SA a,c (resistant)
Cefotaxime
PA a,c (32)
[72]
Zosima absinthifolia
(Plant)
SA (5000)EC (5000) Gentamycin
SA (8)
Gentamycin
EC (8)
[73]
Bi-coumarin (Dicoumarin)DaphnoretinLoeselia mexicana
(Plant)
CA (50),
TR (25),
AN (100)
Nystatin
CA (8), TR (-)
AN (-)
Miconazole
CA (-), TR (4)
AN (8)
[58]
FlavonoidsFlavonolsMyricetinDiospyros virginiana L.
(Plant)
LM (10),
SA (5),
BC c (2.5)
EC (15),
ST (15),
PA (150)
AN (5),
AV (2.5),
AF (2.5)
Streptomycin
LM (150),
SA (250),
BC c (50)
Streptomycin
EC (100),
ST (50),
PA (50)
Ketoconazole
AN (200),
AV (200),
AF (200)
[50]
QuercetinDiospyros virginiana L. (Plant)LM (10),
SA (1),
BC c (2.5)
EC (15),
ST (15),
PA (200)
AN (5),
AV (2.5),
AF (2.5)
Streptomycin
LM (150),
SA (250),
BC c (50)
Streptomycin
EC (100),
ST (50),
PA (50)
Ketoconazole
AN (200),
AV (200),
AF (200)
[50]
Nauclea latifolia
(Plant)
SA (156)EC (2500), PA (1250) Streptomycin
SA (125)
Streptomycin
EC (125),
PA (500)
[55]
Euphorbia schimperiana
(Plant)
LM (450), SA (420), BC (430)EC (430), PA (420) -- [74]
Macaranga conglomerate
(Plant)
SA (500)EC (500), PA (500) Ciprofloxacin
SA (15.6)
Ciprofloxacin
EC (1.0),
PA (15.6)
[75]
Monanthotaxis littoralis
(Plant)
SA (16)EC (16),
PA (16)
CA (16),
CN (8)
Vancomycin
SA (0.5)
Vancomycin
EC (32),
PA (16)
Fluconazole
CA (1.0),
CN (2.0)
[76]
Flavones6,7,4′-trimethyl flavoneWulfenia amherstiana (Plant)SA (127.06–128.94)PA (510.98–513.02)CA (127.37–128.63),
CG (255.18–256.82),
FS (511.02–512.98)
---[77]
LuteolinDiospyros virginiana L.
(Plant)
LM (1.5), SA (1.5),
BC c (2.5)
EC (15),
ST (20),
PA (200)
AN (10),
AV (5),
AF (2.5)
Streptomycin
LM (150),
SA (250),
BC c (50)
Streptomycin
EC (100),
ST (50),
PA (50)
Ketoconazole
AN (200),
AV (200),
AF (200)
[50]
Flavanols
(Flavan-3-ols)
(+)-Catechin-3′-O-rhamnopyranosideNeocarya macrophylla (Sabine) Prance (Chrysobalanaceae)
(Plant)
SA c (25)PA c (25), EC c (25)CA c (6.25)---[78]
(−)-CatechinPrunus avium L.
(Plant)
LM (100), SA a (100), BC a (100)EC (100) -- [79]
IsoflavonesMyrsininone AFicus auriculata (Plant)BC (2.03), SE (0.51)EC (2.03), PA (4.06) Streptomycin sulfate
BC (0.23),
SE (0.23)
Streptomycin sulfate
EC (0.45),
PA (0.45)
[80]
DaidzeinSpatholobus parviflorus
(Plant)
BC (64)PA (128) Vancomycin
BC (0.25)
Gentamycin
PA (1.0)
[81]
LupalbigeninMaclura cochinchinensis (Lour.) Corner (Plant)SA (1), MRSA (1) CA (4)Vancomycin
SA (0.5), MRSA (1.0)
Ampicillin
CA (0.25)
[82]
FlavanonesLupinifolinDerris reticulata Craib
(Plant)
SA (12.5), BC (12.5), SE (25) Penicillin G
SA (0.05),
BC (ND),
SE (0.05)
[83]
7-O-(2,2-dimethylallyl)-aromadendrinMaclura cochinchinensis (Lour.) Corner (Plant)SA (32), MRSA (32) CA (64)Vancomycin
SA (0.5), MRSA (1.0)
Ampicillin
CA (0.25)
[82]
TanninsGallotanninsPenta-O-galloylglucoseRhus trichocarpa Miquel
(Plant)
SA (64–128),
MRSA (64–128),
BC (32), SE (32)
CA (64)Vancomycin
SA (0.25–1), MRSA (0.25–1),
BC (>64), SE (1)
Vancomycin
CA (32)
[84]
EllagitanninsPunicalaginPunica granatum L. (Plant)SA (0.6),
SE (0.6)
EC (1.2),
PA (0.6)
CA (1.2)---[85]
3,3′-di-O-methylellagic acidEuphorbia schimperiana
(Plant)
LM (450), SA (450), BC (450)EC (450), PA (430) -- [74]
Isorugosins BLiquidambar formosana
(Plant)
MRSA (32.46–63.96) Oxacillin
(128.05–256.1)
[86]
VescalaginCork
(Plant)
SA (500), MRSA (125)PA (1000) - -[87]
Castalagin- -
Condensed tanninsA type-proanthocyanidinQuercus ilex
(Plant)
LM (100.72), SA (100.72), BC c (100.72)EC (100.72), ST (100.72), PA (100.72)AN (100.72), AF (100.72), AV (100.72)Streptomycin
LM (150.04), SA (100.03), BC c (25.01)
Streptomycin
EC (100.03), ST (100.03), PA (100.03)
Ketoconazole
AN (201.94), AF (201.94), AV (201.94)
[88]
PhlorotanninsFucofuroeckol-AEisenia bicyclis (Marine algae) CA b,c (512) Fluconazole
CA b,c (512–8197)
[89]
DieckolEcklonia stolonifera (Marine algae)MRSAa,c (64–128)EC (256),
ST (256),
SF (256)
Ampicillin
MRSA a,c (128–512)
Vancomycin
EC (512),
ST (512),
SF (256)
[90]
StilbenesStilbene MonomersResveratrolMezoneuron benthamianum
(Plant)
SAc (25)ECc (25), PA c (25),
PA (200)
-- [51]
Nauclea pobeguiinii (Plant) ECb,c (32–128),
PA a,c (256)
Chloramphenicol
EC b,c (64),
PA a,c (256)
[91]
Gnetum gnemon L. (Plant) EC (>3000)SC (2000) --[92]
Bacillus sp. N strain (Microorganism)SA (32)EC (32),
PA (64)
CA (64)Ciprofloxacin
SA (5)
Cefotaxime
SA (250)
Ciprofloxacin
EC (5),
PA (10)
Cefotaxime
EC (100),
PA (500)
Amphotericin B
CA (50)
[93]
PiceatannolMezoneuron benthamianum
(Plant)
SAc (25)EC c (25), PA c (300), -- [51]
Spirotropis longifolia
(Plant)
CAc (2),
CG c (4),
TR c (8)
Fluconazole
CA c (>64),
CG c (8),
TR c (2)
[94]
PterostilbeneCommercial ProductLM (64),
SA (4),
BC (16)
EC (512), PA (512) Chlorhexidine
LM (8),
SA (32),
BC (8)
Chlorhexidine
EC (32),
PA (32)
[95]
OxyresveratrolSpirotropis longifolia
(Plant)
CAc (>64),
CG c (8),
TR c (16)
Fluconazole
CA c (>64),
CG c (8),
TR c (2)
[94]
Morus alba L.
(Plant)
TR (500)Miconazole nitrat
TR (1)
[96]
Chiricanine AArachis hypogaea
(Plant)
MRSA (12.5) - [97]
3,5-Dihydroxy-4-isopropylstilbeneBacillus sp. N strain (Microorganism)SA (8)EC (>1000), PA (>1000)CA (24)Ciprofloxacin
SA (5)
Cefotaxime
SA (250)
Ciprofloxacin
EC (5),
PA (10)
Cefotaxime
EC (100),
PA (500)
Amphotericin B
CA (50)
[93]
Photorhabdus luminescens
(Microorganism)
CN (12), AF (12) -[98]
Stilbene DimersMonalittorinMonanthotaxis littoralis
(Plant)
SA (64)EC (65),
PA (64)
CA (16),
CN (16),
Vancomycin
SA (0.5)
Vancomycin
EC (32),
PA (16)
Fluconazole
CA (1.0),
CN (2.0)
[76]
Gnetin DSpirotropis longifolia
(Plant)
CAc (64),
CG c (32),
TR c (8)
Fluconazole
CA c (>64),
CG c (8),
TR c (2)
[94]
Gnetin CGnetum gnemon L. (Plant) EC (1000)SC (500) --[92]
Longistylin ACajanus cajan (Plant)SA (1.56),
BC (25), MRSA (1.56)
EC (>100) Vancomycin
SA (1.56),
BC (50), MRSA (0.78)
Vancomycin
EC (50)
[99]
MonalittorinMonanthotaxis littoralis
(Plant)
SA (64)EC (64),
PA (64)
CA (16),
CN (16)
Vancomycin
SA (0.5)
Vancomycin
EC (32),
PA (16)
Fluconazole
CA (1.0),
CN (2.0)
[76]
Stilbene OligomersRockiol A and Rockiol BPaeonia rockii
(Plant)
SA (25)EC (200), PA (200) Penicillin G
SA (10)
Penicillin G
EC (20),
PA (10)
[100]
Upunaphenol DDryobalanops lanceolata
(Plant)
SA (45.3), SE (22.7)EC (>906.9), ST (>906.9), SF (453.4) Chloramphenicol
SA (0.008), SE (0.008)
Chloramphenicol
EC (323.132), ST (323.132), SF (0.010)
[101]
Heyneanol AVitis thunbergii var. taiwaniana
(Plant)
SA (2), MRSA (2) Vancomycin
SA (1),
MRSA (1)
Oxacillin
SA (2),
MRSA (64–128)
[102]
LignansTetrahydrofuran LignansMatairesinolCentaurea scabiosa
(Plant)
SA (10), MRSA (1000), SE (10)EC (10),
PA (10)
Ciprofloxacin
SA (2.5 × 10−4), MRSA (2.5 × 10−4),
SE (2.5 × 10−5)
Ciprofloxacin
EC (2.5 × 10−4), PA (0.0025)
[103]
Centaurea raphanina ssp. Mixta
(Plant)
AN (100),
AV (100)
Miconazole
AN (1.5),
AV (2)
[104]
LariciresinolRubia philippinensis
(Plant)
SA (125)EC (250) -- [105]
Sambucus williamsii
(Plant)
CA (25) Amphotericin B
CA (6.25)
[106]
Iso-hydroxymatairesinolPunica granatum L.
(Plant)
SA (1500), SE (190)EC (560), PA (1500) -- [107]
Punicatannin CSA (1500), SE (750)EC (1120) --
Furofuran LignansSesaminZanthoxylum paracanthum Kokwaro
(Plant)
SA (500) Omacilin
(0.49)
[108]
Phillyrigeninside BForsythia suspensa
(Plant)
SA (10)EC (20)CA (20)Gentamicin
SA (4)
Gentamicin
EC (4)
Gentamicin
CA (4)
[109]
PinoresinolCinnamomum Camphora
(Plant)
SA (15.60)EC (31.25),
PA (7.80)
-- [110]
Sambucus williamsii
(Plant)
CA (12.5) Amphotericin B
CA (6.5)
[111]
Arylnaphthalene Lignan2,3-dimethyl-4-(4′-hydroxy-3′,5′-dimethoxyphenyl)-6-hydroxy-7-methoxy-naphthaleneGanoderma lipsiense
(Microorganism)
SA (1.25), SE (>10)EC (10)CA (>10)Ciprofloxacin
SA (0.156),
SE (0.156)
Ciprofloxacin
EC (0.156)
Ciprofloxacin
CA (0.156)
[112]
Arylnaphthalenelactone LignanJusticidin BNocardia sp.
(Microorganism)
SA (1),
BC (2.5)
EC (0.5),
PA (0.2)
CA (4.5),
CN (0.5),
AN (0.2)
---[113]
Dibenzocyclooctadiene LignanManglisin BManglietiastrum sinicum
(Plant)
SA (0.025), MRSA (0.025) Vancomycin hydrochloride
SA (1.63 × 10−3),
MRSA (8.02 × 10−4)
[114]
QuinonesBenzoquinonesOncocalyxone AAuxemma oncocalyx (Allem) Taub
(Plant)
LM (37.75), SA (18.87), MRSA (18.87–37.75),
SE (9.43–37.75)
EC (>151), PA (>151)CA (>151),
CN (>151),
AF (>151)
Vancomycin
LM (<2.0),
SA (1.0), MRSA (1.0), SE (2.0)
Meropenem
EC (<0.1), PA (<0.39)
Itraconazole
CA (0.25),
CN (0.06),
AF (0.125)
[115]
2-methyl-6-(-3-methyl-2-butenyl)benzo-1,4-quinoneGunnera perpensa
(Plant)
SA (39), BC (18),
SE (9.8)
EC (>6250)CA (130), CN (70)Ciproflaxin
SA (0.31),
BC (2.5),
SE (1.25)
Ciproflaxin
EC (0.63)
Amphotericin B
CA (1.25),
CN (2.5)
[116]
3,5- dimethoxy-2- methylthio)cyclohexa-2,5 diene-1,4-dioneDiplocentrus melici
(Animal)
SA (4) Ampicillin
SA (0.5)
[117]
2,6-Dimethoxy-1,4-BenzoquinoneWood tar
(Plant)
SA (32)EC (64),
ST (32)
Chloramphenicol
SA (32)
Chloramphenicol
EC (32),
ST (32)
[118]
NaphthoquinonesPlumbaginDiospyros bipindensis
(Plant)
SA (20) Ampicillin
SA (0.7)
[119]
Plumbago zeylanica L.
(Plant)
MRSA (4–8) [120]
Diospyros crassiflora (Plant) CAc (0.78),
CG c (3.12),
CN c (1.56), AN c (0.78)
Ketoconazole
CA c (0.25),
CG c (5),
CN c (0.25), AN c (0.25)
[121]
Plumbago zeylanica (Plant)SA (0.5)EC (8),
PA (8)
CA (2)Ciprofloxacin
SA (1.0)
Amoxicillin
SA (0.5)
Ciprofloxacin
EC (0.5),
PA (0.5)
Amoxicillin
EC (4),
PA (128)
Ketoconazole
CA (256)
[122]
Plumbago indica (Plant)SA (3.12), SE (0.018) Tetracycline HCl
SA (0.38),
SE (0.048)
[123]
2-methyl-1,4-naphthoquinone (vitamin K3)Pulsatilla koreana (Plant)SA (2.6–4)PA (4)CA (32–96), CG (8)Tetracycline
SA (0.5)
Tetracycline
PA (0.22–0.38)
Ketoconazole
CA (10.6–16), CG (8–13.4)
[124]
2-Methoxy-1,4-naphthoquinoneImpatiens balsamina L.
(Plant)
SA (16),
BC (64)
CAc (0.62–2.50),
CA a,c (0.62–1.25),
AF c (0.31)
Chloramphenicol
SA (8),
BC (8)
Amphotericin B
CA c (1.1),
CA a,c (90), AF c (1.1)
[125]
BluemomycinStreptomyces sp.
(Microorganism)
SA (NA), MRSA c (10.6–39.4), SE (35.6–64.4)ECa,c (8.9–39.4),
ST (8.9–16.1),
SF (5.3–19.7),
PA (5.3–19.7)
CAc (46.4–53.6), TR (NA)Streptomycin
SA (2.65–9.85), MRSA c (6.25–20.65), SE (17.8–32.2)
Streptomycin
EC a,c (10.6–39.4),
ST (17.8–32.2), SF (2.65–9.85), PA (10.6–39.4)
Ketoconazole
CA c (10.6–39.4),
TR (<26.9)
[126]
5-hydroxy-3,6-dimethoxy-7-methyl-1,4-naphthalenedioneXanthium sibiricum
(Plant)
SA (2.78), BC (22.2)EC (5.55) Ciprofloxacin
SA (1.39), BC (5.55)
Ciprofloxacin
EC (0.69)
[127]
AnthraquinonesZenkequinone AStereospermum zenkeri
(Plant)
ECa (37.50), PA a (18.75) Ampicillin
EC a (0.40),
PA a (0.80)
[56]
EmodinRumex abyssinicus (Plant)SA (8), MRSA (32)SF (8),
PA (16)
CA (8), CN (8)Ciprofloxacin
SA (0.5), MRSA (4)
Ciprofloxacin
SF (8),
PA (0.5)
Fluconazole
CA (1), CN (2)
[128]
Cassia occidentalis (Plant)SA (3.9)EC (>50) Neomycin
SA (6.3)
Neomycin
EC (1.6)
[129]
PhyscionRumex abyssinicus (Plant)SA (8), MRSA (16)SF (8),
PA (8)
CA (8), CN (8) Ciprofloxacin
SA (0.5), MRSA (4)
Ciprofloxacin
SF (8),
PA (0.5)
Fluconazole
CA (1), CN (2)
[128]
Isoversicolorin CAspergillus nidulans
(Microorganism)
EC (32) Chloramphenicol
EC (1)
[130]
2,3-dihydroxy-9,10-anthraquinoneStreptomyces galbus (Microorganism)SA (>100),
MRSA c (12.5),
SE (>100)
ECc (50),
ST (12.5),
SF (25),
PA (12.5)
CA (50)Streptomycin
SA (6.25), MRSA c (6.25), SE (12.5)
Streptomycin
EC c (25),
ST (6.25),
SF (6.25),
PA (25)
Ketoconazole
CA (25)
[131]
5-Hydroxy ericamycinActinoplanes sp.
(Microorganism)
SAb,c (<0.06) MRSA (0.016)
MRSA c (<0.06), SE b,c (<0.06)
EC (4),
EC b,c (16),
PA (16)
Vancomycin
SA b,c (1.0–8.0), MRSA (2.0),
MRSA c (1.0), SEb,c (2.0)
Vancomycin
EC (>64),
EC b,c (>64),
PA (ND)
[132]
CurcuminoidsCurcuminZingiber spectabile (Plant)SA (500), BC (125)EC (NA) Tetracycline
SA (3.91),
BC (1.95)
Tetracycline
EC (NA)
[133]
Curcuma longa LinnéMRSA (125–250), MRSA c (125–250) Oxacillin
MRSA (500–1000, >1000), MRSA c (500–1000)
Ciprofloxacin
MRSA (7.8–250.0),
MRSA c (1.95–15.6)
[134]
Commercial productSA (25)PA (50) - [135]
Commercial product from Curcuma longa L. (Plant)SA (125–500),
MRSA (>4500),
SE (500–2000)
EC (2000), EC c (1500), PA (62.5–5000)CA (1000–5000),
SC (5000)
---[136]
Commercial productSA (450)PA (500) -- [137]
Commercial product from Curcuma longa CA (1000),
CG (125)
Ketoconazole
CA (62.5),
CG (1.95)
[138]
Commercial productSA (0.03), BC (0.05)EC (0.225) -- [139]
DemethoxycurcuminZingiber spectabile (Plant)SA (125), BC (125)EC (500) Tetracycline
SA (3.91),
BC (1.95)
Tetracycline
EC (NA)
[133]
(-): not tested; a: multidrug-resistant strain; b: drug-resistant strain; c: clinical isolate; ND: not determine; NA: no activity; LM: Listeria monocytogenes; SA: Staphylococcus aureus; MRSA: Methicillin-resistant Staphylococcus aureus; BC: Bacillus cereus; SE: Staphylococcus epidermidis; EC: Escherichia coli; ST: Salmonella typhimurium; SF: Shigella flexneri; PA: Pseudomonas aeruginosa; CA: Candida albicans; CG: Candida glabrata; SC: Saccharomyces cerevisiae; CN: Cryptococcus neoformans; AN: Aspergillus niger; AF: Aspergillus fumigatus; AV: Aspergillus versicolor; FS: Fusarium solani, TR: Trichophyton rubrum.

2.2. Coumarin

Coumarin (1-benzopyran-2-one) derivatives are chemical compounds of the benzopyrone class, comprised of fused benzene and α-pyrone rings, that can be discovered in bacteria, fungi, and plants [140]. Coumarin is poorly water-soluble; however, thanks to its 4-hydroxy substitution, the compound is water-soluble in slightly alkali conditions [141]. To date, over 1300 coumarin types have been described and obtained through plant extraction or microbial synthesis [142,143]. Natural coumarins can be categorized into four primary classes, including simple coumarins, pyrano coumarins, furanocoumarins, and bicoumarins (dicoumarin), as illustrated in Figure 2 [144,145]. These compounds are mostly isolated from plants, nonetheless, some of them can also be produced by microorganisms [146]. The most prominent coumarin members obtained by microbial sources are novobiocin, clorobiocin, and coumermycin from Streptomyces species. Novobiocin, coumermycin A1 and chlorobiocin are amino-coumarin antibiotics with an antibacterial mechanism of action based on bacteriostatic action through the bacterial DNA gyrase inhibition [147,148]. Novobiocin was confirmed as an effective antibiotic for the handling of infections triggered by multiple resistant Gram-positive bacteria, specifically Staphylococcus epidermidis (S. epidermidis) and S. aureus [149].
The MIC values of coumarins in different subclasses, such as umbelliferone, osthol, peucedanin, imperatorin, etc., isolated from natural sources are presented in Table 1. It can be easily seen that the antimicrobial activity of these representative compounds was less than those of the antibiotics given in the related studies. Although the several studies on the antibacterial mechanism of action of coumarin and its derivatives have depicted that their antibacterial action occurs primarily through bacteriostatic effects by binding the B subunit of microbial DNA gyrase and preventing DNA supercoiling by blocking ATPase activity [150,151,152], further work is necessary to understand the antimicrobial mechanism of these nature-derived coumarins. Additionally, recent studies based on the toxicological properties of coumarins in humans indicates that these compounds have a tolerable dose intake (TDI) of 0.1 mg/kg of body weight [153]. However, it should be noted that the toxicity of natural and synthetic coumarins depends on the position and chemical structure of the substituents groups connected to the coumarin core [154].
Various studies on the antibacterial action and structure relationship of coumarins and their derivatives revealed the action of these compounds on the antibacterial effects of the number and binding position of substituents such as the thiazole ring, halogen, methyl, methoxy, hydroxyl, and amino, attached to the basic skeleton [155,156]. For example, the electron-donating substituents of the phenyl ring such as –OCH3, –CH3, and electron-withdrawing substituents such as NO2 and halogen groups play a remarkable effect on their action [157]. Ben Jannet et al. isolated two coumarins, marmesin and scopoletin, from a plant, Ferula lutea (Poir.) Maire and synthesized their synthetic derivatives from these natural compounds. The isolated and synthesized compounds were evaluated for antibacterial activity against Escherichia coli (E. coli), Pseudomonas aeruginosa (P. aeruginosa), Enterococcus faecalis and S. aureus. Due to the nature of the acyl group and the aryl ring attached to the isoxazole moiety included in both marmesin and scopoletin, the synthesized derivatives exhibited potent antibacterial activity when compared to nature-derived coumarins. Marmesin was esterified with a series of acid chlorides, resulting in the formation of new ester derivatives. The phenyl group substituted with the ester moiety improved the activity compared to the methyl group against the tested Gram-positive bacteria. In addition, the presence of one or more chlorine atoms on branched-chain methyl esters provided more antibacterial activity; the chlorination of straight-chain methyl esters did not. For scopoletin and their derivatives, antibacterial data indicated that introducing the isoxazole moiety into scopoletin improved the antibacterial action. In addition, the antibacterial activity of a synthesized compound bearing p-Cl-phenyl attached to its isoxazole moiety is profitable because of the introduction of a halogen, which is an electron-withdrawing substituent, into the structure. On the other hand, the study revealed that the derivatives that maintain electron-donating substituents, isopropyl, ethyl, methyl, and furan, exhibited less antibacterial activity [158]. In accordance with conducted structure-activity studies, naturally derived coumarins and their substitution with various functional groups can be considered an essential step for the development of antibacterial agents [159,160,161].

2.3. Flavonoids

Flavonoids are a wide class of polyphenolic compounds based on a basic structure of 2-phenyl chroman [162]. On the other side, isoflavonoids own a basic structure of 3-phenyl-chroman which is biogenetically derived from the 2-phenyl chroman skeleton of flavonoids [163]. Until now, more than 8000 flavonoid derivatives have been recognized in nature, as both free state and conjugated state, as ester or glycosidic derivatives [164,165,166]. Flavonoids are generally discovered in plant sources, but they may innately occur in certain microalgae and fungi [167]. As described by Jin et al., based on the oxidation degree of the main heterocycle, flavonoids are categorized into seven subclasses: flavonols, flavones, isoflavones, anthocyanidins, flavanones, flavanols, and chalcones (see Figure 3) [168]. Flavonoids and ısoflavanoids are promising antimicrobial mediators that target different microbial cells and can inhibit virulence features in drug-resistant strains [169,170]. Various studies have proposed that the compounds in this class can display antimicrobial activity through both bacteriostatic and bactericidal effects. Their bacteriostatic effects are associated with their ability to form complexes with the bacterial cell wall to inhibit the growth of bacteria. In detail, they can suppress cell growth by inhibiting microbial cell energy metabolism, nucleic acid synthesis, or cytoplasmic membrane function [171]. Their bactericidal activities are considered to be associated with irreversible damage to the cytoplasmic membrane. As a representative example, Voutquenne-Nazabadioko et al. reported the antimicrobial skills of the purified flavonoid glycosides obtained from Graptophyllum grandulosum plant against Vibrio cholerae, S. aureus, Candida albicans (C. albicans), and Cryptococcus neoformans (C. neoformans) [172]. Their antibacterial mechanism is based on cytoplasmic membrane damage by disturbing the membrane permeability and causing the leakage of cellular constituents. In another study scrutinizing the anti-P. aeruginosa activity and possible mechanism of a flavonoid isolated from Trianthema decandra, 2—(3′, 4′ dihydroxyphenyl) 3, 5, 7—trihydroxy-chromen-4 displayed a bactericidal effect by constraining the FabZ enzyme, according to molecular docking studies, although additional findings are wanted to fully elucidate its mechanism [173]. Additionally, some flavonoids exhibit the aptitude to boost the therapeutic effect when pooled with existing antibiotic drugs [174]. For example, Sathiya Deepika et al. investigated the anti-biofilm efficacy of rutin from Citrus sinensis peels and its synergistic effects in combination with conventional bactericidal antibiotic gentamicin against multidrug-resistant P. aeruginosa. Rutin and a combination of rutin-gentamicin prevented biofilm development by inducing reactive oxygen species (ROS) generation in P. aeruginosa, which led to oxidative stress, induction of cell wall disruption, and eventually to bacteria killing. A synergistic effect was further observed by combining gentamicin with rutin. Herein the bactericidal effect of the flavonoids is related to their antioxidant properties. They act as pro-oxidants against microbial pathogens and cause oxidative stress by generating ROS to induce cell death [175]. Additionally, Table 1 provides the list of MIC values of flavonoids obtained from different plant species. As can be seen in Table 1, some flavonoids exhibit higher antimicrobial activity than conventional antibiotics, indicating their potential to prevent microbial infections. On the other hand, it should be noted that flavonoids with higher MICs than antibiotics can display possible enhanced effects in combination with antibiotics due to their multiple target mechanisms.

2.4. Tannin

Tannins are a heterogeneous class of high-molecular-weight polyphenolic substances [176]. Previously, tannins have been categorized as hydrolyzable tannins and condensed tannins. Accordingly, it was assumed that hydrolyzable tannins included two sub-groups, as gallotannins and ellagitannins. Nevertheless, the existence of some ellagitannins, which cannot be hydrolyzed on account of more additional C-C bonding of the polyphenolic moieties with the polyol unit, caused an update in their classification. Thus, the revised classification of these compounds is sub-categorized into five groups, gallotannins, ellagitannins, condensed tannins (proanthocyanidins), complex tannins and phlorotannins (see Figure 4) [177]. Gallotannins form by one or more galloyl units bounded to a polyol, triterpenoid or catechin unit, while ellagitannins are composed of hexahydroxydiphenoyl esters coupled to sugar, mostly glucose [178,179]. Condensed tannins exist in the plants as a form that is free or bound to protein and fiber [180]. Their chemical structure is formulated of flavan-3-ols that are bound through single C-C bonds, which are typically C4 → C6′ or C4 → C8′ (B-type) or doubly coupled with a further bond at C2 → O → C7′ (A-type) [181,182]. On the other hand, complex tannins are a class of tannins with high molecular weight in which a catechin unit is linked to either gallotannins or ellagitannins [176]. In the last two decades, investigations on their isolations and biological activities have been limited because of, presumably, their complex structure. Melasquanins A–D can be exemplified as complex tannins, which were isolated from Melaleuca squarrosa, but their biological properties need to be elucidated [183]. Additionally, the particular type of tannins, commonly isolated from marine algae, is named phlorotannins due to their occurrence by polymerizing phloroglucinol units [184]. Tannins are naturally found in higher plants and marine algea and also possess a defensive function for the plant against diverse environmental factors, pathogens, or herbivores [179]. Inspired by the biochemical shield against herbivores and pathogens offered by tannins existing in plants, the interest of the scientific community arises on their usage as an antimicrobial agent [185,186,187,188]. Previous studies have revealed the tannins to display several biological features, such as antimicrobial activity. This feature is rooted in their chemical structure, allowing them to own antimicrobial activity through bacteriostatic or bactericidal actions [189]. In detail, the chemical nature of tannins owns plenty of hydroxyl groups, providing them with a hydrophilic character. This mainly allows the tannins to form complexes with proteins or enzymes of microbial cell membrane by hydrogen bonds and hydrophobic interactions, which can affect the morphology of the cell wall and increasing membrane permeability [190]. Another purposed antimicrobial mechanism is the generation of complexes between metal ions and tannins. Tannins may chelate many metal ions, hindering the accessibility of such indispensable ions for microorganisms [189,191].
As mentioned earlier, the phlorotannins discovered in marine algae led to a need for more investigations into their biological potential. Recently, Kim et al. reported that phlorofucofuroeckol-A, extracted from brown alga Eisenia bicyclis, displayed anti-MRSA activity by blocking the production or function of penicillin-binding protein 2a, which is regarded as the primary reason for methicillin resistance [192]. Hereby, this compound can be considered a promising candidate due to its potential in inhibiting the growth of antibiotic resistance related to mediating suppressive effects on methicillin resistance-associated genes. In another study, persimmon tannins from young astringent persimmon fruit showed antibacterial activity with an MIC value of 1000 μg/mL against some MRSA isolates from pork. Performed studies on the mechanism of antibacterial action indicated that persimmon tannins showed bactericidal and bacteriostatic activities by multiple mechanisms, including damage to cell wall and membrane, leading to membrane hyperpolarization, reduction of intracellular ATP concentration, losing bacterial membrane integrity, whole cell protein, and cell cycle depression [187]. Anti-MRSA molecular mechanisms of persimmon tannins were elucidated deeply using transcriptome and metabolome analyses by the same research group. Results demonstrated that persimmon tannins adversely affected the cell membrane permeability and integrity, amino acid, and energy metabolism and also caused iron deprivation [193]. A survey of recently reported antimicrobial activities of representative compounds in different classes of tannins is given in Table 1. Considering the MIC values of the tannin compounds and the relevant antibiotics, further investigations should be concentrated on in vivo assays and clinical trials to depict the effectiveness of these antimicrobial agents in clinical settings.

2.5. Stilbenes

Stilbenes are widely found in plants, but also, their basic forms or various substitutions can be isolated from the pathogenic strains [194]. Although stilbenes are typically encountered in the plant kingdom, several studies on their isolation from microorganisms and marine organisms have also been reported [98,195,196]. In general, these metabolites are unearthed in plants as both free and glycosylated forms [197]. Their chemical structure comprises a 1,2-diphenylethylene core with substituted hydroxyl groups on the aromatic rings [198]. The sorting of different subclasses of stilbenes is a challenge owing to their broad structural diversity; we basically classified them into four main subclasses: monomeric, dimeric, oligomeric, and miscellaneous stilbenes. Although stilbenes stand out in multiple fields owing to their antitumoral [199], antioxidant [200], cardioprotective [201], hypolipidemic [202], and immunosuppressive [203] activities, their antimicrobial properties occupy a noteworthy position in combating various microbial infections. Resveratrol, piceatannol, isorhapontigenin, pinosylvin, and oxyresveratrol are widely recognized monomeric stilbenes (see Figure 5). Among the representative monomeric stilbenes, resveratrol and pterostilbene were commercially available on the bench due to their prominent properties such as antihypertensive, antioxidant, anti-inflammatory, and anti-cancer activities. Thus, investigations on the discovery and biological potential of naturally derived stilbenes are in progress. Various reports on their antimicrobial mechanism have documented that stilbenes induce cell membrane damage and DNA degradation mediated by oxidative stress and increase cell membrane permeability, causing the leakage of intracellular nucleic acids and proteins [204,205,206,207,208]. As a representative example, Longistylin A, a pinosylvin-derived monomeric stilbene isolated from the leaves of Cajanus cajan (L.) Millspaugh, was tested against MRSA, S. aureus, E. coli. and Bacillus cereus (B. cereus). This compound exhibited notable antibacterial activity against tested gram-positive bacterial strains (see Table 1). Moreover, studies on the underlying mechanism of anti-MRSA action revealed that Longistylin A demonstrated bactericidal activity by disrupting bacterial membranes and increasing membrane permeability. Furthermore, Longistylin A exhibited much faster bactericidal activity (3-log reduction in MRSA survival within 8 h) compared to vancomycin used as a positive control, indicating the promising potency of Longistylin A in fighting with MRSA-associated infections [99].
Dimeric stilbenes occur from two monomeric units; for instance, Gnetin C is formed from two resveratrol monomers, or Longusol A arose from two distinct monomeric units, a resveratrol unit and a piceatannol unit [209]. Gnetin D, a dimeric stilbene, was isolated from the roots of Spirotropis longifolia and showed effective antifungal activity against, especially, C. albicans, Candida parapsilosis, and Candida krusei (C. krusei) strains among the ten different tested fungal stains (see Table 1) [94]. Another study on the antibacterial activity of fifteen resveratrol-derived stilbenoids verified that the dehydro-δ-viniferin, a stilbene dimer, displayed the most potent antibacterial activity among others. The mechanism of action of this compound against L. monocytogenes was demonstrated to be accomplished by more than one specific mechanism, including membrane depolarization followed by damaging the cytoplasmic membrane and the destruction of membrane integrity and severe morphological changes [95]. Oligomeric stilbenes are generated by a coupling reaction between monomeric units of stilbenes following the pattern of a homogeneous or heterogeneous oligomerization [210]. Miscellaneous oligomeric stilbenes were described as complex stilbene oligomers with diverse structural skeletons comprising distinct stilbene units excepting resveratrol and oxyresveratrol units in the comprehensive study where Shen et al. updated the classification of stilbenes [211]. In recent decades, prenylated stilbenes, which are considered a class of miscellaneous stilbenes, have been attracting widespread interest due to their unique structures and biological potential. For instance, denticulatains A and B, prenylated stilbenes with stilbene-diterpene type skeleton were isolated from a plant species Macaranga denticulata [210]. In another study, prenylated stilbenes, cajanusins A-D and their derivatives were isolated from Cajanus cajan (See the structure of cajanusin B in Figure 5) [212]. Besides the extensive research on the isolation of these compounds from various natural sources, the studies on their antimicrobial aptitude is still scarce.
Interactions of stilbenoids, particularly resveratrol, with conventional antibiotics have been investigated as combinatorial therapy, which can potentially improve the effectiveness of antimicrobials and hinder the emergence of resistant strains due to the synergistic effect. In vitro antibacterial activity of resveratrol was assessed alone and in combination with the bactericidal antibiotic, colistin, against a collection of colistin-resistant (COL-R) Gram-negative pathogens, including E. coli, Klebsiella pneumoniae, Enterobacter cloacae, Stenotrophomonas maltophilia, Citrobacter braakii, and polymyxin-resistant enterobacterial species [213]. The results revealed that the 512 mg/L concentration of resveratrol did not show antimicrobial activity against all Gram-negative pathogens tested. Nevertheless, in the combination, resveratrol (at 128 mg/L) potentiated the bactericidal effect of colistin (0.5 × MIC and 1 × MIC) against all strains tested except for one of the E. coli strains [213]. However, the synergistic mechanism of resveratrol and colistin against COL-R strains has not been elucidated. In a study investigating the antibacterial activity of the structural analogues of resveratrol, the dimeric compound of the 4,4′ dihydroxy stilbene revealed a strong antibacterial effect with 10 µg/mL of MIC value against S. aureus [208]. Thereupon, the same research group investigated the synergistic effect of this dimeric stilbene compound with antibiotics in their further study [214]. The combination of dimeric stilbene compound with the antibiotics targeting protein synthesis led to the decreased MIC values of antibiotics, including kanamycin, tobramycin, chloramphenicol, tetracycline, musiporin, and erythromycin, indicating the in vitro synergistic effect. The same synergistic effect was also established by combining the dimeric stilbene and kanamycin against kanamycin-resistant laboratory strains and kanamycin-resistant clinical strains. Furthermore, in vivo studies confirmed that the dimeric stilbene ameliorates S. aureus infection in mice, both alone and in combination with kanamycin [214]. However, the mechanism underlying this synergistic outcome has not been ascribed.
In addition, stilbenes have potential synergistic activities to combat microbial infection in combination with conventional antibiotics; it has also been reported that combining these compounds with antibiotics can lead to antagonistic interactions [215,216]. The mechanism of antagonism is proposed to involve a reduction in ROS by stilbenes due to their antioxidant properties. Increased ROS production by antibiotics causes oxidative stress, DNA damage, and, eventually, cell death in bacteria. However, in some cases, the ROS produced by bactericidal antibiotics may be suppressed through the scavenging of free radicals by stilbenes, depending on the concentration of stilbene and the target bacterial species.
To provide deep insight into the antimicrobial activity of stilbenes, several structure-activity relationship studies have been performed to date [194,208,217]. The study on resveratrol structural analogs concluded that the presence of hydroxyl groups on the aromatic rings of stilbenes plays a key role on their antimicrobial activity [208]. However, an increasing number of hydroxy groups did not provoke higher antimicrobial activity. Additionally, the presence of the methoxy group along with the hydroxy group resulted in more potent antibacterial activity. Converting all the active hydroxy groups to acetoxy group or methoxy group caused a drastically reduced antibacterial activity. The partial transformation of hydroxy group to methoxy group resulted in enhanced antibacterial activity as a result of oxidative stress and membrane damage. Additionally, the studied stilbenes are less active against Gram-negative bacteria, as they are taken out of the cell by the efflux pump in Gram-negative bacteria. Though, surprisingly, pinosylvin and 4-Bromo resveratrol were effective even in the presence of efflux pump. This was presumably related to the fact that these compounds cause cell damage within a short time before being pumped out by the efflux pump or they are weak substrates for the efflux pump [208]. Additionally, halogenation and dimerization cause enhanced antibacterial and antifungal properties [218]. However, it should be noted that halogenation might increase cytotoxicity [219].

2.6. Lignans

Lignans are formed from the dimerization of two phenylpropanoid units through oxidative coupling reactions [220]. Their wide structural diversity has resulted in differences in their nomenclature and classification. In detail, lignans can be categorized as lignans, neolignans, or norlignans in accordance with the bonding positions of the two phenylpropanoid units and the case of lack of carbon from the parent lignan skeleton. As can be seen in Figure 6, while the basic lignan structure is composed of two phenylpropanoid (C6C3) units linked by a β-β’ (C8-C8′) bond, neolignan contains dimerization of C6C3 units linked in a form other than β-β’ (C8–C8′) bond [221]. On the other hand, the term ‘norlignans’ is defined as lignans that couple two phenylpropanoid units with a β-β’ bond and have one or more carbon atoms missing than those of the basic lignan skeleton. Interestingly, it should be noted that the missing carbon is assigned in the nomenclature of norlignans, as shown in Figure 1 [222,223,224]. According to the dissimilarities of carbon skeletons, a recent study by Tan et al. classified lignans into six subclasses: dibenzylbutane, tetrahydrofuran, arylnaphthalene, arylnaphthalenelactone, furofuran, and dibenzocyclooctadiene [222]. These compounds are frequently discovered in plants, but they may also be metabolized by gut microbiota in mammals [225]. Plant lignans exist as secoisolariciresinol, matairesinol, lariciresinol, and pinoresinol in a diversity of food sources [226,227]. Lignans from plants are metabolized into mammalian lignans, such as enterolignans, enterodiol, and enterolactone, by intestinal bacteria [228]. The mammalian and plant lignans are distinguished from each other by the presence of phenolic hydroxy groups only in the meta-position of the aromatic rings [229]. Additionally, mammalian lignans are gifted to bind to estrogen receptors owing to their chemical structural similarity to estrogen and thus can serve as antioxidant agents [230]. Many efforts are still being made on the antimicrobial potential of lignans obtained from diverse origins such as microorganisms, and plants because of their various biological potential [231,232]. As a representative example, pinoresinol, one of the structurally basic lignans, was extracted from Cinnamomum camphora leaves and exhibited more potent antibacterial activity against Bacillus subtilis (B. subtilis) and P. aeruginosa than S. aureus, E. coli, and Salmonella enterica [110]. In addition, studies on its mechanism of antibacterial action have reported that pinoresinol displayed bactericidal activity by increasing the permeability of bacterial plasma membrane and damaging the cell wall of B. subtilis and P. aeruginosa [110]. As an example of an arylnaphthalide lignans, justicidin B was previously extracted from various plant families, including Justicia pectoralis, Linum leonii, Phyllanthus polyphyllus, Sesbania drummondii, and Justicia procumbens [233,234,235,236]. However, Jaspars et al. isolated for the first time this compound from a microbial species, a marine-derived bacterium Nocardia sp. ALAA 2000 and reported its remarkable and potent antimicrobial properties (see Table 1) against fourteen microbial strains, demonstrating the promising bioactive aspect of compounds from different sources [113]. In another study, Hwang et al. [106] also isolated an enterolignan precursor, Lariciresinol, from Sambucus williamsii herb and confirmed it to exhibit fungicidal activities by disrupting the fungal plasma membrane of C. albicans (see Table 1). To date, many researchers have demonstrated the antimicrobial activity of lignans and their derivatives from natural sources [237,238], but research to elucidate the mechanisms of action at the molecular level remains unclear.
An in-depth evaluation of their structure-antimicrobial activity relationship is essential for the assessment of these compounds and their use in the design as antimicrobial agents. Koba et al. investigated the association of antimicrobial activity and structure among compounds containing different benzylic oxidation degree and stereochemistry. These compounds exhibit antibacterial activity against the tested Gram-positive bacteria. However, the presence of a carbonyl group at C-9′ of tetrahydrofuran lignans decreased antibacterial activity in the absence of benzylic oxygen. In addition, full oxidation of the benzylic positions on 2,3-dibenzyl-4-butanolide triggers more potent antibacterial activity than that of 2,3-dibenzyl-4-butanolide, thus indicating the importance of benzylic carbonyl groups for a prominent antibacterial effect [239]. Another study investigating the structure-activity relationship of tetrahydrofuran lignan, 9-O,9′-O-demethyl (+)-virgatusin indicated that the antibacterial activity might vary on the presence and location of methoxy substitutions. For example, the 3’-methoxy group contributed to higher effective antibacterial activity than that of the 4’-substituent. Additionally, the existence of the 3,4-methylenedioxy group on the 7-phenyl group played an essential role in the enhanced antibacterial activity [240]. Similarly, a study on virgatusin and its derivatives supported the aforementioned results. For these tetrahydrofuran lignans, two methoxy groups at C-9/9′ and a 3,4-methylenedioxyphenyl group at C-7 improved antifungal activity. Additionally, among virgatusin and its derivatives, the substitution of the 4-methoxyphenyl group at C-7′ resulted in the highest antifungal activity [241]. Further investigations are in need of elucidating the structure-activity relationship, especially addressing other subclasses of lignans.

2.7. Quinones

Quinones are structurally defined as cyclohexadiendiones possessing carbonyl groups in the 1,2 or 1,4 positions relative to each other [242]. These compounds are found in numerous natural sources, including plants, bacteria, fungi [243], and marine organisms [244] but can also be found in some animals such as aphids, sea urchins, lac insects, and certain scale insects [245,246]. Quinones are divided into four subclasses that include benzoquinones, naphthoquinones, anthraquinones and phenanthraquinones (see Figure 7) [14]. To date, benzoquinones and naphthoquinones derivatives have been isolated from innumerable plant sources and reported to show significant antimicrobial activities [116,124,247]. Phenanthrenequinone derivatives have also been extracted from various plants, such as Pleione bulbocodioides and Cannabis sativa [248,249]. Nevertheless, a major deficit in the literature subsists regarding the antimicrobial activities of phenanthrenequinones.
According to previously reported studies, quinones display antibacterial activity by bacteriostatic and/or bactericidal modes [250,251,252]. Although the mechanism by which quinone causes antimicrobial activity is complex, it has been reported that quinones generate ROS through redox cycling with their semiquinone radicals, causing intracellular oxidative stress and, thus, cell membrane damage [253,254]. A list of recently published investigations indicating good and moderate antimicrobial properties of representative compounds in different quinone classes is provided in Table 1. The discovery of these compounds provides new insight into novel antimicrobial agents for pharmaceutical development. However, more analyses are desirable on the structure-function and antimicrobial mechanisms of quinones.

2.8. Curcuminoids

Curcuminoids are linear, diarylheptanoid molecules consisting of curcumin, demethoxycurcumin, bisdemethoxycurcumin and their analogues [255,256]. Curcuminoids are commonly isolated from turmeric, a member of the ginger family (Zingiberaceae), and other plant species, such as Curcuma zedoaria, Curcuma manga, Costus speciosus, Curcuma aromatica, Curcuma xanthorrhiza, Curcuma phaeocaulis, Zingiber cassumunar and Etlingera elatior [257,258,259]. Among the curcuminoids, curcumin has been applied in several studies as an outcome of its various biological properties, such as anti-inflammatory [260,261], antimicrobial [262,263], and antitumoral [264,265] activities. Clinical studies disclosed that curcumin is non-toxic to humans at high dosages, but its low bioavailability hinders its therapeutic applications [266,267]. Several strategies have been performed to enhance the bioavailability of curcumin, including the usage of liposomal curcumin, nanocurcumin and more recently deep eutectic solvents [268,269,270,271]. As mentioned above, curcumin, which stands out with its various biological benefits, is commercially available. Therefore, Table 1 includes MIC values that evaluate the antibacterial activity of commercial curcumin along with curcumin isolated from natural sources. The antimicrobial mechanism of curcuminoids has been widely investigated [272,273,274,275,276]. For example, Sivasothy et al. [133] isolated five different flavonoids and curcuminoids from the rhizomes of Zingiber spectabile. The antibacterial data indicated that curcuminoids present higher antibacterial activity than flavonoid-derived compounds [133]. In another promising study, Adamczak et al. [136] evaluated the usefulness of commercial curcumin from Curcuma longa L. against more than 100 strains from 19 different species, as represented in Table 1.
To date, curcuminoids have been the subject of numerous investigations and have been reviewed deeply. In particular, curcumin has been reported to own antibacterial activity by bacteriostatic and bactericidal modes against a wide range of bacterial strains. The bactericidal action of curcumin has been attributed to its diffusion across the bacterial membrane of S. aureus and E. coli due to its amphipathic and lipophilic character, increasing membrane permeability, leaking of intracellular constituents, and ultimately causing cell death [276].
Investigations on the bacteriostatic mechanism of curcumin at the molecular level proposed that curcumin inhibits bacterial growth by attacking different targets such as the DNA, protein, cell wall, cell membrane and quorum-sensing systems in the bacteria. As a representative example, the bacteriostatic action of curcumin against B. subtilis was reported to occur by increasing the GTPase activity of protein FtsZ, which plays an essential role in the division of bacterial cells, depending on the concentration [274]. Additionally, as curcumin is a photosensitizer, several reports surveyed its usage in antimicrobial photodynamic therapy [269,277,278,279,280,281,282,283]. The bactericidal action of curcumin-based photodynamic inactivation therapy against L. monocytogenes was proven to be membrane protein degradation and increased membrane permeability triggered by oxidative stress through intracellular ROS production [284]. In addition, another study elucidating the antibacterial mechanism of L. monocytogenes ascertained that the bactericidal efficacy of curcumin-based photodynamic inactivation therapy is a result of cytoplasmic DNA and protein damage [285]. According to the previous studies on curcumin, it should be noted that the mechanism of antibacterial action of curcumin may differ depending on the strain tested.
Besides their antibacterial activities, many studies have also investigated the fungistatic and fungicidal activities of curcumins against various fungal strains as potential antifungal agents. A study on the in vitro antifungal activity of curcumin has demonstrated its antifungal activity against Candida species, including C. albicans, Candida glabrata (C. glabrata), and C. krusei, and also indicated that curcumin displayed fungistatic activity by binding to the membrane ergosterol of C. albicans. However, the same mechanism was not observed against C. glabrata and C. krusei, indicating that the interaction of curcumin-ergosterol is not a single mechanism of action for curcumin [138]. As a fungicidal agent, it has been proposed that curcuminoids bind to residues on the fungal cell membrane, causing cell membrane disruption, leakage of intracellular contents, and, eventually, cell death.
In addition to functioning as an antimicrobial compound by itself, curcumin has also been scrutinized for potential effects in combination with conventional antibiotics. Curcumin may have synergistic effects in combination with bacteriostatic antibiotics and fungistatic drugs to improve antimicrobial activity. For C. albicans, Ferreira-Pereira et al. reported that curcumin potentiates synergistically the antifungal effect of fluconazole against a clinical isolate of C. albicans possessing a multiple drug resistance phenotype [286]. in another study, the combination of curcumin with conventional fungicidal drugs, including fluconazole, ketoconazole, miconazole, itraconazole, voriconazole, amphotericin B, and nystatin proved a 10–35-fold reduction in the MIC80 values of drugs against 21 clinical C. albicans isolates [287]. The mechanism of synergistic activity of curcumin with amphotericin B and fluconazole was hypothesized to occur through the accumulation of ROS since the addition of an antioxidant could reverse it [287]. Moreover, various in vitro studies revealed the synergistic activity of curcumin with conventional antibiotics against several bacterial strains, such as S. aureus, MRSA, E. coli, P. aeruginosa [134,288,289,290,291,292,293].
On the other side, the combination of curcumin with the bactericidal antibiotic, ciprofloxacin, antagonized the bactericidal activity of ciprofloxacin against Salmonella Typhi and Salmonella Typhimurium (S. Typhimurium). Furthermore, the studies on the elucidation of the antogination mechanism indicated that curcumin reduces the antibacterial effect of the antibiotic due to its antioxidant properties. Accordingly, the oxidative stress induced by ciprofloxacin is suppressed by lowering ROS-induced filamentation in S. Typhimurium in the presence of curcumin [294]. These results are a significant warning against the unexpected consequences of the combination of antibiotics, which function by increasing oxidative stress, with antioxidants such as curcumin.
In accordance with the molecular structure of curcuminoids, the presence of phenolic hydroxyl groups acts as an electron donating group, interacts with the bacterial membrane, and thus increases the permeability of the bacterial membrane, enabling the targeting of antibacterial agents to bacterial cells [135]. In addition, the β-diketone moiety of curcuminoids might form a hydrogen bond containing a six-membered ring through keto−enol tautomerism as illustrated in Figure 8 [295,296]. The formed six-membered ring is a potential substrate for the aldo-keto reductase (AKR) enzymes, which are NADPH-dependent oxidoreductases. So, curcuminoids can be bound to the bacterial membrane and cause an increase in its AKR activity, the deficit of intracellular NADPH, and henceforward increase membrane permeability and bacterial cell death [297]. Although there are various studies on the contribution of the phenolic methoxy group in demethoxycurcumin and curcumin to their anti-inflammatory, anticancer and antioxidant activity [298,299,300,301], there remains a need for elucidation of the contribution of the methoxy group to antimicrobial activity.

3. Limitations in the Therapeutic Usage of Natural Phenolic Compounds

Many natural compounds show various biological properties such as antimicrobial, antioxidant, antitumor, and anti-inflammatory activities. However, as a result of their poor water solubility and stability, these compounds can not be transported in the organism to the target site, significantly limiting their application. Alternative strategies for enhancing the efficacy of natural antimicrobials, or overcoming their limitations to use, have been explored and are still being developed. These have included encapsulation of the antimicrobial compound in varied manners, and modification of polar functional groups chemically or enzymatically to the relative compound, etc.
On the other hand, the conspicuous challenge is to overcome existing limitations on the use of natural phenolic compounds for therapeutic use, such as supply and identification of materials, scaled-up production, high throughput screening assays and possible safety issues [302,303,304,305,306]. In the isolation of large quantities of a particular natural compound, some circumstances may be challenging, such as low product yields or long growing periods [307]. In addition, considering the problematic issue of the extinction of plants and other organisms due to environmental change, the large-scale supply of plants, marine organisms, or animals for industrial-scale production could have dire consequences on the ecological balance. Although progress has been made, the supply problem affecting the industrial-scale manufacture can be unraveled by developing an artificial biosynthetic pathway in cooperation with genetic engineering strategies.
Validation, characterization, and standardization of discovered natural compounds are critical for their approval into mainstream medicine. The quality of chemical components in a plant species can be impacted by different factors such as the age of the plant, geographical and seasonal variations, time, method of collection, etc. Hence, quality evaluation of the source product is time-consuming and costly, as the safety, efficacy, and quality of the isolated compound depend on the quality of the source product [308]. Efforts to provide a reliable and sustainable source product for the desired quantity and quality of the pharmacologically active substance are needed in combination with modern genetic engineering and agricultural technological methods [309].

4. Conclusions

Natural phenolic compounds with multiple-target mechanisms stand out as promising candidates for microbial infections. The mechanisms surveyed and specified in detail for each subclass include multiple mechanisms such as the hindrance of microbial cell wall biosynthesis, protein synthesis, nucleic acids synthesis, metabolic pathways, and disruption of cell membrane integrity. Hence, natural phenolic compounds do not present a specific target mechanism but multiple antibacterial mechanisms that have already been discovered and/or have yet to be discovered. The latest improvements in the therapeutic usage of nature-based compounds reveal their tremendous potential. The clinical applicability of these bioactive compounds requires a multi-disciplinary approach and synchronized actions of various fields. To date, there are still many aspects to be clarified regarding the structure-function relationships of many compounds that have been discovered or are yet to be discovered. The consequence of such novel therapeutic endeavors will open new doors in the health and pharmaceutical sector to cope with many diseases, including infections instigated by multidrug-resistant organisms.

Author Contributions

Conceptualization, K.E., J.M.S., A.A.B. and R.L.R.; methodology, K.E.; investigation, K.E. and J.M.S.; resources, R.L.R.; data curation, K.E., J.M.S. and A.A.B.; writing—original draft preparation, K.E.; writing—review and editing, K.E., J.M.S., A.A.B. and R.L.R.; visualization, K.E.; supervision, R.L.R., A.A.B. and J.M.S.; funding acquisition, R.L.R. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial support by the Portuguese Foundation for Science and Technology (FCT) through the doctoral grant and junior research contract with the reference number PD/BD/150521/2019 (K.E.) and CEECIND/01026/2018 (J.M.S.), respectively.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Harvey, A.L.; Edrada-Ebel, R.; Quinn, R.J. The Re-Emergence of Natural Products for Drug Discovery in the Genomics Era. Nat. Rev. Drug Discov. 2015, 14, 111–129. [Google Scholar] [CrossRef] [Green Version]
  2. 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]
  3. Yoshikawa, T.T. Antimicrobial Resistance and Aging: Beginning of the End of the Antibiotic Era? J. Am. Geriatr. Soc. 2002, 50, 226–229. [Google Scholar] [CrossRef] [Green Version]
  4. Laws, M.; Shaaban, A.; Rahman, K.M. Antibiotic Resistance Breakers: Current Approaches and Future Directions. FEMS Microbiol. Rev. 2019, 43, 490–516. [Google Scholar] [CrossRef] [Green Version]
  5. Brown, D. Antibiotic Resistance Breakers: Can Repurposed Drugs Fill the Antibiotic Discovery Void? Nat. Rev. Drug Discov. 2015, 14, 821–832. [Google Scholar] [CrossRef]
  6. Pancu, D.F.; Scurtu, A.; Macasoi, I.G.; Marti, D.; Mioc, M.; Soica, C.; Coricovac, D.; Horhat, D.; Poenaru, M.; Dehelean, C. Antibiotics: Conventional Therapy and Natural Compounds with Antibacterial Activity—A Pharmaco-Toxicological Screening. Antibiotics 2021, 10, 401. [Google Scholar] [CrossRef]
  7. McKeny, P.T.; Nessel, T.A.; Zito, P.M. Antifungal Antibiotics; StatPearls Publishing: Tampa, FL, USA, 2022. [Google Scholar]
  8. Kapoor, G.; Saigal, S.; Elongavan, A. Action and Resistance Mechanisms of Antibiotics: A Guide for Clinicians. J. Anaesthesiol. Clin. Pharmacol. 2017, 33, 300–305. [Google Scholar] [CrossRef]
  9. Bautista-Baños, S.; Hernández-Lauzardo, A.N.; Velázquez-del Valle, M.G.; Hernández-López, M.; Ait Barka, E.; Bosquez-Molina, E.; Wilson, C.L. Chitosan as a Potential Natural Compound to Control Pre and Postharvest Diseases of Horticultural Commodities. Crop Prot. 2006, 25, 108–118. [Google Scholar] [CrossRef]
  10. Chew, J.; Peh, S.-C.; Sin Yeang, T. Non-Microbial Natural Products That Inhibit Drug-Resistant Staphylococcus Aureus. In Staphylococcus Aureus; Hemeg, H., Ozbak, H., Afrin, F., Eds.; IntechOpen: London, UK, 2019; pp. 1–30. ISBN 978-1-78984-592-1. [Google Scholar]
  11. Rahim, M.A.; Kristufek, S.L.; Pan, S.; Richardson, J.J.; Caruso, F. Phenolic Building Blocks for the Assembly of Functional Materials. Angew. Chem. Int. Ed. 2019, 58, 1904–1927. [Google Scholar] [CrossRef]
  12. Oliver, S.; Vittorio, O.; Cirillo, G.; Boyer, C. Enhancing the Therapeutic Effects of Polyphenols with Macromolecules. Polym. Chem. 2016, 7, 1529–1544. [Google Scholar] [CrossRef]
  13. Shavandi, A.; Bekhit, A.E.-D.A.; Saeedi, P.; Izadifar, Z.; Bekhit, A.A.; Khademhosseini, A. Polyphenol Uses in Biomaterials Engineering. Biomaterials 2018, 167, 91–106. [Google Scholar] [CrossRef] [PubMed]
  14. Gan, R.-Y.; Chan, C.-L.; Yang, Q.-Q.; Li, H.-B.; Zhang, D.; Ge, Y.-Y.; Gunaratne, A.; Ge, J.; Corke, H. Bioactive Compounds and Beneficial Functions of Sprouted Grains. In Sprouted Grains; Elsevier: Duxford, UK, 2019; pp. 191–246. ISBN 978-0-12-811525-1. [Google Scholar]
  15. Özçelik, B.; Kartal, M.; Orhan, I. Cytotoxicity, Antiviral and Antimicrobial Activities of Alkaloids, Flavonoids, and Phenolic Acids. Pharm. Biol. 2011, 49, 396–402. [Google Scholar] [CrossRef] [PubMed]
  16. Arif, T.; Bhosale, J.D.; Kumar, N.; Mandal, T.K.; Bendre, R.S.; Lavekar, G.S.; Dabur, R. Natural Products—Antifungal Agents Derived from Plants. J. Asian Nat. Prod. Res. 2009, 11, 621–638. [Google Scholar] [CrossRef]
  17. Jiang, F.; Dusting, G.J. Natural Phenolic Compounds as Cardiovascular Therapeutics: Potential Role of Their Antiinflammatory Effects. Curr. Vasc. Pharmacol. 2003, 1, 135–156. [Google Scholar] [CrossRef]
  18. Medeiros, K.C.P.; Figueiredo, C.A.V.; Figueredo, T.B.; Freire, K.R.L.; Santos, F.A.R.; Alcantara-Neves, N.M.; Silva, T.M.S.; Piuvezam, M.R. Anti-Allergic Effect of Bee Pollen Phenolic Extract and Myricetin in Ovalbumin-Sensitized Mice. J. Ethnopharmacol. 2008, 119, 41–46. [Google Scholar] [CrossRef]
  19. Hahn, D.; Bae, J.-S. Recent Progress in the Discovery of Bioactive Components from Edible Natural Sources with Antithrombotic Activity. J. Med. Food 2018, 22, 109–120. [Google Scholar] [CrossRef]
  20. Brglez Mojzer, E.; Knez Hrnčič, M.; Škerget, M.; Knez, Ž.; Bren, U. Polyphenols: Extraction Methods, Antioxidative Action, Bioavailability and Anticarcinogenic Effects. Molecules 2016, 21, 901. [Google Scholar] [CrossRef]
  21. Gutiérrez-Larraínzar, M.; Rúa, J.; Caro, I.; de Castro, C.; de Arriaga, D.; García-Armesto, M.R.; del Valle, P. Evaluation of Antimicrobial and Antioxidant Activities of Natural Phenolic Compounds against Foodborne Pathogens and Spoilage Bacteria. Food Control 2012, 26, 555–563. [Google Scholar] [CrossRef]
  22. Madrigal-Santillán, E.; Madrigal-Bujaidar, E.; Álvarez-González, I.; Sumaya-Martínez, M.T.; Gutiérrez-Salinas, J.; Bautista, M.; Morales-González, Á.; García-Luna y González-Rubio, M.; Aguilar-Faisal, J.L.; Morales-González, J.A. Review of Natural Products with Hepatoprotective Effects. World J. Gastroenterol. 2014, 20, 14787–14804. [Google Scholar] [CrossRef]
  23. 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]
  24. Lin, D.; Xiao, M.; Zhao, J.; Li, Z.; Xing, B.; Li, X.; Kong, M.; Li, L.; Zhang, Q.; Liu, Y.; et al. An Overview of Plant Phenolic Compounds and Their Importance in Human Nutrition and Management of Type 2 Diabetes. Molecules 2016, 21, 1374. [Google Scholar] [CrossRef]
  25. Kougan, G.B.; Tabopda, T.; Kuete, V.; Verpoorte, R. Simple Phenols, Phenolic Acids, and Related Esters from the Medicinal Plants of Africa. In Medicinal Plant Research in Africa; Elsevier: London, UK, 2013; pp. 225–249. ISBN 978-0-12-405927-6. [Google Scholar]
  26. Stich, H.F.; Rosin, M.P.; Wu, C.H.; Powrie, W.D. The Action of Transition Metals on the Genotoxicity of Simple Phenols, Phenolic Acids and Cinnamic Acids. Cancer Lett. 1981, 14, 251–260. [Google Scholar] [CrossRef]
  27. Kumar, N.; Goel, N. Phenolic Acids: Natural Versatile Molecules with Promising Therapeutic Applications. Biotechnol. Rep. 2019, 24, e00370. [Google Scholar] [CrossRef]
  28. Robbins, R.J. Phenolic Acids in Foods:  An Overview of Analytical Methodology. J. Agric. Food Chem. 2003, 51, 2866–2887. [Google Scholar] [CrossRef]
  29. Muller, A.G.; Sarker, S.D.; Saleem, I.Y.; Hutcheon, G.A. Delivery of Natural Phenolic Compounds for the Potential Treatment of Lung Cancer. Daru 2019, 27, 433–449. [Google Scholar] [CrossRef] [Green Version]
  30. Liwa, A.C.; Barton, E.N.; Cole, W.C.; Nwokocha, C.R. Chapter 15—Bioactive Plant Molecules, Sources and Mechanism of Action in the Treatment of Cardiovascular Disease. In Pharmacognosy; Badal, S., Delgoda, R., Eds.; Academic Press: Boston, MA, USA, 2017; pp. 315–336. ISBN 978-0-12-802104-0. [Google Scholar]
  31. Ablikim, G.; Bobakulov, K.; Li, J.; Yadikar, N.; Aisa, H.A. Two New Glucoside Derivatives of Truxinic and Cinnamic Acids from Lavandula angustifolia mill. Nat. Prod. Res. 2021, 35, 2526–2534. [Google Scholar] [CrossRef]
  32. Buxton, T.; Takahashi, S.; Eddy Doh, A.-M.; Baffoe-Ansah, J.; Owusu, E.O.; Kim, C.-S. Insecticidal Activities of Cinnamic Acid Esters Isolated from Ocimum gratissimum L. and Vitellaria paradoxa Gaertn Leaves against Tribolium castaneum Hebst (Coleoptera: Tenebrionidae). Pest Manag. Sci. 2020, 76, 257–267. [Google Scholar] [CrossRef]
  33. Pietta, P.; Minoggio, M.; Bramati, L. Plant Polyphenols: Structure, Occurrence and Bioactivity. In Studies in Natural Products Chemistry; Rahman, A., Ed.; Bioactive Natural Products (Part I); Elsevier: Amsterdam, The Netherlands, 2003; Volume 28, pp. 257–312. [Google Scholar]
  34. Deng, Z.; Li, C.; Luo, D.; Teng, P.; Guo, Z.; Tu, X. A New Cinnamic Acid Derivative from Plant-Derived Endophytic Fungus Pyronema sp. Nat. Prod. Res. 2017, 31, 2413–2419. [Google Scholar] [CrossRef]
  35. Parthasarathy, R.; Chandrika, M.; Yashavantha Rao, H.C.; Kamalraj, S.; Jayabaskaran, C.; Pugazhendhi, A. Molecular Profiling of Marine Endophytic Fungi from Green Algae: Assessment of Antibacterial and Anticancer Activities. Process Biochem. 2020, 96, 11–20. [Google Scholar] [CrossRef]
  36. Ramalingam, V.; Narendra Kumar, N.; Harshavardhan, M.; Sampath Kumar, H.M.; Tiwari, A.K.; Suresh Babu, K.; Mudiam, M.K.R. Chemical Profiling of Marine Seaweed Halimeda gracilis Using UPLC-ESI-Q-TOF-MSE and Evaluation of Anticancer Activity Targeting PI3K/AKT and Intrinsic Apoptosis Signaling Pathway. Food Res. Int. 2022, 157, 111394. [Google Scholar] [CrossRef]
  37. de Siqueira, K.A.; Liotti, R.G.; de Sousa, J.R.; Vendruscullo, S.J.; de Souza, G.B.; de Vasconcelos, L.G.; Januário, A.H.; de Oliveira Mendes, T.A.; Soares, M.A. Streptomyces griseocarneus R132 Expresses Antimicrobial Genes and Produces Metabolites That Modulate Galleria mellonella Immune System. 3 Biotech 2021, 11, 396. [Google Scholar] [CrossRef]
  38. Keman, D.; Soyer, F. Antibiotic-Resistant Staphylococcus Aureus Does Not Develop Resistance to Vanillic Acid and 2-Hydroxycinnamic Acid after Continuous Exposure in Vitro. ACS Omega 2019, 4, 15393–15400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. El-Zawawy, N.A.; Ali, S.S.; Khalil, M.A.; Sun, J.; Nouh, H.S. Exploring the Potential of Benzoic Acid Derived from the Endophytic Fungus Strain Neurospora crassa SSN01 as a Promising Antimicrobial Agent in Wound Healing. Microbiol. Res. 2022, 262, 127108. [Google Scholar] [CrossRef] [PubMed]
  40. 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, e7413504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Sánchez-Maldonado, A.F.; Schieber, A.; Gänzle, M.G. Structure–Function Relationships of the Antibacterial Activity of Phenolic Acids and Their Metabolism by Lactic Acid Bacteria. J. Appl. Microbiol. 2011, 111, 1176–1184. [Google Scholar] [CrossRef]
  42. Bouarab-Chibane, L.; Forquet, V.; Lantéri, P.; Clément, Y.; Léonard-Akkari, L.; Oulahal, N.; Degraeve, P.; Bordes, C. Antibacterial Properties of Polyphenols: Characterization and QSAR (Quantitative Structure–Activity Relationship) Models. Front. Microbiol. 2019, 10, 829. [Google Scholar] [CrossRef] [Green Version]
  43. Campos, F.M.; Couto, J.A.; Figueiredo, A.R.; Tóth, I.V.; Rangel, A.O.S.S.; Hogg, T.A. Cell Membrane Damage Induced by Phenolic Acids on Wine Lactic Acid Bacteria. Int. J. Food Microbiol. 2009, 135, 144–151. [Google Scholar] [CrossRef]
  44. Lou, Z.; Wang, H.; Rao, S.; Sun, J.; Ma, C.; Li, J. P-Coumaric Acid Kills Bacteria through Dual Damage Mechanisms. Food Control 2012, 25, 550–554. [Google Scholar] [CrossRef]
  45. Wen, A.; Delaquis, P.; Stanich, K.; Toivonen, P. Antilisterial Activity of Selected Phenolic Acids. Food Microbiol. 2003, 20, 305–311. [Google Scholar] [CrossRef]
  46. Cueva, C.; Moreno-Arribas, M.V.; Martín-Álvarez, P.J.; Bills, G.; Vicente, M.F.; Basilio, A.; Rivas, C.L.; Requena, T.; Rodríguez, J.M.; Bartolomé, B. Antimicrobial Activity of Phenolic Acids against Commensal, Probiotic and Pathogenic Bacteria. Res. Microbiol. 2010, 161, 372–382. [Google Scholar] [CrossRef]
  47. Andrade, M.; Benfeito, S.; Soares, P.; Magalhães e Silva, D.; Loureiro, J.; Borges, A.; Borges, F.; Simões, M. Fine-Tuning of the Hydrophobicity of Caffeic Acid: Studies on the Antimicrobial Activity against Staphylococcus aureus and Escherichia coli. RSC Adv. 2015, 5, 53915–53925. [Google Scholar] [CrossRef] [Green Version]
  48. Araújo, M.O.; Freire Pessoa, H.L.; Lira, A.B.; Castillo, Y.P.; de Sousa, D.P. Synthesis, Antibacterial Evaluation, and QSAR of Caffeic Acid Derivatives. J. Chem. 2019, 2019, 1–9. [Google Scholar] [CrossRef]
  49. Chanwitheesuk, A.; Teerawutgulrag, A.; Kilburn, J.D.; Rakariyatham, N. Antimicrobial Gallic Acid from Caesalpinia mimosoides Lamk. Food Chem. 2007, 100, 1044–1048. [Google Scholar] [CrossRef]
  50. Rashed, K.; Ćirić, A.; Glamočlija, J.; Soković, M. Antibacterial and Antifungal Activities of Methanol Extract and Phenolic Compounds from Diospyros virginiana L. Ind. Crops Prod. 2014, 59, 210–215. [Google Scholar] [CrossRef]
  51. Osamudiamen, P.M.; Oluremi, B.B.; Oderinlo, O.O.; Aiyelaagbe, O.O. Trans-Resveratrol, Piceatannol and Gallic Acid: Potent Polyphenols Isolated from Mezoneuron benthamianum Effective as Anticaries, Antioxidant and Cytotoxic Agents. Sci. Afr. 2020, 7, e00244. [Google Scholar] [CrossRef]
  52. Heleno, S.A.; Ferreira, I.C.F.R.; Esteves, A.P.; Ćirić, A.; Glamočlija, J.; Martins, A.; Soković, M.; Queiroz, M.J.R.P. Antimicrobial and Demelanizing Activity of Ganoderma lucidum Extract, p-Hydroxybenzoic and Cinnamic Acids and Their Synthetic Acetylated Glucuronide Methyl Esters. Food Chem. Toxicol. 2013, 58, 95–100. [Google Scholar] [CrossRef]
  53. Ma, Y.-M.; Qiao, K.; Kong, Y.; Guo, L.-X.; Li, M.-Y.; Fan, C. A New P-Hydroxybenzoic Acid Derivative from an Endophytic Fungus Penicillium sp. of Nerium indicum. J. Asian Nat. Prod. Res. 2017, 19, 1245–1251. [Google Scholar] [CrossRef]
  54. Ren, B.; Xia, B.; Li, W.; Wu, J.; Zhang, H. Two Novel Phenolic Compounds from Stenoloma chusanum and Their Antifungal Activity. Chem. Nat. Compd. 2009, 45, 182–186. [Google Scholar] [CrossRef]
  55. Ajayi, O.S.; Aderogba, M.A.; Akinkunmi, E.O.; Obuotor, E.M.; Majinda, R.R.T. Bioactive Compounds from Nauclea latifolia Leaf Extracts. J. King Saud Univ.-Sci. 2020, 32, 2419–2425. [Google Scholar] [CrossRef]
  56. Lenta, B.N.; Weniger, B.; Antheaume, C.; Noungoue, D.T.; Ngouela, S.; Assob, J.C.N.; Vonthron-Sénécheau, C.; Fokou, P.A.; Devkota, K.P.; Tsamo, E.; et al. Anthraquinones from the Stem Bark of Stereospermum zenkeri with Antimicrobial Activity. Phytochemistry 2007, 68, 1595–1599. [Google Scholar] [CrossRef]
  57. Pham, T.V.; Hoang, H.N.T.; Nguyen, H.T.; Nguyen, H.M.; Huynh, C.T.; Vu, T.Y.; Do, A.T.; Nguyen, N.H.; Do, B.H. Anti-Inflammatory and Antimicrobial Activities of Compounds Isolated from Distichochlamys benenica. BioMed Res. Int. 2021, 2021, e6624347. [Google Scholar] [CrossRef] [PubMed]
  58. Navarro-García, V.M.; Rojas, G.; Avilés, M.; Fuentes, M.; Zepeda, G. In Vitro Antifungal Activity of Coumarin Extracted from Loeselia mexicana Brand. Mycoses 2011, 54, e569–e571. [Google Scholar] [CrossRef] [PubMed]
  59. Karakaya, S.; Şimşek, D.; Özbek, H.; Güvenalp, Z.; Altanlar, N.; Kazaz, C.; Kılıc, C. Antimicrobial Activities of Extracts and Isolated Coumarins from the Roots of Four Ferulago Species Growing in Turkey. Iran. J. Pharm. Res. 2019, 18, 1516–1529. [Google Scholar] [CrossRef]
  60. Rosselli, S.; Maggio, A.; Bellone, G.; Formisano, C.; Basile, A.; Cicala, C.; Alfieri, A.; Mascolo, N.; Bruno, M. Antibacterial and Anticoagulant Activities of Coumarins Isolated from the Flowers of Magydaris tomentosa. Planta Med. 2007, 73, 116–120. [Google Scholar] [CrossRef] [PubMed]
  61. Tan, N.; Yazıcı-Tütüniş, S.; Bilgin, M.; Tan, E.; Miski, M. Antibacterial Activities of Pyrenylated Coumarins from the Roots of Prangos hulusii. Molecules 2017, 22, 1098. [Google Scholar] [CrossRef] [PubMed]
  62. Tada, Y.; Shikishima, Y.; Takaishi, Y.; Shibata, H.; Higuti, T.; Honda, G.; Ito, M.; Takeda, Y.; Kodzhimatov, O.K.; Ashurmetov, O.; et al. Coumarins and γ-Pyrone Derivatives from Prangos pabularia: Antibacterial Activity and Inhibition of Cytokine Release. Phytochemistry 2002, 59, 649–654. [Google Scholar] [CrossRef]
  63. Tian, S.-Z.; Pu, X.; Luo, G.; Zhao, L.-X.; Xu, L.-H.; Li, W.-J.; Luo, Y. Isolation and Characterization of New P-Terphenyls with Antifungal, Antibacterial, and Antioxidant Activities from Halophilic Actinomycete Nocardiopsis gilva YIM 90087. J. Agric. Food Chem. 2013, 61, 3006–3012. [Google Scholar] [CrossRef]
  64. Dalisay, D.S.; Williams, D.E.; Wang, X.L.; Centko, R.; Chen, J.; Andersen, R.J. Marine Sediment-Derived Streptomyces Bacteria from British Columbia, Canada Are a Promising Microbiota Resource for the Discovery of Antimicrobial Natural Products. PLoS ONE 2013, 8, e77078. [Google Scholar] [CrossRef]
  65. Karunai Raj, M.; Balachandran, C.; Duraipandiyan, V.; Agastian, P.; Ignacimuthu, S. Antimicrobial Activity of Ulopterol Isolated from Toddalia asiatica (L.) Lam.: A Traditional Medicinal Plant. J. Ethnopharmacol. 2012, 140, 161–165. [Google Scholar] [CrossRef]
  66. Widelski, J.; Luca, S.V.; Skiba, A.; Chinou, I.; Marcourt, L.; Wolfender, J.-L.; Skalicka-Wozniak, K. Isolation and Antimicrobial Activity of Coumarin Derivatives from Fruits of Peucedanum luxurians tamamsch. Molecules 2018, 23, 1222. [Google Scholar] [CrossRef]
  67. Mileski, K.S.; Trifunović, S.S.; Ćirić, A.D.; Šakić, Ž.M.; Ristić, M.S.; Todorović, N.M.; Matevski, V.S.; Marin, P.D.; Tešević, V.V.; Džamić, A.M. Research on Chemical Composition and Biological Properties Including Antiquorum Sensing Activity of Angelica Pancicii vandas Aerial Parts and Roots. J. Agric. Food Chem. 2017, 65, 10933–10949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Widelski, J.; Popova, M.; Graikou, K.; Glowniak, K.; Chinou, I. Coumarins from Angelica lucida L.—Antibacterial Activities. Molecules 2009, 14, 2729–2734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Dongfack, M.D.J.; Lallemand, M.-C.; Kuete, V.; Mbazoa, C.D.; Wansi, J.-D.; Trinh-van-Dufat, H.; Michel, S.; Wandji, J. A New Sphingolipid and Furanocoumarins with Antimicrobial Activity from Ficus exasperata. Chem. Pharm. Bull. 2012, 60, 1072–1075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Walasek, M.; Grzegorczyk, A.; Malm, A.; Skalicka-Woźniak, K. Bioactivity-Guided Isolation of Antimicrobial Coumarins from Heracleum mantegazzianum Sommier & Levier (Apiaceae) Fruits by High-Performance Counter-Current Chromatography. Food Chem. 2015, 186, 133–138. [Google Scholar] [CrossRef] [PubMed]
  71. Ngameni, B.; Kuete, V.; Simo, I.K.; Mbaveng, A.T.; Awoussong, P.K.; Patnam, R.; Roy, R.; Ngadjui, B.T. Antibacterial and Antifungal Activities of the Crude Extract and Compounds from Dorstenia turbinata (Moraceae). S. Afr. J. Bot. 2009, 75, 256–261. [Google Scholar] [CrossRef] [Green Version]
  72. Basile, A.; Sorbo, S.; Spadaro, V.; Bruno, M.; Maggio, A.; Faraone, N.; Rosselli, S. Antimicrobial and Antioxidant Activities of Coumarins from the Roots of Ferulago campestris (Apiaceae). Molecules 2009, 14, 939–952. [Google Scholar] [CrossRef] [Green Version]
  73. Razavi, S.M.; Imanzadeh, G.; Jahed, F.S.; Zarrini, G. Pyranocoumarins from Zosima absinthifolia (Vent) Link Roots. Russ. J. Bioorg. Chem. 2013, 39, 215–217. [Google Scholar] [CrossRef]
  74. Aljubiri, S.M.; Mahmoud, K.; Mahgoub, S.A.; Almansour, A.I.; Shaker, K.H. Bioactive Compounds from Euphorbia schimperiana with Cytotoxic and Antibacterial Activities. S. Afr. J. Bot. 2021, 141, 357–366. [Google Scholar] [CrossRef]
  75. Hashim, I.; Onyari, J.M.; Omosa, L.K.; Maru, S.M.; Nchiozem-Ngnitedem, V.-A.; Karpoormath, R. Conglomeratin: A New Antibacterial Flavonol Derivative from Macaranga conglomerata Brenan (Euphorbiaceae). Nat. Prod. Res. 2022, 1–9. [Google Scholar] [CrossRef]
  76. Dongmo, A.J.N.; Ekom, S.E.; Tamokou, J.-D.; Tagousop, C.N.; Harakat, D.; Voutquenne-Nazabadioko, L.; Ngnokam, D. A New Stilbene Dimer and Other Chemical Constituents from Monanthotaxis littoralis with Their Antimicrobial Activities. Nat. Prod. Sci. 2020, 26, 9. [Google Scholar]
  77. Kakar, M.; Amin, M.U.; Alghamdi, S.; Sahibzada, M.U.K.; Ahmad, N.; Ullah, N. Antimicrobial, Cytotoxic, and Antioxidant Potential of a Novel Flavone “6,7,4′-Trimethyl Flavone” Isolated from Wulfenia amherstiana. Evid.-Based Complement. Altern. Med. 2020, 2020, e3903682. [Google Scholar] [CrossRef]
  78. AJ, Y.; MI, A.; AM, M.; AK, H.A. Bioactive (+)-Catechin-3′-O-Rhamnopyranoside from Neocarya macrophylla (Sabine) Prance (Chrysobalanaceae). Egypt. J. Basic Appl. Sci. 2019, 6, 124–136. [Google Scholar] [CrossRef] [Green Version]
  79. Ortega-Vidal, J.; Cobo, A.; Ortega-Morente, E.; Gálvez, A.; Alejo-Armijo, A.; Salido, S.; Altarejos, J. Antimicrobial and Antioxidant Activities of Flavonoids Isolated from Wood of Sweet Cherry Tree (Prunus avium L.). J. Wood Chem. Technol. 2021, 41, 104–117. [Google Scholar] [CrossRef]
  80. Shao, T.-M.; Liao, H.-X.; Li, X.-B.; Chen, G.-Y.; Song, X.-P.; Han, C.-R. A New Isoflavone from the Fruits of Ficus auriculata and Its Antibacterial Activity. Nat. Prod. Res. 2022, 36, 1191–1196. [Google Scholar] [CrossRef]
  81. Promchai, T.; Janhom, P.; Maneerat, W.; Rattanajak, R.; Kamchonwongpaisan, S.; Pyne, S.G.; Limtharakul, T. Antibacterial and Cytotoxic Activities of Phenolic Constituents from the Stem Extracts of Spatholobus parviflorus. Nat. Prod. Res. 2020, 34, 1394–1398. [Google Scholar] [CrossRef] [Green Version]
  82. Polbuppha, I.; Suthiphasilp, V.; Maneerat, T.; Charoensup, R.; Limtharakul, T.; Cheenpracha, S.; Pyne, S.G.; Laphookhieo, S. Macluracochinones A-E, Antimicrobial Flavonoids from Maclura cochinchinensis (Lour.) Corner. Phytochemistry 2021, 187, 112773. [Google Scholar] [CrossRef]
  83. Issarachot, P.; Sangkaew, W.; Sianglum, W.; Saeloh, D.; Limsuwan, S.; Voravuthikunchai, S.P.; Joycharat, N. α-Glucosidase Inhibitory, Antibacterial, and Antioxidant Activities of Natural Substances from the Wood of Derris reticulata Craib. Nat. Prod. Res. 2021, 35, 2858–2865. [Google Scholar] [CrossRef]
  84. Cho, J.-Y.; Sohn, M.-J.; Lee, J.; Kim, W.-G. Isolation and Identification of Pentagalloylglucose with Broad-Spectrum Antibacterial Activity from Rhus Trichocarpa miquel. Food Chem. 2010, 123, 501–506. [Google Scholar] [CrossRef]
  85. Gosset-Erard, C.; Zhao, M.; Lordel-Madeleine, S.; Ennahar, S. Identification of Punicalagin as the Bioactive Compound behind the Antimicrobial Activity of Pomegranate (Punica granatum L.) Peels. Food Chem. 2021, 352, 129396. [Google Scholar] [CrossRef]
  86. Shimozu, Y.; Kuroda, T.; Tsuchiya, T.; Hatano, T. Structures and Antibacterial Properties of Isorugosins H–J, Oligomeric Ellagitannins from Liquidambar formosana with Characteristic Bridging Groups between Sugar Moieties. J. Nat. Prod. 2017, 80, 2723–2733. [Google Scholar] [CrossRef]
  87. Araújo, A.R.; Araújo, A.C.; Reis, R.L.; Pires, R.A. Vescalagin and Castalagin Present Bactericidal Activity toward Methicillin-Resistant Bacteria. ACS Biomater. Sci. Eng. 2021, 7, 1022–1030. [Google Scholar] [CrossRef]
  88. Karioti, A.; Sokovic, M.; Ciric, A.; Koukoulitsa, C.; Bilia, A.R.; Skaltsa, H. Antimicrobial Properties of Quercus ilex L. Proanthocyanidin Dimers and Simple Phenolics: Evaluation of Their Synergistic Activity with Conventional Antimicrobials and Prediction of Their Pharmacokinetic Profile. J. Agric. Food Chem. 2011, 59, 6412–6422. [Google Scholar] [CrossRef]
  89. Kim, K.-H.; Yu, D.; Eom, S.-H.; Kim, H.-J.; Kim, D.-H.; Song, H.-S.; Kim, D.-M.; Kim, Y.-M. Fucofuroeckol-A from Edible Marine Alga Eisenia bicyclis to Restore Antifungal Activity of Fluconazole against Fluconazole-Resistant Candida albicans. J. Appl. Phycol. 2018, 30, 605–609. [Google Scholar] [CrossRef]
  90. Lee, D.-S.; Kang, M.-S.; Hwang, H.-J.; Eom, S.-H.; Yang, J.-Y.; Lee, M.-S.; Lee, W.-J.; Jeon, Y.-J.; Choi, J.-S.; Kim, Y.-M. Synergistic Effect between Dieckol from Ecklonia stolonifera and β-Lactams against Methicillin-Resistant Staphylococcus aureus. Biotechnol. Bioproc. E 2008, 13, 758–764. [Google Scholar] [CrossRef]
  91. Seukep, J.A.; Sandjo, L.P.; Ngadjui, B.T.; Kuete, V. Antibacterial and Antibiotic-Resistance Modifying Activity of the Extracts and Compounds from Nauclea pobeguinii against Gram-Negative Multi-Drug Resistant Phenotypes. BMC Complement Altern. Med. 2016, 16, 193. [Google Scholar] [CrossRef] [Green Version]
  92. Kato, E.; Tokunaga, Y.; Sakan, F. Stilbenoids Isolated from the Seeds of Melinjo (Gnetum gnemon L.) and Their Biological Activity. J. Agric. Food Chem. 2009, 57, 2544–2549. [Google Scholar] [CrossRef]
  93. Kumar, S.n.; Siji, J.v.; Rajasekharan, K.n.; Nambisan, B.; Mohandas, C. Bioactive Stilbenes from a Bacillus sp. N Strain Associated with a Novel Rhabditid Entomopathogenic Nematode. Lett. Appl. Microbiol. 2012, 54, 410–417. [Google Scholar] [CrossRef]
  94. Basset, C.; Rodrigues, A.M.S.; Eparvier, V.; Silva, M.R.R.; Lopes, N.P.; Sabatier, D.; Fonty, E.; Espindola, L.S.; Stien, D. Secondary Metabolites from Spirotropis longifolia (DC) Baill and Their Antifungal Activity against Human Pathogenic Fungi. Phytochemistry 2012, 74, 166–172. [Google Scholar] [CrossRef]
  95. Mattio, L.M.; Dallavalle, S.; Musso, L.; Filardi, R.; Franzetti, L.; Pellegrino, L.; D’Incecco, P.; Mora, D.; Pinto, A.; Arioli, S. Antimicrobial Activity of Resveratrol-Derived Monomers and Dimers against Foodborne Pathogens. Sci. Rep. 2019, 9, 19525. [Google Scholar] [CrossRef] [Green Version]
  96. Lu, H.-P.; Jia, Y.-N.; Peng, Y.-L.; Yu, Y.; Sun, S.-L.; Yue, M.-T.; Pan, M.-H.; Zeng, L.-S.; Xu, L. Oxyresveratrol, a Stilbene Compound from Morus alba L. Twig Extract Active Against Trichophyton rubrum. Phytother. Res. 2017, 31, 1842–1848. [Google Scholar] [CrossRef]
  97. de Bruijn, W.J.C.; Araya-Cloutier, C.; Bijlsma, J.; de Swart, A.; Sanders, M.G.; de Waard, P.; Gruppen, H.; Vincken, J.-P. Antibacterial Prenylated Stilbenoids from Peanut (Arachis hypogaea). Phytochem. Lett. 2018, 28, 13–18. [Google Scholar] [CrossRef]
  98. Li, J.; Chen, G.; Wu, H.; Webster, J.M. Identification of Two Pigments and a Hydroxystilbene Antibiotic from Photorhabdus luminescens. Appl. Environ. Microbiol. 1995, 61, 4329–4333. [Google Scholar] [CrossRef] [Green Version]
  99. Wu, J.; Li, B.; Xiao, W.; Hu, J.; Xie, J.; Yuan, J.; Wang, L. Longistylin A, a Natural Stilbene Isolated from the Leaves of Cajanus cajan, Exhibits Significant Anti-MRSA Activity. Int. J. Antimicrob. Agents 2020, 55, 105821. [Google Scholar] [CrossRef]
  100. Liu, P.; Li, X.-F.; Gao, J.-Y.; Liu, Y.; Hou, X.-W.; Yin, W.-P.; Deng, R.-X. Two New Resveratrol Trimers with Antibacterial Activities from Seed Cake of Paeonia rockii. Chem. Nat. Compd. 2017, 53, 51–55. [Google Scholar] [CrossRef]
  101. Wibowo, A.; Ahmat, N.; Hamzah, A.S.; Low, A.L.M.; Mohamad, S.A.S.; Khong, H.Y.; Sufian, A.S.; Manshoor, N.; Takayama, H. Malaysianol B, an Oligostilbenoid Derivative from Dryobalanops lanceolata. Fitoterapia 2012, 83, 1569–1575. [Google Scholar] [CrossRef]
  102. Peng, S.-C.; Cheng, C.-Y.; Sheu, F.; Su, C.-H. The Antimicrobial Activity of Heyneanol A Extracted from the Root of Taiwanese Wild Grape. J. Appl. Microbiol. 2008, 105, 485–491. [Google Scholar] [CrossRef]
  103. Kumarasamy, Y.; Nahar, L.; Cox, P.J.; Dinan, L.N.; Ferguson, C.A.; Finnie, D.A.; Jaspars, M.; Sarker, S.D. Biological Activity of Lignans from the Seeds of Centaurea scabiosa. Pharm. Biol. 2003, 41, 203–206. [Google Scholar] [CrossRef]
  104. Panagouleas, C.; Skaltsa, H.; Lazari, D.; Skaltsounis, A.-L.; Sokovic, M. Antifungal Activity of Secondary Metabolites of Centaurea raphanina ssp. Mixta, Growing Wild in Greece. Pharm. Biol. 2003, 41, 266–270. [Google Scholar] [CrossRef]
  105. Bajpai, V.K.; Shukla, S.; Paek, W.K.; Lim, J.; Kumar, P.; Kumar, P.; Na, M. Efficacy of (+)-Lariciresinol to Control Bacterial Growth of Staphylococcus aureus and Escherichia coli O157:H7. Front. Microbiol. 2017, 8, 804. [Google Scholar] [CrossRef] [Green Version]
  106. Hwang, B.; Cho, J.; Hwang, I.; Jin, H.-G.; Woo, E.-R.; Lee, D.G. Antifungal Activity of Lariciresinol Derived from Sambucus williamsii and Their Membrane-Active Mechanisms in Candida Albicans. Biochem. Biophys. Res. Commun. 2011, 410, 489–493. [Google Scholar] [CrossRef]
  107. Nazeam, J.A.; AL-Shareef, W.A.; Helmy, M.W.; El-Haddad, A.E. Bioassay-Guided Isolation of Potential Bioactive Constituents from Pomegranate Agrifood by-Product. Food Chem. 2020, 326, 126993. [Google Scholar] [CrossRef] [PubMed]
  108. Kaigongi, M.M.; Lukhoba, C.W.; Yaouba, S.; Makunga, N.P.; Githiomi, J.; Yenesew, A. In Vitro Antimicrobial and Antiproliferative Activities of the Root Bark Extract and Isolated Chemical Constituents of Zanthoxylum paracanthum Kokwaro (Rutaceae). Plants 2020, 9, 920. [Google Scholar] [CrossRef] [PubMed]
  109. Deng, R.; Zhou, S.; Yang, X.; Zhao, S.; Liu, P. Two New Furofuran Lignan Glycosides from Forsythia suspensa Leaves. Phytochem. Lett. 2021, 41, 34–37. [Google Scholar] [CrossRef]
  110. Zhou, H.; Ren, J.; Li, Z. Antibacterial Activity and Mechanism of Pinoresinol from Cinnamomum camphora Leaves against Food-Related Bacteria. Food Control 2017, 79, 192–199. [Google Scholar] [CrossRef]
  111. Hwang, B.; Lee, J.; Liu, Q.-H.; Woo, E.-R.; Lee, D.G. Antifungal Effect of (+)-Pinoresinol Isolated from Sambucus williamsii. Molecules 2010, 15, 3507–3516. [Google Scholar] [CrossRef]
  112. Bai, M.; Wu, S.-Y.; Zhang, W.-F.; Song, X.-P.; Han, C.-R.; Zheng, C.-J.; Chen, G.-Y. One New Lignan Derivative from the Fruiting Bodies of Ganoderma lipsiense. Nat. Prod. Res. 2019, 33, 2784–2788. [Google Scholar] [CrossRef]
  113. El-Gendy, M.M.A.; Hawas, U.W.; Jaspars, M. Novel Bioactive Metabolites from a Marine Derived Bacterium Nocardia sp. ALAA 2000. J. Antibiot. 2008, 61, 379–386. [Google Scholar] [CrossRef] [Green Version]
  114. Ding, J.-Y.; Yuan, C.-M.; Cao, M.-M.; Liu, W.-W.; Yu, C.; Zhang, H.-Y.; Zhang, Y.; Di, Y.-T.; He, H.-P.; Li, S.-L.; et al. Antimicrobial Constituents of the Mature Carpels of Manglietiastrum sinicum. J. Nat. Prod. 2014, 77, 1800–1805. [Google Scholar] [CrossRef]
  115. da Silva, R.E.; de Oliveira Silva Ribeiro, F.; de Carvalho, A.M.A.; Daboit, T.C.; Marinho-Filho, J.D.B.; Matos, T.S.; Pessoa, O.D.L.; de Souza de Almeida Leite, J.R.; de Araújo, A.R.; dos Santos Soares, M.J. Antimicrobial and Antibiofilm Activity of the Benzoquinone Oncocalyxone A. Microb. Pathog. 2020, 149, 104513. [Google Scholar] [CrossRef]
  116. Drewes, S.E.; Khan, F.; van Vuuren, S.F.; Viljoen, A.M. Simple 1,4-Benzoquinones with Antibacterial Activity from Stems and Leaves of Gunnera perpensa. Phytochemistry 2005, 66, 1812–1816. [Google Scholar] [CrossRef]
  117. Carcamo-Noriega, E.N.; Sathyamoorthi, S.; Banerjee, S.; Gnanamani, E.; Mendoza-Trujillo, M.; Mata-Espinosa, D.; Hernández-Pando, R.; Veytia-Bucheli, J.I.; Possani, L.D.; Zare, R.N. 1,4-Benzoquinone Antimicrobial Agents against Staphylococcus aureus and Mycobacterium tuberculosis Derived from Scorpion Venom. Proc. Natl. Acad. Sci. USA 2019, 116, 12642–12647. [Google Scholar] [CrossRef] [PubMed]
  118. Lana, E.J.L.; Carazza, F.; Takahashi, J.A. Antibacterial Evaluation of 1,4-Benzoquinone Derivatives. J. Agric. Food Chem. 2006, 54, 2053–2056. [Google Scholar] [CrossRef] [PubMed]
  119. Cesari, I.; Hoerlé, M.; Simoes-Pires, C.; Grisoli, P.; Queiroz, E.F.; Dacarro, C.; Marcourt, L.; Moundipa, P.F.; Carrupt, P.A.; Cuendet, M.; et al. Anti-Inflammatory, Antimicrobial and Antioxidant Activities of Diospyros bipindensis (Gürke) Extracts and Its Main Constituents. J. Ethnopharmacol. 2013, 146, 264–270. [Google Scholar] [CrossRef] [PubMed]
  120. 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] [CrossRef]
  121. Dzoyem, J.P.; Tangmouo, J.G.; Lontsi, D.; Etoa, F.X.; Lohoue, P.J. In Vitro Antifungal Activity of Extract and Plumbagin from the Stem Bark of Diospyros Crassiflora Hiern (Ebenaceae). Phytother. Res. 2007, 21, 671–674. [Google Scholar] [CrossRef]
  122. Adusei, E.B.A.; Adosraku, R.K.; Oppong-Kyekyeku, J.; Amengor, C.D.K.; Jibira, Y. Resistance Modulation Action, Time-Kill Kinetics Assay, and Inhibition of Biofilm Formation Effects of Plumbagin from Plumbago Zeylanica Linn. J. Trop. Med. 2019, 2019, e1250645. [Google Scholar] [CrossRef] [Green Version]
  123. Kaewbumrung, S.; Panichayupakaranant, P. Isolation of Three Antibacterial Naphthoquinones from Plumbago indica Roots and Development of a Validated Quantitative HPLC Analytical Method. Nat. Prod. Res. 2012, 26, 2020–2023. [Google Scholar] [CrossRef]
  124. Cho, S.-C.; Sultan, M.Z.; Moon, S.-S. Anti-Acne Activities of Pulsaquinone, Hydropulsaquinone, and Structurally Related 1, 4-Quinone Derivatives. Arch. Pharm. Res. 2009, 32, 489–494. [Google Scholar] [CrossRef]
  125. Yang, X.; Summerhurst, D.K.; Koval, S.F.; Ficker, C.; Smith, M.L.; Bernards, M.A. Isolation of an Antimicrobial Compound from Impatiens Balsamina L. Using Bioassay-Guided Fractionation. Phytother. Res. 2001, 15, 676–680. [Google Scholar] [CrossRef]
  126. Balachandran, C.; Al-Dhabi, N.A.; Duraipandiyan, V.; Ignacimuthu, S. Bluemomycin, a New Naphthoquinone Derivative from Streptomyces sp. with Antimicrobial and Cytotoxic Properties. Biotechnol. Lett. 2021, 43, 1005–1018. [Google Scholar] [CrossRef]
  127. Chen, W.-H.; Liu, W.-J.; Wang, Y.; Song, X.-P.; Chen, G.-Y. A New Naphthoquinone and Other Antibacterial Constituents from the Roots of Xanthium sibiricum. Nat. Prod. Res. 2015, 29, 739–744. [Google Scholar] [CrossRef] [PubMed]
  128. Kengne, I.C.; Feugap, L.D.T.; Njouendou, A.J.; Ngnokam, C.D.J.; Djamalladine, M.D.; Ngnokam, D.; Voutquenne-Nazabadioko, L.; Tamokou, J.-D.-D. Antibacterial, Antifungal and Antioxidant Activities of Whole Plant Chemical Constituents of Rumex abyssinicus. BMC Complement. Med. Ther. 2021, 21, 164. [Google Scholar] [CrossRef]
  129. Chukwujekwu, J.C.; Coombes, P.H.; Mulholland, D.A.; van Staden, J. Emodin, an Antibacterial Anthraquinone from the Roots of Cassia occidentalis. S. Afr. J. Bot. 2006, 72, 295–297. [Google Scholar] [CrossRef] [Green Version]
  130. Yang, S.-Q.; Li, X.-M.; Xu, G.-M.; Li, X.; An, C.-Y.; Wang, B.-G. Antibacterial Anthraquinone Derivatives Isolated from a Mangrove-Derived Endophytic Fungus Aspergillus nidulans by Ethanol Stress Strategy. J. Antibiot. 2018, 71, 778–784. [Google Scholar] [CrossRef] [PubMed]
  131. Balachandran, C.; Arun, Y.; Duraipandiyan, V.; Ignacimuthu, S.; Balakrishna, K.; Al-Dhabi, N.A. Antimicrobial and Cytotoxicity Properties of 2,3-Dihydroxy-9,10-Anthraquinone Isolated from Streptomyces galbus (ERINLG-127). Appl. Biochem. Biotechnol. 2014, 172, 3513–3528. [Google Scholar] [CrossRef]
  132. Rhea, J.; Craig Hopp, D.; Rabenstein, J.; Smith, C.; Lucas, S.; Romari, K.; Clarke, M.; Francis, L.; Irigoyen, M.; Luche, M.; et al. 5-Hydroxy Ericamycin, a New Anthraquinone with Potent Antimicrobial Activity. J. Antibiot. 2012, 65, 623–625. [Google Scholar] [CrossRef]
  133. Sivasothy, Y.; Sulaiman, S.F.; Ooi, K.L.; Ibrahim, H.; Awang, K. Antioxidant and Antibacterial Activities of Flavonoids and Curcuminoids from Zingiber spectabile Griff. Food Control 2013, 30, 714–720. [Google Scholar] [CrossRef]
  134. Mun, S.-H.; Joung, D.-K.; Kim, Y.-S.; Kang, O.-H.; Kim, S.-B.; Seo, Y.-S.; Kim, Y.-C.; Lee, D.-S.; Shin, D.-W.; Kweon, K.-T.; et al. Synergistic Antibacterial Effect of Curcumin against Methicillin-Resistant Staphylococcus aureus. Phytomedicine 2013, 20, 714–718. [Google Scholar] [CrossRef]
  135. Targhi, A.A.; Moammeri, A.; Jamshidifar, E.; Abbaspour, K.; Sadeghi, S.; Lamakani, L.; Akbarzadeh, I. Synergistic Effect of Curcumin-Cu and Curcumin-Ag Nanoparticle Loaded Niosome: Enhanced Antibacterial and Anti-Biofilm Activities. Bioorganic Chem. 2021, 115, 105116. [Google Scholar] [CrossRef]
  136. Adamczak, A.; Ożarowski, M.; Karpiński, T.M. Curcumin, a Natural Antimicrobial Agent with Strain-Specific Activity. Pharmaceuticals 2020, 13, 153. [Google Scholar] [CrossRef]
  137. Aslam, Z.; Roome, T.; Razzak, A.; Aslam, S.M.; Zaidi, M.B.; Kanwal, T.; Sikandar, B.; Bertino, M.F.; Rehman, K.; Shah, M.R. Investigation of Wound Healing Potential of Photo-Active Curcumin-ZnO-Nanoconjugates in Excisional Wound Model. Photodiagnosis Photodyn. Ther. 2022, 39, 102956. [Google Scholar] [CrossRef] [PubMed]
  138. Andrade, J.T.; Fantini de Figueiredo, G.; Cruz, L.F.; Eliza de Morais, S.; Souza, C.D.F.; Pinto, F.C.H.; Ferreira, J.M.S.; de Araújo, M.G.F. Efficacy of Curcumin in the Treatment of Experimental Vulvovaginal candidiasis. Rev. Iberoam. Micol. 2019, 36, 192–199. [Google Scholar] [CrossRef] [PubMed]
  139. Parvathy, K.S.; Negi, P.S.; Srinivas, P. Antioxidant, Antimutagenic and Antibacterial Activities of Curcumin-β-Diglucoside. Food Chem. 2009, 115, 265–271. [Google Scholar] [CrossRef]
  140. Hoult, J.R.S.; Payá, M. Pharmacological and Biochemical Actions of Simple Coumarins: Natural Products with Therapeutic Potential. Gen. Pharmacol. Vasc. Syst. 1996, 27, 713–722. [Google Scholar] [CrossRef]
  141. Jain, P.K.; Joshi, H. Coumarin: Chemical and Pharmacological Profile. J. Appl. Pharm. Sci. 2012, 236–240. [Google Scholar] [CrossRef]
  142. Costa, T.M.; Tavares, L.B.B.; de Oliveira, D. Fungi as a Source of Natural Coumarins Production. Appl. Microbiol. Biotechnol. 2016, 100, 6571–6584. [Google Scholar] [CrossRef]
  143. Bor, T.; Aljaloud, S.O.; Gyawali, R.; Ibrahim, S.A. Chapter 26—Antimicrobials from Herbs, Spices, and Plants. In Fruits, Vegetables, and Herbs; Watson, R.R., Preedy, V.R., Eds.; Academic Press: London, UK, 2016; pp. 551–578. ISBN 978-0-12-802972-5. [Google Scholar]
  144. Carneiro, A.; Matos, M.J.; Uriarte, E.; Santana, L. Trending Topics on Coumarin and Its Derivatives in 2020. Molecules 2021, 26, 501. [Google Scholar] [CrossRef]
  145. Wu, Y.; Xu, J.; Liu, Y.; Zeng, Y.; Wu, G. A Review on Anti-Tumor Mechanisms of Coumarins. Front. Oncol. 2020, 10, 2720. [Google Scholar]
  146. Lacy, A. Studies on Coumarins and Coumarin-Related Compounds to Determine Their Therapeutic Role in the Treatment of Cancer. CPD 2004, 10, 3797–3811. [Google Scholar] [CrossRef] [Green Version]
  147. Steffensky, M.; Mühlenweg, A.; Wang, Z.-X.; Li, S.-M.; Heide, L. Identification of the Novobiocin Biosynthetic Gene Cluster of Streptomyces spheroides NCIB 11891. Antimicrob. Agents Chemother. 2000, 44, 1214–1222. [Google Scholar] [CrossRef] [Green Version]
  148. Eustáquio, A.S.; Gust, B.; Luft, T.; Li, S.-M.; Chater, K.F.; Heide, L. Clorobiocin Biosynthesis in Streptomyces: Identification of the Halogenase and Generation of Structural Analogs. Chem. Biol. 2003, 10, 279–288. [Google Scholar] [CrossRef] [Green Version]
  149. Wang, Z.-X.; Li, S.-M.; Heide, L. Identification of the Coumermycin A1 Biosynthetic Gene Cluster of Streptomyces rishiriensis DSM 40489. Antimicrob. Agents Chemother. 2000, 44, 3040–3048. [Google Scholar] [CrossRef]
  150. Fournier, B.; Hooper, D.C. Mutations in Topoisomerase IV and DNA Gyrase of Staphylococcus aureus: Novel Pleiotropic Effects on Quinolone and Coumarin Activity. Antimicrob. Agents Chemother. 1998, 42, 121–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Confreres, A.; Maxwell, A. GyrB Mutations Which Confer Coumarin Resistance Also Affect DNA Supercoiling and ATP Hydrolysis by Escherichia coli DNA Gyrase. Mol. Microbiol. 1992, 6, 1617–1624. [Google Scholar] [CrossRef] [PubMed]
  152. Maxwell, A. The Interaction between Coumarin Drugs and DNA Gyrase. Mol. Microbiol. 1993, 9, 681–686. [Google Scholar] [CrossRef]
  153. Stefanachi, A.; Leonetti, F.; Pisani, L.; Catto, M.; Carotti, A. Coumarin: A Natural, Privileged and Versatile Scaffold for Bioactive Compounds. Molecules 2018, 23, 250. [Google Scholar] [CrossRef] [Green Version]
  154. Finn, G.J.; Creaven, B.; Egan, D.A. Study of the in Vitro Cytotoxic Potential of Natural and Synthetic Coumarin Derivatives Using Human Normal and Neoplastic Skin Cell Lines. Melanoma Res. 2001, 11, 461–467. [Google Scholar] [CrossRef]
  155. KhanYusufzai, S.; Osman, H.; Khan, M.S.; Mohamad, S.; Sulaiman, O.; Parumasivam, T.; Gansau, J.A.; Johansah, N. Noviany Design, Characterization, in Vitro Antibacterial, Antitubercular Evaluation and Structure–Activity Relationships of New Hydrazinyl Thiazolyl Coumarin Derivatives. Med. Chem. Res. 2017, 26, 1139–1148. [Google Scholar] [CrossRef] [Green Version]
  156. Qin, H.-L.; Zhang, Z.-W.; Ravindar, L.; Rakesh, K.P. Antibacterial Activities with the Structure-Activity Relationship of Coumarin Derivatives. Eur. J. Med. Chem. 2020, 207, 112832. [Google Scholar] [CrossRef]
  157. Ranjan Sahoo, C.; Sahoo, J.; Mahapatra, M.; Lenka, D.; Kumar Sahu, P.; Dehury, B.; Nath Padhy, R.; Kumar Paidesetty, S. Coumarin Derivatives as Promising Antibacterial Agent(s). Arab. J. Chem. 2021, 14, 102922. [Google Scholar] [CrossRef]
  158. Znati, M.; Zardi-Bergaoui, A.; Daami-Remadi, M.; Ben Jannet, H. Semi-Synthesis, Antibacterial, Anticholinesterase Activities, and Drug Likeness Properties of New Analogues of Coumarins Isolated from Ferula lutea (Poir.) Maire. Chem. Afr. 2020, 3, 635–645. [Google Scholar] [CrossRef]
  159. Mamidala, S.; Peddi, S.R.; Aravilli, R.K.; Jilloju, P.C.; Manga, V.; Vedula, R.R. Microwave Irradiated One Pot, Three Component Synthesis of a New Series of Hybrid Coumarin Based Thiazoles: Antibacterial Evaluation and Molecular Docking Studies. J. Mol. Struct. 2021, 1225, 129114. [Google Scholar] [CrossRef]
  160. Feng, D.; Zhang, A.; Yang, Y.; Yang, P. Coumarin-Containing Hybrids and Their Antibacterial Activities. Arch. Der Pharm. 2020, 353, 1900380. [Google Scholar] [CrossRef]
  161. Song, P.-P.; Zhao, J.; Liu, Z.-L.; Duan, Y.-B.; Hou, Y.-P.; Zhao, C.-Q.; Wu, M.; Wei, M.; Wang, N.-H.; Lv, Y.; et al. Evaluation of Antifungal Activities and Structure–Activity Relationships of Coumarin Derivatives. Pest Manag. Sci. 2017, 73, 94–101. [Google Scholar] [CrossRef]
  162. Nabavi, S.M.; Šamec, D.; Tomczyk, M.; Milella, L.; Russo, D.; Habtemariam, S.; Suntar, I.; Rastrelli, L.; Daglia, M.; Xiao, J.; et al. Flavonoid Biosynthetic Pathways in Plants: Versatile Targets for Metabolic Engineering. Biotechnol. Adv. 2020, 38, 107316. [Google Scholar] [CrossRef] [PubMed]
  163. Miadoková, E. Isoflavonoids—An Overview of Their Biological Activities and Potential Health Benefits. Interdiscip. Toxicol. 2009, 2, 211–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Niaz, K.; Khan, F. Chapter 3—Analysis of Polyphenolics. In Recent Advances in Natural Products Analysis; Sanches Silva, A., Nabavi, S.F., Saeedi, M., Nabavi, S.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 39–197. ISBN 978-0-12-816455-6. [Google Scholar]
  165. Mukherjee, P.K. Chapter 4—Qualitative Analysis for Evaluation of Herbal Drugs. In Quality Control and Evaluation of Herbal Drugs; Mukherjee, P.K., Ed.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 79–149. ISBN 978-0-12-813374-3. [Google Scholar]
  166. Yan, S.; Xie, M.; Wang, Y.; Xiao, Q.; Ding, N.; Li, Y. Semi-Synthesis of a Series Natural Flavonoids and Flavonoid Glycosides from Scutellarin. Tetrahedron 2020, 76, 130950. [Google Scholar] [CrossRef]
  167. Orsat, V.; Routray, W. Chapter 8—Microwave-Assisted Extraction of Flavonoids. In Water Extraction of Bioactive Compounds; Dominguez González, H., González Muñoz, M.J., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 221–244. ISBN 978-0-12-809380-1. [Google Scholar]
  168. Shen, N.; Wang, T.; Gan, Q.; Liu, S.; Wang, L.; Jin, B. Plant Flavonoids: Classification, Distribution, Biosynthesis, and Antioxidant Activity. Food Chem. 2022, 383, 132531. [Google Scholar] [CrossRef]
  169. Dixon, R.A.; Steele, C.L. Flavonoids and Isoflavonoids—A Gold Mine for Metabolic Engineering. Trends Plant Sci. 1999, 4, 394–400. [Google Scholar] [CrossRef]
  170. Wu, S.-C.; Liu, F.; Zhu, K.; Shen, J.-Z. Natural Products That Target Virulence Factors in Antibiotic-Resistant Staphylococcus aureus. J. Agric. Food Chem. 2019, 67, 13195–13211. [Google Scholar] [CrossRef]
  171. Cushnie, T.P.T.; Lamb, A.J. Antimicrobial Activity of Flavonoids. Int. J. Antimicrob. Agents 2005, 26, 343–356. [Google Scholar] [CrossRef] [PubMed]
  172. Tagousop, C.N.; Tamokou, J.-D.; Ekom, S.E.; Ngnokam, D.; Voutquenne-Nazabadioko, L. Antimicrobial Activities of Flavonoid Glycosides from Graptophyllum grandulosum and Their Mechanism of Antibacterial Action. BMC Complement. Altern. Med. 2018, 18, 252. [Google Scholar] [CrossRef] [PubMed]
  173. Geethalakshmi, R.; Sundaramurthi, J.C.; Sarada, D.V.L. Antibacterial Activity of Flavonoid Isolated from Trianthema decandra against Pseudomonas aeruginosa and Molecular Docking Study of FabZ. Microb. Pathog. 2018, 121, 87–92. [Google Scholar] [CrossRef] [PubMed]
  174. Abreu, A.C.; Serra, S.C.; Borges, A.; Saavedra, M.J.; Mcbain, A.J.; Salgado, A.J.; Simões, M. Combinatorial Activity of Flavonoids with Antibiotics Against Drug-Resistant Staphylococcus aureus. Microb. Drug Resist. 2015, 21, 600–609. [Google Scholar] [CrossRef]
  175. Sathiya Deepika, M.; Thangam, R.; Sakthidhasan, P.; Arun, S.; Sivasubramanian, S.; Thirumurugan, R. Combined Effect of a Natural Flavonoid Rutin from Citrus Sinensis and Conventional Antibiotic Gentamicin on Pseudomonas aeruginosa Biofilm Formation. Food Control 2018, 90, 282–294. [Google Scholar] [CrossRef]
  176. Fraga-Corral, M.; Otero, P.; Echave, J.; Garcia-Oliveira, P.; Carpena, M.; Jarboui, A.; Nuñez-Estevez, B.; Simal-Gandara, J.; Prieto, M.A. By-Products of Agri-Food Industry as Tannin-Rich Sources: A Review of Tannins’ Biological Activities and Their Potential for Valorization. Foods 2021, 10, 137. [Google Scholar] [CrossRef]
  177. Dhawale, P.V.; Vineeth, S.K.; Gadhave, R.V.; Mj, J.F.; Supekar, M.V.; Thakur, V.K.; Raghavan, P. Tannin as a Renewable Raw Material for Adhesive Applications: A Review. Mater. Adv. 2022, 3, 3365–3388. [Google Scholar] [CrossRef]
  178. Yoshida, T.; Yoshimura, M.; Amakura, Y. Chemical and Biological Significance of Oenothein B and Related Ellagitannin oligomers with Macrocyclic Structure. Molecules 2018, 23, 552. [Google Scholar] [CrossRef] [Green Version]
  179. Khanbabaee, K.; Ree, T. van Tannins: Classification and Definition. Nat. Prod. Rep. 2001, 18, 641–649. [Google Scholar] [CrossRef]
  180. Rira, M.; Morgavi, D.P.; Popova, M.; Maxin, G.; Doreau, M. Microbial Colonisation of Tannin-Rich Tropical Plants: Interplay between Degradability, Methane Production and Tannin Disappearance in the Rumen. Animal 2022, 16, 100589. [Google Scholar] [CrossRef]
  181. Yoshida, T.; Hatano, T.; Ito, H. Chapter Seven—High Molecular Weight Plant Poplyphenols (Tannins): Prospective Functions. In Recent Advances in Phytochemistry; Romeo, J.T., Ed.; Chemical Ecology and Phytochemistry of Forest Ecosystems; Elsevier: Amsterdam, The Netherlands, 2005; Volume 39, pp. 163–190. [Google Scholar]
  182. Frazier, R.A.; Deaville, E.R.; Green, R.J.; Stringano, E.; Willoughby, I.; Plant, J.; Mueller-Harvey, I. Interactions of Tea Tannins and Condensed Tannins with Proteins. J. Pharm. Biomed. Anal. 2010, 51, 490–495. [Google Scholar] [CrossRef] [PubMed]
  183. Yoshida, T.; Ito, H.; Yoshimura, M.; Miyashita, K.; Hatano, T. C-Glucosidic Ellagitannin Oligomers from Melaleuca Squarrosa Donn Ex Sm., Myrtaceae. Phytochemistry 2008, 69, 3070–3079. [Google Scholar] [CrossRef] [PubMed]
  184. Singh, I.P.; Sidana, J. 5—Phlorotannins. In Functional Ingredients from Algae for Foods and Nutraceuticals; Domínguez, H., Ed.; Woodhead Publishing Series in Food Science, Technology and Nutrition; Woodhead Publishing: Sawston, Cambridge, UK, 2013; pp. 181–204. ISBN 978-0-85709-512-1. [Google Scholar]
  185. Maisetta, G.; Batoni, G.; Caboni, P.; Esin, S.; Rinaldi, A.C.; Zucca, P. Tannin Profile, Antioxidant Properties, and Antimicrobial Activity of Extracts from Two Mediterranean Species of Parasitic Plant Cytinus. BMC Complement. Altern. Med. 2019, 19, 82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Al-Harbi, R.; Shaaban, M.; Al-wegaisi, R.; Moharram, F.; El-Rahman, O.A.; El-Messery, S. Antimicrobial Activity and Molecular Docking of Tannins from Pimenta dioica. Lett. Drug Des. Discov. 2018, 15, 508–515. [Google Scholar] [CrossRef]
  187. Liu, M.; Yang, K.; Wang, J.; Zhang, J.; Qi, Y.; Wei, X.; Fan, M. Young Astringent Persimmon tannin Inhibits Methicillin-Resistant Staphylococcus Aureus Isolated from Pork. LWT 2019, 100, 48–55. [Google Scholar] [CrossRef]
  188. de Freitas, A.L.D.; Kaplum, V.; Rossi, D.C.P.; da Silva, L.B.R.; Melhem, M.; de Souza Carvalho Melhem, M.; Taborda, C.P.; de Mello, J.C.P.; Nakamura, C.V.; Ishida, K. Proanthocyanidin Polymeric Tannins from Stryphnodendron adstringens Are Effective against Candida spp. Isolates and for Vaginal Candidiasis Treatment. J. Ethnopharmacol. 2018, 216, 184–190. [Google Scholar] [CrossRef]
  189. Akiyama, H.; Fujii, K.; Yamasaki, O.; Oono, T.; Iwatsuki, K. Antibacterial Action of Several Tannins against Staphylococcus aureus. J. Antimicrob. Chemother. 2001, 48, 487–491. [Google Scholar] [CrossRef] [Green Version]
  190. Huang, Q.; Liu, X.; Zhao, G.; Hu, T.; Wang, Y. Potential and Challenges of Tannins as an Alternative to In-Feed Antibiotics for Farm Animal Production. Anim. Nutr. 2018, 4, 137–150. [Google Scholar] [CrossRef]
  191. Scalbert, A. Antimicrobial Properties of Tannins. Phytochemistry 1991, 30, 3875–3883. [Google Scholar] [CrossRef]
  192. Eom, S.-H.; Lee, D.-S.; Jung, Y.-J.; Park, J.-H.; Choi, J.-I.; Yim, M.-J.; Jeon, J.-M.; Kim, H.-W.; Son, K.-T.; Je, J.-Y.; et al. The Mechanism of Antibacterial Activity of Phlorofucofuroeckol-A against Methicillin-Resistant Staphylococcus aureus. Appl. Microbiol. Biotechnol. 2014, 98, 9795–9804. [Google Scholar] [CrossRef]
  193. Liu, M.; Feng, M.; Yang, K.; Cao, Y.; Zhang, J.; Xu, J.; Hernández, S.H.; Wei, X.; Fan, M. Transcriptomic and Metabolomic Analyses Reveal Antibacterial Mechanism of Astringent Persimmon Tannin against Methicillin-Resistant Staphylococcus aureus Isolated from Pork. Food Chem. 2020, 309, 125692. [Google Scholar] [CrossRef] [PubMed]
  194. De Filippis, B.; Ammazzalorso, A.; Amoroso, R.; Giampietro, L. Stilbene Derivatives as New Perspective in Antifungal Medicinal Chemistry. Drug Dev. Res. 2019, 80, 285–293. [Google Scholar] [CrossRef] [PubMed]
  195. Joyce, S.A.; Brachmann, A.O.; Glazer, I.; Lango, L.; Schwär, G.; Clarke, D.J.; Bode, H.B. Bacterial Biosynthesis of a Multipotent Stilbene. Angew. Chem. Int. Ed. 2008, 47, 1942–1945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Jayatilake, G.S.; Baker, B.J.; McClintock, J.B. Isolation and Identification of a Stilbene Derivative from the Antarctic Sponge Kirkpatrickia variolosa. J. Nat. Prod. 1995, 58, 1958–1960. [Google Scholar] [CrossRef]
  197. Błaszczyk, A.; Sady, S.; Sielicka, M. The Stilbene Profile in Edible Berries. Phytochem. Rev. 2019, 18, 37–67. [Google Scholar] [CrossRef] [Green Version]
  198. Morabito, G.; Miglio, C.; Peluso, I.; Serafini, M. Chapter 85—Fruit Polyphenols and Postprandial Inflammatory Stress. In Polyphenols in Human Health and Disease; Watson, R.R., Preedy, V.R., Zibadi, S., Eds.; Academic Press: San Diego, CA, USA, 2014; pp. 1107–1126. ISBN 978-0-12-398456-2. [Google Scholar]
  199. Garbicz, D.; Tobiasz, P.; Borys, F.; Pilżys, T.; Marcinkowski, M.; Poterała, M.; Grzesiuk, E.; Krawczyk, H. The Stilbene and Dibenzo[b,f]Oxepine Derivatives as Anticancer Compounds. Biomed. Pharmacother. 2020, 123, 109781. [Google Scholar] [CrossRef]
  200. Lv, L.; Gu, X.; Tang, J.; Ho, C.-T. Antioxidant Activity of Stilbene Glycoside from Polygonum multiflorum Thunb in Vivo. Food Chem. 2007, 104, 1678–1681. [Google Scholar] [CrossRef]
  201. Boccellino, M.; Donniacuo, M.; Bruno, F.; Rinaldi, B.; Quagliuolo, L.; Ambruosi, M.; Pace, S.; De Rosa, M.; Olgaç, A.; Banoglu, E.; et al. Protective Effect of Piceatannol and Bioactive Stilbene Derivatives against Hypoxia-Induced Toxicity in H9c2 Cardiomyocytes and Structural Elucidation as 5-LOX Inhibitors. Eur. J. Med. Chem. 2019, 180, 637–647. [Google Scholar] [CrossRef]
  202. Sham, T.-T.; Li, M.-H.; Chan, C.-O.; Zhang, H.; Chan, S.-W.; Mok, D.K.-W. Cholesterol-Lowering Effects of Piceatannol, a Stilbene from Wine, Using Untargeted Metabolomics. J. Funct. Foods 2017, 28, 127–137. [Google Scholar] [CrossRef]
  203. Lai, X.; Pei, Q.; Song, X.; Zhou, X.; Yin, Z.; Jia, R.; Zou, Y.; Li, L.; Yue, G.; Liang, X.; et al. The Enhancement of Immune Function and Activation of NF-ΚB by Resveratrol-Treatment in Immunosuppressive Mice. Int. Immunopharmacol. 2016, 33, 42–47. [Google Scholar] [CrossRef]
  204. Lee, W.; Lee, D.G. Resveratrol Induces Membrane and DNA Disruption via Pro-Oxidant Activity against Salmonella typhimurium. Biochem. Biophys. Res. Commun. 2017, 489, 228–234. [Google Scholar] [CrossRef] [PubMed]
  205. Yang, S.-C.; Tseng, C.-H.; Wang, P.-W.; Lu, P.-L.; Weng, Y.-H.; Yen, F.-L.; Fang, J.-Y. Pterostilbene, a Methoxylated Resveratrol Derivative, Efficiently Eradicates Planktonic, Biofilm, and Intracellular MRSA by Topical Application. Front. Microbiol. 2017, 8, 1103. [Google Scholar] [CrossRef] [PubMed]
  206. Ren, X.; An, P.; Zhai, X.; Wang, S.; Kong, Q. The Antibacterial Mechanism of Pterostilbene Derived from Xinjiang Wine Grape: A Novel Apoptosis Inducer in Staphyloccocus aureus and Escherichia coli. LWT 2019, 101, 100–106. [Google Scholar] [CrossRef]
  207. Joung, D.-K.; Mun, S.-H.; Choi, S.-H.; Kang, O.-H.; Kim, S.-B.; Lee, Y.-S.; Zhou, T.; Kong, R.; Choi, J.-G.; Shin, D.-W.; et al. Antibacterial Activity of Oxyresveratrol against Methicillin-resistant Staphylococcus aureus and Its Mechanism. Exp. Ther. Med. 2016, 12, 1579–1584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Singh, D.; Mendonsa, R.; Koli, M.; Subramanian, M.; Nayak, S.K. Antibacterial Activity of Resveratrol Structural Analogues: A Mechanistic Evaluation of the Structure-Activity Relationship. Toxicol. Appl. Pharmacol. 2019, 367, 23–32. [Google Scholar] [CrossRef]
  209. Rivière, C.; Pawlus, A.D.; Mérillon, J.-M. Natural Stilbenoids: Distribution in the Plant Kingdom and Chemotaxonomic Interest in Vitaceae. Nat. Prod. Rep. 2012, 29, 1317. [Google Scholar] [CrossRef]
  210. Yang, D.-S.; Li, Z.-L.; Wang, X.; Yan, H.; Yang, Y.-P.; Luo, H.-R.; Liu, K.-C.; Xiao, W.-L.; Li, X.-L. Denticulatains A and B: Unique Stilbene–Diterpene Heterodimers from Macaranga denticulata. RSC Adv. 2015, 5, 13886–13890. [Google Scholar] [CrossRef]
  211. Shen, T.; Wang, X.-N.; Lou, H.-X. Natural Stilbenes: An Overview. Nat. Prod. Rep. 2009, 26, 916–935. [Google Scholar] [CrossRef]
  212. Wu, G.-Y.; Zhang, X.; Guo, X.-Y.; Huo, L.-Q.; Liu, H.-X.; Shen, X.-L.; Qiu, S.-X.; Hu, Y.-J.; Tan, H.-B. Prenylated Stilbenes and Flavonoids from the Leaves of Cajanus cajan. Chin. J. Nat. Med. 2019, 17, 381–386. [Google Scholar] [CrossRef]
  213. Cannatelli, A.; Principato, S.; Colavecchio, O.L.; Pallecchi, L.; Rossolini, G.M. Synergistic Activity of Colistin in Combination with Resveratrol Against Colistin-Resistant Gram-Negative Pathogens. Front. Microbiol. 2018, 9, 1808. [Google Scholar] [CrossRef]
  214. Singh, D.; Chauhan, N.; Koli, M.; Nayak, S.K.; Subramanian, M. Dimer Stilbene, a Resveratrol Analogue Exhibits Synergy with Antibiotics That Target Protein Synthesis in Eradicating Staphylococcus aureus Infection. Biochimie 2022, 201, 128–138. [Google Scholar] [CrossRef] [PubMed]
  215. Liu, Y.; Zhou, J.; Qu, Y.; Yang, X.; Shi, G.; Wang, X.; Hong, Y.; Drlica, K.; Zhao, X. Resveratrol Antagonizes Antimicrobial Lethality and Stimulates Recovery of Bacterial Mutants. PLoS ONE 2016, 11, e0153023. [Google Scholar] [CrossRef] [PubMed]
  216. Basri, D.F.; Xian, L.W.; Abdul Shukor, N.I.; Latip, J. Bacteriostatic Antimicrobial Combination: Antagonistic Interaction between Epsilon-Viniferin and Vancomycin against Methicillin-Resistant Staphylococcus aureus. BioMed Res. Int. 2014, 2014, e461756. [Google Scholar] [CrossRef] [Green Version]
  217. Sheng, J.-Y.; Chen, T.-T.; Tan, X.-J.; Chen, T.; Jia, A.-Q. The Quorum-Sensing Inhibiting Effects of Stilbenoids and Their Potential Structure–Activity Relationship. Bioorganic Med. Chem. Lett. 2015, 25, 5217–5220. [Google Scholar] [CrossRef] [PubMed]
  218. Karki, S.S.; Bhutle, S.R.; Pedgaonkar, G.S.; Zubaidha, P.K.; Shaikh, R.M.; Rajput, C.G.; Shendarkar, G.S. Synthesis and Biological Evaluation of Some Stilbene-Based Analogues. Med. Chem. Res. 2011, 20, 1158–1163. [Google Scholar] [CrossRef]
  219. Zhang, Z.; Zhu, Q.; Huang, C.; Yang, M.; Li, J.; Chen, Y.; Yang, B.; Zhao, X. Comparative Cytotoxicity of Halogenated Aromatic DBPs and Implications of the Corresponding Developed QSAR Model to Toxicity Mechanisms of Those DBPs: Binding Interactions between Aromatic DBPs and Catalase Play an Important Role. Water Res. 2020, 170, 115283. [Google Scholar] [CrossRef]
  220. Saleem, M.; Kim, H.J.; Ali, M.S.; Lee, Y.S. An Update on Bioactive Plant Lignans. Nat. Prod. Rep. 2005, 22, 696–716. [Google Scholar] [CrossRef]
  221. Ward, R.S. Recent Advances in the Chemistry of Lignans. In Studies in Natural Products Chemistry; Elsevier: Amsterdam, The Netherlands, 2000; Volume 24, pp. 739–798. ISBN 978-0-444-50643-6. [Google Scholar]
  222. Wang, L.-X.; Wang, H.-L.; Huang, J.; Chu, T.-Z.; Peng, C.; Zhang, H.; Chen, H.-L.; Xiong, Y.-A.; Tan, Y.-Z. Review of Lignans from 2019 to 2021: Newly Reported Compounds, Diverse Activities, Structure-Activity Relationships and Clinical Applications. Phytochemistry 2022, 202, 113326. [Google Scholar] [CrossRef]
  223. Eklund, P.; Raitanen, J.-E. 9-Norlignans: Occurrence, Properties and Their Semisynthetic Preparation from Hydroxymatairesinol. Molecules 2019, 24, 220. [Google Scholar] [CrossRef] [Green Version]
  224. Moss, G.P. Nomenclature of Lignans and Neolignans (IUPAC Recommendations 2000). Pure Appl. Chem. 2000, 72, 1493–1523. [Google Scholar] [CrossRef]
  225. Wcislo, G.; Szarlej-Wcislo, K. Chapter 8—Colorectal Cancer Prevention by Wheat Consumption: A Three-Valued Logic—True, False, or Otherwise? In Wheat and Rice in Disease Prevention and Health; Watson, R.R., Preedy, V.R., Zibadi, S., Eds.; Academic Press: San Diego, CA, USA, 2014; pp. 91–111. ISBN 978-0-12-401716-0. [Google Scholar]
  226. Aldred, E.M.; Buck, C.; Vall, K. (Eds.) Chapter 21—Phenols. In Pharmacology; Churchill Livingstone: Edinburgh, UK, 2009; pp. 149–166. ISBN 978-0-443-06898-0. [Google Scholar]
  227. Yoder, S.C.; Lancaster, S.M.; Hullar, M.A.J.; Lampe, J.W. Chapter 7—Gut Microbial Metabolism of Plant Lignans: Influence on Human Health. In Diet-Microbe Interactions in the Gut; Tuohy, K., Del Rio, D., Eds.; Academic Press: San Diego, CA, USA, 2015; pp. 103–117. ISBN 978-0-12-407825-3. [Google Scholar]
  228. Pathak, S.; Kesavan, P.; Banerjee, A.; Banerjee, A.; Celep, G.S.; Bissi, L.; Marotta, F. Chapter 25—Metabolism of Dietary Polyphenols by Human Gut Microbiota and Their Health Benefits. In Polyphenols: Mechanisms of Action in Human Health and Disease, 2nd ed.; Watson, R.R., Preedy, V.R., Zibadi, S., Eds.; Academic Press: London, UK, 2018; pp. 347–359. ISBN 978-0-12-813006-3. [Google Scholar]
  229. Wang, L.-Q. Mammalian Phytoestrogens: Enterodiol and Enterolactone. J. Chromatogr. B 2002, 777, 289–309. [Google Scholar] [CrossRef]
  230. Desmawati, D.; Sulastri, D. Phytoestrogens and Their Health Effect. Open Access Maced. J. Med. Sci. 2019, 7, 495–499. [Google Scholar] [CrossRef] [PubMed]
  231. Mahendra Kumar, C.; Singh, S.A. Bioactive Lignans from Sesame (Sesamum indicum L.): Evaluation of Their Antioxidant and Antibacterial Effects for Food Applications. J. Food Sci. Technol. 2015, 52, 2934–2941. [Google Scholar] [CrossRef] [Green Version]
  232. Mousa, W.; Raizada, M. The Diversity of Anti-Microbial Secondary Metabolites Produced by Fungal Endophytes: An Interdisciplinary Perspective. Front. Microbiol. 2013, 4, 65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Hui, Y.H.; Chang, C.J.; McLaughlin, J.L.; Powell, R.G. Justicidin B, a Bioactive Trace Lignan from the Seeds of Sesbania drummondii. J. Nat. Prod. 1986, 49, 1175–1176. [Google Scholar] [CrossRef]
  234. Rao, Y.K.; Fang, S.-H.; Tzeng, Y.-M. Anti-Inflammatory Activities of Constituents Isolated from Phyllanthus polyphyllus. J. Ethnopharmacol. 2006, 103, 181–186. [Google Scholar] [CrossRef] [PubMed]
  235. Vasilev, N.; Elfahmi; Bos, R.; Kayser, O.; Momekov, G.; Konstantinov, S.; Ionkova, I. Production of Justicidin B, a Cytotoxic Arylnaphthalene Lignan from Genetically Transformed Root Cultures of Linum leonii. J. Nat. Prod. 2006, 69, 1014–1017. [Google Scholar] [CrossRef]
  236. Chen, C.-C.; Hsin, W.-C.; Ko, F.-N.; Huang, Y.-L.; Ou, J.-C.; Teng, C.-M. Antiplatelet Arylnaphthalide lignans from Justicia procumbens. J. Nat. Prod. 1996, 59, 1149–1150. [Google Scholar] [CrossRef]
  237. Vargas-Arispuro, I.; Reyes-Báez, R.; Rivera-Castañeda, G.; Martínez-Téllez, M.A.; Rivero-Espejel, I. Antifungal Lignans from the Creosotebush (Larrea Tridentata). Ind. Crops Prod. 2005, 22, 101–107. [Google Scholar] [CrossRef]
  238. Li, K.-M.; Dong, X.; Ma, Y.-N.; Wu, Z.-H.; Yan, Y.-M.; Cheng, Y.-X. Antifungal Coumarins and Lignans from Artemisia annua. Fitoterapia 2019, 134, 323–328. [Google Scholar] [CrossRef]
  239. Akiyama, K.; Maruyama, M.; Yamauchi, S.; Nakashima, Y.; Nakato, T.; Tago, R.; Sugahara, T.; Kishida, T.; Koba, Y. Antimicrobiological Activity of Lignan: Effect of Benzylic Oxygen and Stereochemistry of 2,3-Dibenzyl-4-Butanolide and 3,4-Dibenzyltetrahydrofuran Lignans on Activity. Biosci. Biotechnol. Biochem. 2007, 71, 1745–1751. [Google Scholar] [CrossRef] [PubMed]
  240. Tago, R.; Yamauchi, S.; Maruyama, M.; Akiyama, K.; Sugahara, T.; Kishida, T.; Koba, Y. Structure-Antibacterial Activity Relationship for 9- O,9′- O -Demethyl (+)-Virgatusin. Biosci. Biotechnol. Biochem. 2008, 72, 1032–1037. [Google Scholar] [CrossRef] [PubMed]
  241. Akiyama, K.; Yamauchi, S.; Nakato, T.; Maruyama, M.; Sugahara, T.; Kishida, T. Antifungal Activity of Tetra-Substituted Tetrahydrofuran lignan, (−)-Virgatusin, and Its Structure-Activity Relationship. Biosci. Biotechnol. Biochem. 2007, 71, 1028–1035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Ouellette, R.J.; Rawn, J.D. 25 - Aryl Halides, Phenols, and Anilines. In Organic Chemistry, 2nd ed.; Ouellette, R.J., Rawn, J.D., Eds.; Academic Press: Cambridge, MA, USA, 2018; pp. 801–828. ISBN 978-0-12-812838-1. [Google Scholar]
  243. Narmani, A.; Teponno, R.B.; Arzanlou, M.; Surup, F.; Helaly, S.E.; Wittstein, K.; Praditya, D.F.; Babai-Ahari, A.; Steinmann, E.; Stadler, M. Cytotoxic, Antimicrobial and Antiviral Secondary Metabolites Produced by the Plant Pathogenic Fungus Cytospora sp. CCTU A309. Fitoterapia 2019, 134, 314–322. [Google Scholar] [CrossRef]
  244. Kumar, R.; Subramani, R.; Aalbersberg, W. Three Bioactive Sesquiterpene Quinones from the Fijian Marine Sponge of the Genus hippospongia. Nat. Prod. Res. 2013, 27, 1488–1491. [Google Scholar] [CrossRef]
  245. Scott Obach, R.; Kalgutkar, A.S. 1.15—Reactive Electrophiles and Metabolic Activation. In Comprehensive Toxicology, 2nd ed.; McQueen, C.A., Ed.; Elsevier: Oxford, UK, 2010; pp. 309–347. ISBN 978-0-08-046884-6. [Google Scholar]
  246. Alamgir, A.N.M. Therapeutic Use of Medicinal Plants and Their Extracts: Volume 2. In Phytochemistry and Bioactive Compounds; Progress in Drug Research; Springer: Cham, Switzerland, 2018; Volume 74, ISBN 978-3-319-92387-1. [Google Scholar]
  247. Nishina, A.; Uchibori, T. Antimicrobial Activity of 2,6-Dimethoxy-p-Benzoquinone, Isolated from Thick-Stemmed Bamboo, and Its Analogs. Agric. Biol. Chem. 1991, 55, 2395–2398. [Google Scholar] [CrossRef]
  248. Sánchez-Duffhues, G.; Calzado, M.A.; de Vinuesa, A.G.; Caballero, F.J.; Ech-Chahad, A.; Appendino, G.; Krohn, K.; Fiebich, B.L.; Muñoz, E. Denbinobin, a Naturally Occurring 1,4-Phenanthrenequinone, Inhibits HIV-1 Replication through an NF-ΚB-Dependent Pathway. Biochem. Pharmacol. 2008, 76, 1240–1250. [Google Scholar] [CrossRef]
  249. Shao, S.-Y.; Wang, C.; Han, S.-W.; Li, S. Two New Phenanthrenequinones with Cytotoxic Activity from the Tubers of Pleione bulbocodioides. Phytochem. Lett. 2020, 35, 6–9. [Google Scholar] [CrossRef]
  250. Kazmi, M.H.; Malik, A.; Hameed, S.; Akhtar, N.; Noor Ali, S. An Anthraquinone Derivative from Cassia italica. Phytochemistry 1994, 36, 761–763. [Google Scholar] [CrossRef]
  251. Tran, T.; Saheba, E.; Arcerio, A.V.; Chavez, V.; Li, Q.; Martinez, L.E.; Primm, T.P. Quinones as Antimycobacterial Agents. Bioorganic Med. Chem. 2004, 12, 4809–4813. [Google Scholar] [CrossRef]
  252. Bisio, A.; Romussi, G.; Russo, E.; Cafaggi, S.; Schito, A.M.; Repetto, B.; De Tommasi, N. Antimicrobial Activity of the Ornamental Species Salvia corrugata, a Potential New Crop for Extractive Purposes. J. Agric. Food Chem. 2008, 56, 10468–10472. [Google Scholar] [CrossRef] [PubMed]
  253. Wang, Q.; Leong, W.F.; Elias, R.J.; Tikekar, R.V. UV-C Irradiated Gallic Acid Exhibits Enhanced Antimicrobial Activity via Generation of Reactive Oxidative Species and Quinone. Food Chem. 2019, 287, 303–312. [Google Scholar] [CrossRef] [PubMed]
  254. Vaughan, P.P.; Novotny, P.; Haubrich, N.; McDonald, L.; Cochran, M.; Serdula, J.; Amin, R.W.; Jeffrey, W.H. Bacterial Growth Response to Photoactive Quinones. Photochem. Photobiol. 2010, 86, 1327–1333. [Google Scholar] [CrossRef] [PubMed]
  255. Subramani, P.A.; Panati, K.; Lebaka, V.R.; Reddy, D.D.; Narala, V.R. Chapter 21—Nanostructures for Curcumin Delivery: Possibilities and Challenges. In Nano- and Microscale Drug Delivery Systems; Grumezescu, A.M., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 393–418. ISBN 978-0-323-52727-9. [Google Scholar]
  256. Zorofchian Moghadamtousi, S.; Abdul Kadir, H.; Hassandarvish, P.; Tajik, H.; Abubakar, S.; Zandi, K. A Review on Antibacterial, Antiviral, and Antifungal Activity of Curcumin. BioMed Res. Int. 2014, 2014, 1–12. [Google Scholar] [CrossRef] [Green Version]
  257. Carvalho, A.C.; Gomes, A.C.; Pereira-Wilson, C.; Lima, C.F. Chapter 35—Mechanisms of Action of Curcumin on Aging: Nutritional and Pharmacological Applications. In Molecular Basis of Nutrition and Aging; Malavolta, M., Mocchegiani, E., Eds.; Academic Press: San Diego, CA, USA, 2016; pp. 491–511. ISBN 978-0-12-801816-3. [Google Scholar]
  258. Li, L.; Leung, P.S. Chapter 13—Pancreatic Cancer, Pancreatitis, and Oxidative Stress. In Gastrointestinal Tissue; Gracia-Sancho, J., Salvadó, J., Eds.; Academic Press: Cambridge, MA, USA, 2017; pp. 173–186. ISBN 978-0-12-805377-5. [Google Scholar]
  259. Priyadarsini, K.I. Photophysics, Photochemistry and Photobiology of Curcumin: Studies from Organic Solutions, Bio-Mimetics and Living Cells. J. Photochem. Photobiol. C Photochem. Rev. 2009, 10, 81–95. [Google Scholar] [CrossRef]
  260. Jurenka, J.S. Anti-Inflammatory Properties of Curcumin, a Major Constituent of Curcuma Longa: A Review of Preclinical and Clinical Research. Altern. Med. Rev. 2009, 14, 141–153. [Google Scholar]
  261. Fadus, M.C.; Lau, C.; Bikhchandani, J.; Lynch, H.T. Curcumin: An Age-Old Anti-Inflammatory and Anti-Neoplastic Agent. J. Tradit. Complement. Med. 2017, 7, 339–346. [Google Scholar] [CrossRef] [Green Version]
  262. Vieira, T.M.; dos Santos, I.A.; Silva, T.S.; Martins, C.H.G.; Crotti, A.E.M. Antimicrobial Activity of Monoketone Curcuminoids Against Cariogenic Bacteria. Chem. Biodivers. 2018, 15, e1800216. [Google Scholar] [CrossRef]
  263. Preis, E.; Baghdan, E.; Agel, M.R.; Anders, T.; Pourasghar, M.; Schneider, M.; Bakowsky, U. Spray Dried Curcumin Loaded Nanoparticles for Antimicrobial Photodynamic Therapy. Eur. J. Pharm. Biopharm. 2019, 142, 531–539. [Google Scholar] [CrossRef]
  264. Naeini, M.B.; Momtazi, A.A.; Jaafari, M.R.; Johnston, T.P.; Barreto, G.; Banach, M.; Sahebkar, A. Antitumor Effects of Curcumin: A Lipid Perspective. J. Cell. Physiol. 2019, 234, 14743–14758. [Google Scholar] [CrossRef]
  265. Liu, F.; Gao, S.; Yang, Y.; Zhao, X.; Fan, Y.; Ma, W.; Yang, D.; Yang, A.; Yu, Y. Antitumor Activity of Curcumin by Modulation of Apoptosis and Autophagy in Human Lung Cancer A549 Cells through Inhibiting PI3K/Akt/MTOR Pathway. Oncol. Rep. 2018, 39, 1523–1531. [Google Scholar] [CrossRef]
  266. Lestari, M.L.A.D.; Indrayanto, G. Chapter Three—Curcumin. In Profiles of Drug Substances, Excipients and Related Methodology; Brittain, H.G., Ed.; Academic Press: Cambridge, MA, USA, 2014; Volume 39, pp. 113–204. [Google Scholar]
  267. 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] [PubMed] [Green Version]
  268. Anand, P.; Kunnumakkara, A.B.; Newman, R.A.; Aggarwal, B.B. Bioavailability of Curcumin: Problems and Promises. Mol. Pharm. 2007, 4, 807–818. [Google Scholar] [CrossRef] [PubMed]
  269. Wikene, K.O.; Bruzell, E.; Tønnesen, H.H. Characterization and Antimicrobial Phototoxicity of Curcumin Dissolved in Natural Deep Eutectic Solvents. Eur. J. Pharm. Sci. 2015, 80, 26–32. [Google Scholar] [CrossRef] [PubMed]
  270. Battista, S.; Maggi, M.A.; Bellio, P.; Galantini, L.; D’Archivio, A.A.; Celenza, G.; Colaiezzi, R.; Giansanti, L. Curcuminoids-Loaded Liposomes: Influence of Lipid Composition on Their Physicochemical Properties and Efficacy as Delivery Systems. Colloids Surf. A Physicochem. Eng. Asp. 2020, 597, 124759. [Google Scholar] [CrossRef]
  271. Shome, S.; Talukdar, A.D.; Choudhury, M.D.; Bhattacharya, M.K.; Upadhyaya, H. Curcumin as Potential Therapeutic Natural Product: A Nanobiotechnological Perspective. J. Pharm. Pharmacol. 2016, 68, 1481–1500. [Google Scholar] [CrossRef]
  272. Kali, A.; Bhuvaneshwar, D.; Charles, P.M.V.; Seetha, K.S. Antibacterial Synergy of Curcumin with Antibiotics against Biofilm Producing Clinical Bacterial Isolates. J. Basic Clin. Pharm. 2016, 7, 93–96. [Google Scholar] [CrossRef] [Green Version]
  273. Gupta, A.; Mahajan, S.; Sharma, R. Evaluation of Antimicrobial Activity of Curcuma Longa Rhizome Extract against Staphylococcus aureus. Biotechnol. Rep. 2015, 6, 51–55. [Google Scholar] [CrossRef] [Green Version]
  274. Morão, L.G.; Polaquini, C.R.; Kopacz, M.; Torrezan, G.S.; Ayusso, G.M.; Dilarri, G.; Cavalca, L.B.; Zielińska, A.; Scheffers, D.-J.; Regasini, L.O.; et al. A Simplified Curcumin Targets the Membrane of Bacillus subtilis. Microbiologyopen 2019, 8, e00683. [Google Scholar] [CrossRef] [Green Version]
  275. Shlar, I.; Droby, S.; Choudhary, R.; Rodov, V. The Mode of Antimicrobial Action of Curcumin Depends on the Delivery System: Monolithic Nanoparticles vs. Supramolecular Inclusion Complex. RSC Adv. 2017, 7, 42559–42569. [Google Scholar] [CrossRef] [Green Version]
  276. 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]
  277. Hegge, A.B.; Andersen, T.; Melvik, J.E.; Bruzell, E.; Kristensen, S.; Tønnesen, H.H. Formulation and Bacterial Phototoxicity of Curcumin Loaded Alginate Foams for Wound Treatment Applications: Studies on Curcumin and Curcuminoides XLII. J. Pharm. Sci. 2011, 100, 174–185. [Google Scholar] [CrossRef] [PubMed]
  278. Wikene, K.O.; Hegge, A.B.; Bruzell, E.; Tønnesen, H.H. Formulation and Characterization of Lyophilized Curcumin Solid Dispersions for Antimicrobial Photodynamic Therapy (APDT): Studies on Curcumin and Curcuminoids LII. Drug Dev. Ind. Pharm. 2015, 41, 969–977. [Google Scholar] [CrossRef] [PubMed]
  279. Hegge, A.B.; Nielsen, T.T.; Larsen, K.L.; Bruzell, E.; Tønnesen, H.H. Impact of Curcumin Supersaturation in Antibacterial Photodynamic Therapy—Effect of Cyclodextrin Type and Amount: Studies on Curcumin and Curcuminoides XLV. J. Pharm. Sci. 2012, 101, 1524–1537. [Google Scholar] [CrossRef]
  280. Hegge, A.B.; Vukicevic, M.; Bruzell, E.; Kristensen, S.; Tønnesen, H.H. Solid Dispersions for Preparation of Phototoxic Supersaturated Solutions for Antimicrobial Photodynamic Therapy (APDT): Studies on Curcumin and curcuminoides L. Eur. J. Pharm. Biopharm. 2013, 83, 95–105. [Google Scholar] [CrossRef]
  281. Hegge, A.B.; Bruzell, E.; Kristensen, S.; Tønnesen, H.H. Photoinactivation of Staphylococcus Epidermidis Biofilms and Suspensions by the Hydrophobic Photosensitizer Curcumin—Effect of Selected Nanocarrier. Eur. J. Pharm. Sci. 2012, 47, 65–74. [Google Scholar] [CrossRef]
  282. Bruzell, E.M.; Morisbak, E.; Tønnesen, H.H. Studies on Curcumin and Curcuminoids. XXIX. Photoinduced Cytotoxicity of Curcumin in Selected Aqueous Preparations. Photochem. Photobiol. Sci. 2005, 4, 523–530. [Google Scholar] [CrossRef] [PubMed]
  283. Condat, M.; Mazeran, P.-E.; Malval, J.-P.; Lalevée, J.; Morlet-Savary, F.; Renard, E.; Langlois, V.; Andalloussi, S.A.; Versace, D.-L. Photoinduced Curcumin Derivative-Coatings with Antibacterial Properties. RSC Adv. 2015, 5, 85214–85224. [Google Scholar] [CrossRef]
  284. Chai, Z.; Zhang, F.; Liu, B.; Chen, X.; Meng, X. Antibacterial Mechanism and Preservation Effect of Curcumin-Based Photodynamic Extends the Shelf Life of Fresh-Cut Pears. LWT 2021, 142, 110941. [Google Scholar] [CrossRef]
  285. Huang, J.; Chen, B.; Li, H.; Zeng, Q.-H.; Wang, J.J.; Liu, H.; Pan, Y.; Zhao, Y. Enhanced Antibacterial and Antibiofilm Functions of the Curcumin-Mediated Photodynamic Inactivation against Listeria Monocytogenes. Food Control 2020, 108, 106886. [Google Scholar] [CrossRef]
  286. Garcia-Gomes, A.S.; Curvelo, J.A.R.; Soares, R.M.A.; Ferreira-Pereira, A. Curcumin Acts Synergistically with Fluconazole to Sensitize a Clinical Isolate of Candida Albicans Showing a MDR Phenotype. Med. Mycol. 2012, 50, 26–32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Sharma, M.; Manoharlal, R.; Negi, A.S.; Prasad, R. Synergistic Anticandidal Activity of Pure Polyphenol Curcumin I in Combination with Azoles and Polyenes Generates Reactive Oxygen Species Leading to Apoptosis. FEMS Yeast Res. 2010, 10, 570–578. [Google Scholar] [CrossRef] [PubMed]
  288. Batista de Andrade Neto, J.; Pessoa de Farias Cabral, V.; Brito Nogueira, L.F.; Rocha da Silva, C.; Gurgel do Amaral Valente Sá, L.; Ramos da Silva, A.; Barbosa da Silva, W.M.; Silva, J.; Marinho, E.S.; Cavalcanti, B.C.; et al. Anti-MRSA Activity of Curcumin in Planktonic Cells and Biofilms and Determination of Possible Action Mechanisms. Microb. Pathog. 2021, 155, 104892. [Google Scholar] [CrossRef] [PubMed]
  289. Dai, C.; Wang, Y.; Sharma, G.; Shen, J.; Velkov, T.; Xiao, X. Polymyxins–Curcumin Combination Antimicrobial Therapy: Safety Implications and Efficacy for Infection Treatment. Antioxidants 2020, 9, 506. [Google Scholar] [CrossRef]
  290. Teow, S.-Y.; Ali, S. Synergistic Antibacterial Activity of Curcumin with Antibiotics against Staphylococcus aureus. Pak. J. Pharm. Sci. 2015, 28, 2109–2114. [Google Scholar]
  291. Li, Y.; Xu, Y.; Liao, Q.; Xie, M.; Tao, H.; Wang, H.-L. Synergistic Effect of Hypocrellin B and Curcumin on Photodynamic Inactivation of Staphylococcus aureus. Microb. Biotechnol. 2021, 14, 692–707. [Google Scholar] [CrossRef]
  292. Itzia Azucena, R.-C.; José Roberto, C.-L.; Martin, Z.-R.; Rafael, C.-Z.; Leonardo, H.-H.; Gabriela, T.-P.; Araceli, C.-R. Drug Susceptibility Testing and Synergistic Antibacterial Activity of Curcumin with Antibiotics against Enterotoxigenic Escherichia coli. Antibiotics 2019, 8, 43. [Google Scholar] [CrossRef] [Green Version]
  293. Gülen, D.; Şafak, B.; Erdal, B.; Günaydın, B. Curcumin-Meropenem Synergy in Carbapenem Resistant Klebsiella Pneumoniae Curcumin-Meropenem Synergy. Iran J. Microbiol. 2021, 13, 345–351. [Google Scholar] [CrossRef]
  294. Marathe, S.A.; Kumar, R.; Ajitkumar, P.; Nagaraja, V.; Chakravortty, D. Curcumin Reduces the Antimicrobial Activity of Ciprofloxacin against Salmonella Typhimurium and Salmonella typhi. J. Antimicrob. Chemother. 2013, 68, 139–152. [Google Scholar] [CrossRef] [Green Version]
  295. Singh, A.; Singh, J.V.; Rana, A.; Bhagat, K.; Gulati, H.K.; Kumar, R.; Salwan, R.; Bhagat, K.; Kaur, G.; Singh, N.; et al. Monocarbonyl Curcumin-Based Molecular Hybrids as Potent Antibacterial Agents. ACS Omega 2019, 4, 11673–11684. [Google Scholar] [CrossRef] [Green Version]
  296. Sahne, F.; Mohammadi, M.; Najafpour, G.D.; Moghadamnia, A.A. Enzyme-Assisted Ionic Liquid Extraction of Bioactive Compound from Turmeric (Curcuma longa L.): Isolation, Purification and Analysis of Curcumin. Ind. Crops Prod. 2017, 95, 686–694. [Google Scholar] [CrossRef]
  297. Zhou, W.; Wang, Z.; Mo, H.; Zhao, Y.; Li, H.; Zhang, H.; Hu, L.; Zhou, X. Thymol Mediates Bactericidal Activity against Staphylococcus aureus by Targeting an Aldo–Keto Reductase and Consequent Depletion of NADPH. J. Agric. Food Chem. 2019, 67, 8382–8392. [Google Scholar] [CrossRef] [PubMed]
  298. Kaur, G.; Kaur, M.; Bansal, M. New Insights of Structural Activity Relationship of Curcumin and Correlating Their Efficacy in Anticancer Studies with Some Other Similar Molecules. Am. J. Cancer Res. 2021, 11, 3755–3765. [Google Scholar] [PubMed]
  299. Yang, H.; Du, Z.; Wang, W.; Song, M.; Sanidad, K.; Sukamtoh, E.; Zheng, J.; Tian, L.; Xiao, H.; Liu, Z.; et al. Structure–Activity Relationship of Curcumin: Role of the Methoxy Group in Anti-Inflammatory and Anticolitis Effects of Curcumin. J. Agric. Food Chem. 2017, 65, 4509–4515. [Google Scholar] [CrossRef]
  300. Ahsan, H.; Parveen, N.; Khan, N.U.; Hadi, S.M. Pro-Oxidant, Anti-Oxidant and Cleavage Activities on DNA of Curcumin and Its Derivatives Demethoxycurcumin and Bisdemethoxycurcumin. Chem.-Biol. Interact. 1999, 121, 161–175. [Google Scholar] [CrossRef]
  301. Hatamipour, M.; Ramezani, M.; Tabassi, S.A.S.; Johnston, T.P.; Sahebkar, A. Demethoxycurcumin: A Naturally Occurring Curcumin Analogue for Treating Non-Cancerous Diseases. J. Cell. Physiol. 2019, 234, 19320–19330. [Google Scholar] [CrossRef]
  302. Harvey, A. Strategies for Discovering Drugs from Previously Unexplored Natural Products. Drug Discov. Today 2000, 5, 294–300. [Google Scholar] [CrossRef]
  303. Koehn, F.E.; Carter, G.T. The Evolving Role of Natural Products in Drug Discovery. Nat. Rev. Drug Discov. 2005, 4, 206–220. [Google Scholar] [CrossRef]
  304. Koparde, A.A.; Doijad, R.C.; Magdum, C.S. Natural Products in Drug Discovery. In Pharmacognosy—Medicinal Plants; IntechOpen: London, UK, 2019; pp. 1–19. [Google Scholar]
  305. Montaser, R.; Luesch, H. Marine Natural Products: A New Wave of Drugs? Future Med. Chem. 2011, 3, 1475–1489. [Google Scholar] [CrossRef] [Green Version]
  306. Negi, P.S. Plant Extracts for the Control of Bacterial Growth: Efficacy, Stability and Safety Issues for Food Application. Int. J. Food Microbiol. 2012, 156, 7–17. [Google Scholar] [CrossRef]
  307. Wu, S.; Chappell, J. Metabolic Engineering of Natural Products in Plants; Tools of the Trade and Challenges for the Future. Curr. Opin. Biotechnol. 2008, 19, 145–152. [Google Scholar] [CrossRef] [PubMed]
  308. Ginsburg, H.; Deharo, E. A Call for Using Natural Compounds in the Development of New Antimalarial Treatments—An Introduction. Malar. J. 2011, 10, S1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  309. Dey, A. CRISPR/Cas Genome Editing to Optimize Pharmacologically Active Plant Natural Products. Pharmacol. Res. 2021, 164, 105359. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Representative chemical structures of simple phenol, benzoic acid derivatives and cinnamic acid derivatives.
Figure 1. Representative chemical structures of simple phenol, benzoic acid derivatives and cinnamic acid derivatives.
Futurepharmacol 02 00030 g001
Figure 2. Representative chemical structures of different classes of coumarin and nature-derived coumarin.
Figure 2. Representative chemical structures of different classes of coumarin and nature-derived coumarin.
Futurepharmacol 02 00030 g002
Figure 3. Representative chemical structures of different classes of flavonoids.
Figure 3. Representative chemical structures of different classes of flavonoids.
Futurepharmacol 02 00030 g003
Figure 4. Representative chemical structures of different classes of tannins.
Figure 4. Representative chemical structures of different classes of tannins.
Futurepharmacol 02 00030 g004
Figure 5. Representative chemical structures of different classes of stilbene and nature-derived stilbenes.
Figure 5. Representative chemical structures of different classes of stilbene and nature-derived stilbenes.
Futurepharmacol 02 00030 g005
Figure 6. Representative chemical structures of different lignan types and lignan subclasses.
Figure 6. Representative chemical structures of different lignan types and lignan subclasses.
Futurepharmacol 02 00030 g006
Figure 7. Representative chemical structures of different classes of quinones.
Figure 7. Representative chemical structures of different classes of quinones.
Futurepharmacol 02 00030 g007
Figure 8. Representative chemical structures of curcuminoid compounds, and tautomerism of curcumin.
Figure 8. Representative chemical structures of curcuminoid compounds, and tautomerism of curcumin.
Futurepharmacol 02 00030 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ecevit, K.; Barros, A.A.; Silva, J.M.; Reis, R.L. Preventing Microbial Infections with Natural Phenolic Compounds. Future Pharmacol. 2022, 2, 460-498. https://doi.org/10.3390/futurepharmacol2040030

AMA Style

Ecevit K, Barros AA, Silva JM, Reis RL. Preventing Microbial Infections with Natural Phenolic Compounds. Future Pharmacology. 2022; 2(4):460-498. https://doi.org/10.3390/futurepharmacol2040030

Chicago/Turabian Style

Ecevit, Kardelen, Alexandre A. Barros, Joana M. Silva, and Rui L. Reis. 2022. "Preventing Microbial Infections with Natural Phenolic Compounds" Future Pharmacology 2, no. 4: 460-498. https://doi.org/10.3390/futurepharmacol2040030

Article Metrics

Back to TopTop