Next Article in Journal
AlH3 as High-Energy Fuels for Solid Propellants: Synthesis, Thermodynamics, Kinetics, and Stabilization
Previous Article in Journal
Synthesis and Biological Activities of Some Metal Complexes of 2-Thiouracil and Its Derivatives: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antimicrobial Activity of Dimeric Flavonoids

by
Inês Lopes
1,2,
Carla Campos
2,
Rui Medeiros
2,3,4 and
Fátima Cerqueira
2,3,4,5,*
1
School of Health, Polytechnic Institute of Porto, Rua Dr. António Bernardino de Almeida 400, 4200-072 Porto, Portugal
2
Molecular Oncology and Viral Pathology Group, Research Center of IPO Porto (CI-IPOP)/RISE@CI-IPOP (Health Research Network), Portuguese Oncology Institute of Porto (IPO Porto)/Porto Comprehensive Cancer Center (Porto.CCC) Raquel Seruca, Rua António Bernardino de Almeida, 4200-072 Porto, Portugal
3
FP-I3ID, FP-BHS, GIT-LoSa, University Fernando Pessoa, Praça 9 de Abril, 349, 4249-004 Porto, Portugal
4
Faculty of Health Sciences, University Fernando Pessoa, Rua Carlos da Maia, 296, 4200-150 Porto, Portugal
5
CINTESIS.UFP@RISE, Centro de Investigação em Tecnologias e Serviços de Saúde, Rede de Investigação em Saúde, Universidade Fernando Pessoa, Praça de 9 de Abril, 349, 4249-004 Porto, Portugal
*
Author to whom correspondence should be addressed.
Compounds 2024, 4(2), 214-229; https://doi.org/10.3390/compounds4020011
Submission received: 4 December 2023 / Revised: 5 February 2024 / Accepted: 15 March 2024 / Published: 22 March 2024

Abstract

:
Distributed throughout the environment are various microorganisms such as bacteria, fungi, parasites, and viruses. Although many are part of the human microbiome, many are pathogenic and cause infections ranging from mild to severe. In recent years, the identification of multidrug-resistant microorganisms has become a serious public health problem. The resulting infections call into question the therapeutic capacity of health systems and lead to approximately 70,000 deaths annually worldwide. The progressive resistance to antibiotics and antifungals has been a major challenge for the medical and pharmaceutical community, requiring the search for new compounds with antimicrobial properties. Several studies have demonstrated the potential of natural and synthesized flavonoids, especially the dimers of these molecules. In this review are presented many examples of dimeric flavonoids that have demonstrated antimicrobial activity against viruses, like influenza and Human Immunodeficiency Virus (HIV), protozoal infections, such as Leishmaniasis and Malaria, fungal infections by Candida albicans and Cryptococcus neoformans, and bacterial infections caused, for example, by Staphylococcus aureus and Escherichia coli. In the pursuit to find potential safe agents for therapy in microbial infections, natural dimeric flavonoids are an option not only for the antimicrobial activity, but also for the low toxicity usually associated with these compounds when compared to classic antimicrobials.

1. Introduction

During the last few decades, the incidence of microbial infections has increased significantly, which culminates in high morbidity and mortality rates [1]. This is due to the existence of more immunocompromised individuals (due to HIV infection, chemotherapy, and radiotherapy treatments), use of immunosuppressants, an increased number of hospitalized patients, invasive devices and procedures in medical practice and the evolution of virulence and resistance mechanisms to antimicrobial agents [2,3,4].
Although antibiotics and antifungals have a wide spectrum and different mechanisms of action, their incorrect and indiscriminate use has had the consequence of the increase in resistance mechanisms developed by microorganisms [5]. The medical community has limited options for resolving bacterial and fungal infections [6,7].
Parasitic diseases affect over 1 billion people all over the world, specifically, parasitic diseases such as malaria and schistosomiasis, leading to approximately 1 million deaths throughout the world [8]. Drug resistance associated with the treatment of these infections are widespread and, even with some genes already found related to resistance for available therapy, the mechanisms are not completely understood [9].
Antiviral resistance has been widely studied and is commonly related with factors that involve a decrease in host immunity and prolongs the duration of treatment [8]. Consequently, there is an increase in side effects due to the toxicity caused using second-line antivirals and, in cases of serious illness, death due to progressive viral infection when there is no cure available [10].
The efficacy of a therapeutic agent may be affected by the development of withdrawal mechanisms from the first time it is applied [11]. The increasing resistance to antifungals and antibiotics available for clinical practice has been a major challenge for the medical and pharmaceutical communities, requiring the search for new compounds with antimicrobial properties (e.g., of plant origin). Besides that, this strategy may involve the search for new compounds to counteract resistance mechanisms or make it possible to reduce the dose of the antimicrobial and, consequently, its toxicity and adverse effects [11,12].
Several studies have proven the individual or synergistic antimicrobial potential of natural and synthetic flavonoids, against drug-resistant fungi [13], bacteria [14], virus [15] and parasites [16].
Flavonoids are a versatile group of phenolic compounds produced as secondary metabolites by plants and, therefore, existing in the human diet (e.g., present in fruit, vegetables, cereals, wine, and various teas) [17]. These compounds are responsible for the coloration of leaves, flowers and fruits, and have a fundamental role in the protection of plants as oxidizing and microbial agents [18]. As a class of polyphenols, flavonoids can be divided in flavones (apigenin), flavanones (naringenin), flavans (catechin), flavonoid glycosides, flavonols (quercetin), flavonolignan (silibinin), chalcones (butein), isoflavones (genistein), aurones (aureusidin), leucoanthocynidins (leucopelargonidin) and neoflavonoids (neoflavones) [19,20,21].
Flavonoids were described in the literature as processing anti-allergic [22,23,24,25], anti-inflammatory [25,26,27,28,29,30], immunomodulatory [31,32,33], antitumor [33,34,35,36] and antimicrobial [15,21,37,38,39,40,41] properties, which are the reasons explaining their great interest in the food, pharmaceutical and medical industries. Their toxicity levels are greatly reduced and are therefore currently critical for the development of new medicines [11,13,21,42,43,44].
Dimeric flavonoids are a class of flavonoids that consists of the same or diverse flavonoid units connected by C-C bonds or by C-O-C bonds. These dimers are joined in a symmetrical or unsymmetrical manner through an alkyl or an alkoxy-based linker of varying length [45,46]. Mostly, dimeric compounds are formed by dimers of flavone–flavone, flavone–flavonone, and flavonone–flavonone subunits, as well as dimers of chalcones and isoflavones [45]. These compounds are called bis-flavonoids when they have two equal units or biflavonoids when there are two different units in the dimer structure. Since several dimeric compounds have been identified, the scientific community have been interested in their antimicrobial properties [45,46].
The main goal of this review is summarizing the remarks of several published studies on the use of dimeric flavonoids as antimicrobial agents, analyzing their role in aiding or resolving fungal, bacterial, parasitic, and viral infections. Thus, this review aims to outline the potential mechanism of actions of dimeric flavonoids studied in vitro and in vivo, and the perspectives of their use as multi-targets agents or conjugated with antimicrobials already known and applied in the treatment of infections.
A thorough search of the relevant scientific databases, including Web of Science, ScienceDirect, Scopus, PubMed, and Google Scholar, was conducted. The keyword combinations used in all databases were as follows: (antimicrobial resistance AND antibacterial resistance) OR (antimicrobial resistance AND antifungal resistance) OR (antimicrobial resistance AND antiparasitic resistance); (flavonoids AND biological properties) OR (flavonoids and antimicrobial activity) OR (flavonoids AND (dimeric flavonoids OR dimeric compounds); (dimeric flavonoids AND antimicrobial activity) OR (dimeric flavonoids AND antiviral activity) OR (dimeric flavonoids AND antifungal activity) OR (dimeric flavonoids AND (antiprotozoal OR anthelmintic activity) OR (dimeric flavonoids AND antibacterial activity); (dimeric flavonoids AND (clinical studies OR in vivo studies OR pre-clinic tests OR assays in animal models).
Preliminary reading and analysis allowed the selection of several studies published between 2000 and 2023, which were later thoroughly analyzed. Studies written in Portuguese, Brazilian and English were selected.
Abstracts of selected titles were reviewed based on some inclusion and exclusion criteria. The articles that described the antimicrobial assays but did not state the respective control experiments, as well as studies describing it in a contradictory or unclear manner, were excluded from the review.

2. Dimeric Flavonoids

In 1929, ginkgetin, the first dimeric flavonoid, was separated from Ginkgo biloba by Furukawa and opened a new path for the discovery of more than 500,000 of those compounds, such as amenthoflavone, agatisflavone, cupressoflavone, hynoquiflavone and robustaflavone [47]. Due to their chemical and biological properties, there has been a great evolution regarding phytochemical chemical studies for the manipulation, molecular rearrangement strategies, identification, and synthesis of new bioactive dimeric flavonoids with potentiated characteristics [46].
Dimeric flavonoids are extensively studied for their pharmacological properties, as they have low toxicity in human cells [46,48,49,50,51], which has opened new routes to find and synthetize new drugs against pathogens.
Recently, research has extensively reported that the biologic activity of these compounds is higher than monomeric flavonoids, due to the high number of hydroxyl groups that reduce hydrophobicity [46,52,53,54,55,56].
Despite being very promising compounds, there are very few that are completely studied, with elucidated mechanisms of action and with their toxicity investigated [46].

2.1. Antiviral Activity of Dimeric Flavonoids

Dimeric flavonoids can act against many RNA and DNA viruses by blocking different stages of virus life cycle: fixation and entry in the cells, interference with replication and formation, maturation, and liberation of new mature viral particles [46]. In addition, these compounds might be indirect inhibitors by interacting with immune cells of the host [46].
Due to resistance to antiviral drugs, it has become vital to search for more compounds that can reduce the side effects, viral latency, and recurrence of infections. Nonetheless, the emergence of new viruses brings many obstacles to medicine [46]. These dimeric compounds seem to be more promising than flavonoids due to their greater physical–chemical stability during tests of pharmacokinetic parameters [57].
Dimeric flavonoids, such as amentoflavone, have become compounds of interest due to their important antiviral effects, mainly as protease inhibitors [58]. Several reports describe the activity of natural products against coronaviruses (CoV), with the main target being viral replication. Of these, numerous flavonoids, such as quercetin, generated strong antiviral activity, affecting SARS-CoV, MERS-CoV, and SARS-CoV-2 proteases [57]. In addition to the characteristic symptoms of COVID-19, SARS-CoV-2 infection can also result in complications, one of the most worrying being cytokine storm that can lead, in the worst-case scenario, to multiple organ failure and death [59,60].
Bearing this in mind, it is also necessary to use drugs to treat these diseases which, in addition to having a direct antiviral effect, also can modulate the immune response triggered by the infection. Because of their diverse biological activities, dimeric flavonoids may be used in combination with antivirals currently used in the clinic [46]. Table 1 lists some of the compounds that are referred in the literature as possessing effective antiviral activity.
Chaves, O. et al. used not only the dimeric flavonoid agatisflavone, but also its natural monomer apigenin, and demonstrated that the dimeric form increased the antiviral capacity of flavonoids, which might be explained by the top number of hydrophobic contacts by the number of aromatic rings [57]. The study of Y. Lin et al. also showed that the presence of a greater number of hydroxylated groups and at least one flavone unit in dimeric flavonoid compounds are essential for their antiviral activity. On the other hand, the compounds studied can become inactive when the hydroxyl groups are methylated [65].

2.2. Antifungal Activity of Dimeric Flavonoids

The antifungal activity and, respectively, mechanisms of action of dimeric flavonoids were investigated against several pathogenic fungal strains, such as Candida albicans, Cryptococcus neoformans, Aspergillus fumigatus, Penicillium marneffei, Alternaria alternata, Fusarium culmorum, and Cladosporium oxysporum [46]. The diagram of Figure 1 shows the main targets of dimeric flavonoids against fungi described in the literature.
Dimeric flavonoids can form complexes with soluble proteins in fungal cell walls and the lipophilic nature of these compounds makes them capable of disrupting fungal membranes [46,78,79].
Regarding these two types of fungus, there have been dimeric flavonoids, like isoginkgetin, that have demonstrated antifungal activity against Cryptococcus neoformans and Aspergillous fumigatus, a yeast, and a filamentous fungus, respectively. These data show that these types of compounds have a large spectrum of action [80].
According to the literature, one of the main characteristics of these compounds is their ability to inhibit the growth and multiplication of fungus, like Candida albicans and Alternaria alternata, and the growth of spores [81,82].
When evaluating a possible interference of dimeric flavonoid compounds with virulence factors that determine the pathogenicity of fungi, studies have shown that amentoflavone enables Candida albicans to make a dimorphic transition due to a stress response by the accumulation of trehalose and bilobetin, which is able to inhibit the growth of germinating tubes from Cladosporium oxysporum and Fusarium culmorum [82,83]. As far as the production of toxins is concerned, the compounds amentoflavone, 7,7″-Dimethoxyagastisflavone, 6,6″-bigenkwanin, and tetramethoxy-6,6″-bigenkwanin, isolated from the Ouratea species, inhibited the production of aflatoxins B1 and B2 from Aspergillus flavus, and the maximum effect happened at 10 μg/mL [84].
In the case of Alternaria alternata, ginkgetin and 7-O-methylamentoflavone provoked cell wall changes by an hydrophobic interaction [82].
Since biofilms are an enormous obstacle against antifungal agents, Freitas et al. tested if proanthocyadnidin polymeric tannins from the Stryphnodendron adstringens stem bark with antifungal activity against Candida albicans were also active during biofilm formation and on pre-formed biofilms for Candida spp. The best results for Candida spp. were for C. albicans, with MICs of 3.91 and 0,48 mg/L, that represented the inhibition of planktonic and dispersion cells, respectively [85]. In conclusion, their study highlighted the potential of those dimeric compounds to inhibit the formation of those communities of yeasts [85].
Additionally, some synthetic antifungal dimeric flavonoids were generally more active against Aspergillus niger (MICs of 0.2, 0.0013 and 0.4 µmol/mL of dimers) when compared to correspondent monomeric compounds of apigenin [55].
Considering all these findings, dimeric flavonoids that possess inherent antifungal activity (Table 2) could be a strategy for future antifungal therapy [86].
Interestingly, dimeric flavonoids, like amentoflavone and other compounds consisting of flavanone–flavone units (like 2,3-dihydrosciadopitysin) with a methoxyl group absent, were inactive or weakly effective [58].

2.3. Antiparasitic Activity of Dimeric Flavonoids

Parasitic infections are responsible for a great strain on health systems and affect millions of people around the world [90].
According to the literature, some dimeric flavonoids, such as morelloflavone and strychnobiflavone, show activity against both promastigote and amastigote forms [91]. Considering virulence factors of Leishmania spp., studies showed that the dimeric flavonoids lanaroflavone, podocarpusflavone A, amentoflavone, and podocarpusflavone B, inhibited the action of a zinc-dependent metalloprotease, existing in amastigote and proamastigote forms of L. major and L. panamensis, which reduces the ability of parasites to adhere to macrophages by interaction with fibronectin [46,92].
As for malaria disease, although there are medications such as chloroquine, vector control, and vaccines (about 40% effective) capable of controlling transmission, it remains a serious parasitic infection [46]. Dimeric flavonoids, such as lanaroflavone, methylenebissantin and 3″,4′,4‴,5,5″,7,7″-heptahydroxy-3,8-biflavanone, demonstrated high activity against Plasmodium falciparum, in some cases by inhibiting important enzymes [51,93,94]. Weniger et al. and Kunert et al. stated that the patter of methylation of the compounds are determinants for antiplasmodial activity [94,95].
Other dimeric compounds, such as 2″,3″-Dihydroochnaflavone and brachydins B and C, showed important antiparasitic activity against Trypanosoma cruzi amastigotes and trypomastigotes forms, and inhibited its capacity to invade [96,97].
During this review, no dimeric compounds were found with antiparasitic activity against helminths.
Table 3 summarize some dimeric flavonoids that demonstrated ability to act against protozoa.
The fact that dimeric flavonoids tested with commonly used anti-parasitic drugs revealed the absence of competition/interaction may represent an important strategy that allows reducing the dose, adverse effects, time, and cost of treatments, overcoming the weak activity of some medications when administered individually [103].
Additionally, Ichino et al. and Boniface and Ferreira used the liquiritigenin dimer 3,3″-di(7,4″-dihydroxyflavanone-3-yl) and the monomeric liquiritigenin and stated that the monomeric form did not have antiplasmodial activity [98,99]. Also, Thévenin et al. found that the synthetic compounds of methylenebis(chalcone)s were more active against parasites [104]. These extra data comparing dimeric and monomeric forms enhance the potential of dimeric flavonoid investigation for antiparasitic effects [104].

2.4. Antibacterial Activity of Dimeric Flavonoids

Antibacterial resistance has become a problem of public health recognized all around the globe [5]. To find alternatives for resolving infections caused by multi-drug-resistant bacteria, the medical and pharmacological industry need to search for new products that have functions like those of available antibiotics [41].
Although the mechanism of action of antibacterial dimeric flavonoids might not be elucidated, some authors may assume that they may act in a similar way to monomeric compounds, as show in Figure 2 [46].
In general, these compounds are more potent in Gram-positive rather than Gram-negative bacteria, due to the differences between the cell wall of those two groups of bacteria, especially due to the repulsive effect of lipopolysaccharides present in Gram-negative bacteria [80]. One of the mechanisms characterized is the disruption of plasma membranes [46,80,105,106]. For example, isoginkgetin and podocarpusflavone MICs for S. aureus and E. faecalis were 60.0 µg/mL, which showed a moderate activity; for Gram-negative E. coli and P. aeruginosa, isoginkgetin MICs were 130 µg/mL and podocarpusflavone were 250 and 60 µg/mL, respectively, which represented a lower activity [86].
Other compounds, like macrophylloflavone, can interfere with nucleic acid synthesis and other dimeric flavonoids [105]. Additionally, and just like the compound 3″,4′,4‴,5,5″,7,7″-heptahydoxy-3-8″-biflavone, they can also interfere with the metabolism of the bacterial cell and uptake of crucial nutrients, like glucose [105,107].
As already shown for fungus, dimeric flavonoids like agatisflavone, amentoflavone, tetrahydroamentoflavone (THAF) and fukugiside can inhibit the bacterial growth and inhibit the biofilm formation, such as Bacillus subtilis, Staphylococcus carnosus and Streptococcus pyogenes [108,109].
The investigation of Linden et al. investigation a remarkable antibacterial activity of THAF against Gram-positive microorganisms: B. subtilis, with an MIC and MBC of 0.063 mg/mL and a bactericidal effect of 0.125 mg/mL for S. carnosus. In this case, the results stated for the first time that dimerization and a reduced C–ring in dimeric flavonoids, such as in THAF, may be the answers to justify the highest antibacterial activity. Regarding biofilm inhibition, THAF was able to inhibit the biofilm formation of methicillin-resistant S. aureus (MRSA) [108].
The findings of Nandu et al. on fukugiside showed that a concentration of 80 µg/mL reduces an S. pyogenes biofilm by 91% by minimizing the cell surface hydrophobicity, which do not rely on bacterial viability [109]. Furthermore, this dimeric flavonoid was also able to interfere with an important virulence factor—M proteins—that have antiphagocytic functions, enhancing S. pyogenes rate survival in human tissues and fluids. These proteins are encoded by the emm gene, which is positively regulated by mga. Fukugiside downregulated mga, which represented the possible prevention of systemic spread [109].
In short, Table 4 summarizes dimeric flavonoids that have antibacterial activity and the mechanism of action, when elucidate.
The study of Lee et al. showed that the dose-dependent killing of M. aeruginosa KW could be due to another variety of flavonoids in the S. tamariscina extract. For example, in that work, apigenin, a monomer of amentoflavone, also had cyanobacterial-killing effects. However, those effects were insufficient compared to the ones obtained for amentoflavone [106].
Interestingly, Bitchagno et al. found that the antibacterial activity of the dimeric flavonoid ericoside was higher for drug-resistant E. coli AG100 (MIC = 64 μg/mL) and for Klebsiella pneumoniae ATCC11296 (128 μg/mL) than for monomeric taxifolin 3-O-rhamnopyranoside, in which the MICs found were >128 μg/mL [111].

2.5. Potential of Dimeric Flavonoids as Antimicrobials: Form Lab to Clinics

In the literature, it is easy to find studies that have tested many flavonoids compounds in vivo, using animal models, to assess their antimicrobial activity and toxicity levels, and the results are starting to be potentiated in clinical trials [41].
Recent studies demonstrate that topically applied flavonoids, specially flavonols and flavanols are effective when used via oral and vaginal mucosa routes [44]. The work of Araújo et. al, showed that the in vivo tests in mice with a vaginal cream with an extract from Syngonanthus nitens scapes (having flavonoids as the bioactive compounds) eliminated vaginal Candida albicans, with only signs of inflammatory infiltrate and ulcerations that indicated a previous infectious process in the local mucosa [113]. Simonetti et al. also tested in vivo grape seed extract polymeric flavan-3-ols that inhibited C. albicans load in vaginal candidiasis in mice [114]. Furthermore, research from Seleem et al. showed that the compound lichochalcone-A, applied topically in the oral cavities of immunosuppressed mice, not only resulted in an extensively reduced fungal load, but also did not have significantly toxicological effects, with the absence of tissue necrosis [115].
Regarding in vivo activity against parasites, according to Marín et al., none of the nine flavonoids tested in mice infected with T. cruzi had significant toxicity and the parasitic charge was extensively lower when compared with a benznidazole control [116]. Additionally, those compounds changed the levels of the anti-T. cruzi antibody during the chronic stage [116]. An in vivo study with 2′-Hydroxyflavanone showed that this flavonoid reduced the lesion size and L. amazonensis load in a murine model of cutaneous leishmaniasis [117]. Pereira et al. stated the in vivo schistosomicidal activities of oral treatment with chalcones against Schistosoma mansoni worms, and the results showed that in mice there occurred a total worm reduction [118].
The effects on acute lung injury induced by the influenza A virus in mice of an extract from Scutellaria baicalensis root with bioactive flavonoids showed that oral administration protected the infected animals by decreasing the lung virus load by affecting the production of reduced haemagglutinin and inhibiting neuraminidase activity [119]. Based on the Ma et al. study, the authors highlighted the in vivo activity of oxazinyl flavonoids against tobacco mosaic virus [120].
The clinical test conducted with Plantago lanceolata extracts, with the in vitro antimicrobial activity of flavonoids, demonstrated that the individuals that had a P. lanceolata mouth rinse presented a significant decrease in streptococci compared to the placebo group [121]. Even not significantly, the study stated a minor decrease for lactobacilli counts after the treatment [121]. Another in vivo study with mice showed that an ethanolic extract had a powerful antibacterial action against S. aureus, P. aeruginosa and Listeria monocytogenes, attributed to the high content of catechin, epicatechin gallate and epicatechin, and may be useful as an antiseptic solution [122].
Given these data, it is assumed that the study of flavonoids has increased. However, there are few studies with animal models and clinical trials regarding the antimicrobial activity of dimeric flavonoid compounds [46].
Rocha et al. performed a study in mice that showed that brachydin B, a dimeric compound from Arrabidea Brachypoda, reduced the parasitemia in the infected animals with L. amazonensis [102]. The oral administration suggested that the compound is absorbed by the oral route and can reduce parasitemia [102]. The same compound was also tested against T. cruzi in mice and was possible to highlight the low toxicity and decrease in parasitemia and mortality [97,102]. Therefore, brachydin B appears to be a promising lead for treating Leishmaniasis and Chagas disease [97,102].
In vivo analysis in C. elegans evinced low toxicity of the dimeric compound fukugiside and its anti-virulence potential against S. pyogenes [109]. A study with the dimeric flavonoids amentoflavone and robustaflavone demonstrated their ability to reduce the infection by L. amazonensis in mice [101].

3. Conclusions

Besides their nutritional value, flavonoid compounds have gained special interest, given the numerous studies that have pointed out their potential in clinics [46,123]. Several researchers have demonstrated the individual or synergistic anti-microbial potential of natural and synthetic flavonoids against drug-resistant fungi [13], bacteria [14], viruses [15] and parasites [16].
Currently, the dimeric flavonoids offer an opportunity for new therapeutic drugs, as proven by the many compounds studied, not only for biological features, but also by toxicity levels [46]. Regarding viruses, amentoflavone and agathisflavone have shown a high spectrum of anti-viral activity against herpes simplex, influenza, dengue, and SARS-CoV-2, with viral enzymes being the main targets of overall compounds [50,61,68,72,73].
In the group of fungi, dimeric flavonoids have more activity towards C. albicans, like amentoflavone and proanthocyanidin, and the overall targets are enzymes, biofilm and germinative tube formation [85,89].
Amentoflavone and morelloflavone are compounds with promising effects against Leishmania spp. by interacting with enzymes and enhancing antioxidant activity [96,106]. Brachydins discovered in Brazil from the plant Arrabidaea brachypoda have strong activity against Trypanosoma cruzi by inhibiting the parasite invasion and its intracellular multiplication in host cells [101,102]. In bacteria, many compounds, such as macrophylloflavone, isoginkgetin and ericoside, have shown activity against Staphylococcus aureus and Escherichia coli. Although some of the mechanisms of action are unknown, major dimeric compounds with antibacterial activity are showing the ability to interfere with nucleic acid synthesis, cytoplasmic membrane function, energy metabolism, and porins in cell membranes [80,105,111].
Future work on the long road to implement the clinical use of these dimeric compounds is needed to clarify their mechanisms of action and toxicity levels. In vitro findings open great possibility for carrying out tests on animal models and clinical trials [46,97,102,118].
Even with several complicated steps, such as natural isolation, synthesis and modifications, these molecules may be important to fight emergent microbial diseases and especially the threat of antimicrobial resistance [41].

Author Contributions

Conceptualization: I.L., C.C., R.M. and F.C.; writing—original draft preparation: I.L.; writing—review and editing: I.L., C.C., R.M. and F.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded and sponsored by the national funds of FCT/MCTES—Foundation for Science and Technology I.P. from the Ministry of Science, Technology, and Higher Education (PIDDAC) and the European Regional Development Fund (ERDF) by the COMPETE—Programa Operacional Factores de Competitividade (POFC) under the Strategic Funding and the Research Center of the Portuguese Oncology Institute of Porto (project nº. PI86-CI-IPOP-66-2019) and Fundation Fernando Pessoa.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tanwar, J.; Das, S.; Fatima, Z.; Hameed, S. Multidrug resistance: An emerging crisis. Interdiscip. Perspect. Infect. Dis. 2014, 2014, 541340. [Google Scholar] [CrossRef] [PubMed]
  2. Lockhart, S.R.; Guarner, J. Emerging and reemerging fungal infections. Semin. Diagn. Pathol. 2019, 36, 177–181. [Google Scholar] [CrossRef] [PubMed]
  3. Benedict, K.; Richardson, M.; Vallabhaneni, S.; Jackson, B.R.; Chiller, T. Emerging issues, challenges, and changing epidemiology of fungal disease outbreaks. Lancet Infect. Dis. 2017, 17, e403–e411. [Google Scholar] [CrossRef] [PubMed]
  4. Brown, G.D.; Denning, D.W.; Gow, N.A.R.; Levitz, S.M.; Netea, M.G.; White, T.C. Hidden killers: Human fungal infections. Sci. Transl. Med. 2012, 4, 165rv13. [Google Scholar] [CrossRef] [PubMed]
  5. Ferri, M.; Ranucci, E.; Romagnoli, P.; Giaccone, V. Antimicrobial resistance: A global emerging threat to public health systems. Crit. Rev. Food Sci. Nutr. 2015, 57, 2857–2876. [Google Scholar] [CrossRef] [PubMed]
  6. Huemer, M.; Shambat, S.M.; Brugger, S.D.; Zinkernagel, A.S. Antibiotic resistance and persistence—Implications for human health and treatment perspectives. Embo Rep. 2020, 21, e51034. [Google Scholar] [CrossRef] [PubMed]
  7. Larsson, D.G.J.; Flach, C.F. Antibiotic resistance in the environment. Nat. Rev. Microbiol. 2022, 20, 257–269. [Google Scholar] [CrossRef]
  8. Muhaj, F.F.; George, S.J.; Nguyen, C.D.; Tyring, S.K. Antimicrobials and resistance part II: Antifungals, antivirals, and antiparasitics. J. Am. Acad. Dermatol. 2022, 86, 1207–1226. [Google Scholar] [CrossRef]
  9. Pramanik, P.K.; Alam, N.; Chowdhury, D.R.; Chakraborti, T. Drug Resistance in Protozoan Parasites: An Incessant Wrestle for Survival. J. Glob. Antimicrob. Resist. 2019, 18, 1–11. [Google Scholar] [CrossRef]
  10. Strasfeld, L.; Chou, S. Antiviral drug resistance: Mechanisms and clinical implications. Infect. Dis. Clin. N. Am. 2010, 24, 809–833. [Google Scholar] [CrossRef]
  11. Rocha, M.F.G.; Sales, J.A.; da Rocha, M.G.; Galdino, L.M.; de Aguiar, L.; Pereira-Neto, W.d.A.; Cordeiro, R.d.A.; Castelo-Branco, D.d.S.C.M.; Sidrim, J.J.C.; Brilhante, R.S.N. Antifungal effects of the flavonoids kaempferol and quercetin: A possible alternative for the control of fungal biofilms. Biofouling 2019, 35, 320–328. [Google Scholar] [CrossRef] [PubMed]
  12. Cerqueira, F.; Maia, M.; Gabriel, C.; Medeiros, R.; Cravo, S.; Ribeiro, A.I.; Dantas, D.; Dias, A.M.; Saraiva, L.; Raimundo, L.; et al. Mechanism of Antifungal Activity by 5-Aminoimidazole-4-Carbohydrazonamide Derivatives against Candida albicans and Candida krusei. Antibiotics 2021, 10, 183. [Google Scholar] [CrossRef]
  13. Jin, Y.-S. Recent advances in natural antifungal flavonoids and their derivatives. Bioorganic Med. Chem. Lett. 2019, 29, 126589. [Google Scholar] [CrossRef]
  14. Xie, Y.; Yang, W.; Tang, F.; Chen, X.; Ren, L. Antibacterial Activities of Flavonoids: Structure-Activity Relationship and Mechanism. Curr. Med. Chem. 2014, 22, 132–149. [Google Scholar] [CrossRef] [PubMed]
  15. Badshah, S.L.; Faisal, S.; Muhammad, A.; Poulson, B.G.; Emwas, A.H.; Jaremko, M. Antiviral activities of flavonoids. Biomed. Pharmacother. 2021, 140, 111596. [Google Scholar] [CrossRef] [PubMed]
  16. Penna-Coutinho, J.; Aguiar, A.C.; Krettli, A.U. Commercial drugs containing flavonoids are active in mice with malaria and in vitro against chloroquine-resistant Plasmodium falciparum. Mem. Inst. Oswaldo Cruz. 2018, 113, e180279. [Google Scholar] [CrossRef]
  17. Janabi, A.H.W.; Kamboh, A.A.; Saeed, M.; Xiaoyu, L.; BiBi, J.; Majeed, F.; Naveed, M.; Mughal, M.; Korejo, N.A.; Kamboh, R.; et al. Flavonoid-rich foods (FRF): A promising nutraceutical approach against lifespan-shortening diseases. Iran. J. Basic Med. Sci. 2020, 23, 140–153. [Google Scholar]
  18. 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]
  19. Sun, Z.-G.; Li, Z.-N.; Zhang, J.-M.; Hou, X.-Y.; Yeh, S.M.; Ming, X. Recent Developments of Flavonoids with Various Activities. Curr. Top. Med. Chem. 2022, 22, 305–329. [Google Scholar] [CrossRef]
  20. Falcone Ferreyra, M.L.; Rius, S.P.; Casati, P. Flavonoids: Biosynthesis, biological functions, and biotechnological applications. Front. Plant Sci. 2012, 3, 34352. [Google Scholar] [CrossRef]
  21. Cushnie, T.P.T.; Lamb, A.J. Antimicrobial activity of flavonoids. Int. J. Antimicrob. Agents 2005, 26, 343–356. [Google Scholar] [CrossRef] [PubMed]
  22. Liang, Q.; Chen, H.; Zhou, X.; Deng, Q.; Hu, E.; Zhao, C.; Gong, X. Optimized microwave-assistant extraction combined ultrasonic pretreatment of flavonoids from Periploca forrestii Schltr. and evaluation of its anti-allergic activity. Electrophoresis 2017, 38, 1113–1121. [Google Scholar] [CrossRef] [PubMed]
  23. Kawai, M.; Hirano, T.; Higa, S.; Arimitsu, J.; Maruta, M.; Kuwahara, Y.; Ohkawara, T.; Hagihara, K.; Yamadori, T.; Yoshihito, S.; et al. Fla-vonoids and Related Compounds as Anti-Allergic Substances. Allergol. Int. 2007, 56, 113. [Google Scholar] [CrossRef] [PubMed]
  24. Brasiel, P.G.d.A.; Guimarães, F.V.; Rodrigues, P.M.; Bou-Habib, D.C.; Carvalho, V.d.F. Therapeutic Efficacy of Flavonoids in Allergies: A Systematic Review of Randomized Controlled Trials. J. Immunol. Res. 2022, 2022, 8191253. [Google Scholar] [CrossRef]
  25. Rakha, A.; Umar, N.; Rabail, R.; Butt, M.S.; Kieliszek, M.; Hassoun, A.; Aadil, R.M. Anti-inflammatory and anti-allergic potential of dietary flavonoids: A review. Biomed. Pharmacother. 2022, 156, 113945. [Google Scholar] [CrossRef] [PubMed]
  26. Shukitt-Hale, B.; Galli, R.L.; Meterko, V.; Carey, A.; Bielinski, D.F.; McGhie, T.; Joseph, J.A. Dietary supplementation with fruit poly-phenolics ameliorates age-related deficits in behavior and neuronal markers of inflammation and oxidative stress. Age 2005, 27, 49–57. [Google Scholar] [CrossRef] [PubMed]
  27. Serafini, M.; Peluso, I.; Raguzzini, A. Flavonoids as anti-inflammatory agents. Proc. Nutr. Soc. 2010, 69, 273–278. [Google Scholar] [CrossRef]
  28. Maleki, S.J.; Crespo, J.F.; Cabanillas, B. Anti-inflammatory effects of flavonoids. Food Chem. 2019, 299, 125124. [Google Scholar] [CrossRef]
  29. Al-Khayri, J.M.; Sahana, G.R.; Nagella, P.; Joseph, B.V.; Alessa, F.M.; Al-Mssallem, M.Q. Flavonoids as Potential Anti-Inflammatory Molecules: A Review. Molecules 2022, 27, 2901. [Google Scholar] [CrossRef]
  30. Ginwala, R.; Bhavsar, R.; Chigbu, D.G.I.; Jain, P.; Khan, Z.K. Potential role of flavonoids in treating chronic inflammatory diseases with a special focus on the anti-inflammatory activity of apigenin. Antioxidants 2019, 8, 35. [Google Scholar] [CrossRef] [PubMed]
  31. Cerqueira, F.; Cordeiro-Da-Silva, A.; Araújo, N.; Cidade, H.; Kijjoa, A.; Nascimento, M.S.J. Inhibition of lymphocyte proliferation by prenylated flavones: Artelastin as a potent inhibitor. Life Sci. 2003, 73, 2321–2334. [Google Scholar] [CrossRef]
  32. Cerqueira, F.; Cidade, H.; van Ufford, L.; Beukelman, C.; Kijjoa, A.; Nascimento, M.S.J. The natural prenylated flavone artelastin is an inhibitor of ROS and NO production. Int. Immunopharmacol. 2008, 8, 597–602. [Google Scholar] [CrossRef] [PubMed]
  33. Lncap, P.; Hela, C.H.; Horta, B.; Freitas-Silva, J.; Silva, J.; Dias, F.; Lu, A.; Medeiros, R.; Cidade, H.; Pinto, M.; et al. Antitumor Effect of Chalcone Derivatives against Human Macrophage Functions. Molecules 2023, 28, 2159. [Google Scholar]
  34. Kopustinskiene, D.M.; Jakstas, V.; Savickas, A.; Bernatoniene, J. Flavonoids as anticancer agents. Nutrients 2020, 12, 457. [Google Scholar] [CrossRef] [PubMed]
  35. Tuli, H.S.; Garg, V.K.; Bhushan, S.; Uttam, V.; Sharma, U.; Jain, A.; Sak, K.; Yadav, V.; Lorenzo, J.M.; Dhama, K.; et al. Natural flavonoids exhibit potent anticancer activity by targeting microRNAs in cancer: A signature step hinting towards clinical perfection. Transl. Oncol. 2023, 27, 101596. [Google Scholar] [CrossRef] [PubMed]
  36. de Luna, F.C.F.; Ferreira, W.A.S.; Casseb, S.M.M.; de Oliveira, E.H.C. Anticancer Potential of Flavonoids: An Overview with an Emphasis on Tangeretin. Pharmaceuticals 2023, 16, 1229. [Google Scholar] [CrossRef] [PubMed]
  37. Lúcio, M.; Giannino, N.; Barreira, S.; Catita, J.; Gonçalves, H.; Ribeiro, A.; Fernandes, E.; Carvalho, I.; Pinho, H.; Cerqueira, F.; et al. Nanostructured Lipid Carriers Enriched Hydrogels for Skin Topical Administration of Quercetin and Omega-3 Fatty Acid. Pharmaceutics 2023, 15, 2078. [Google Scholar] [CrossRef] [PubMed]
  38. Al Aboody, M.S.; Mickymaray, S. Anti-fungal efficacy and mechanisms of flavonoids. Antibiotics 2020, 9, 45. [Google Scholar] [CrossRef]
  39. Shamsudin, N.F.; Ahmed, Q.U.; Mahmood, S.; Shah, S.A.A.; Khatib, A.; Mukhtar, S.; Alsharif, M.; Parveenn, H.; Zakaria, Z.A. Anti-bacterial Effects of Flavonoids and Their Structure-Activity Relationship Study: A Comparative Interpretation. Molecules 2022, 27, 1149. [Google Scholar] [CrossRef]
  40. Mead, J.R.; McNair, N. Antiparasitic activity of flavonoids and isoflavones against Cryptosporidium parvum and Encephalitozoon intestinalis. FEMS Microbiol. Lett. 2006, 259, 153–157. [Google Scholar] [CrossRef]
  41. Górniak, I.; Bartoszewski, R.; Króliczewski, J. Comprehensive review of antimicrobial activities of plant flavonoids. Phytochem. Rev. 2018, 18, 241–272. [Google Scholar] [CrossRef]
  42. Wang, J.-F.; Liu, S.-S.; Song, Z.-Q.; Xu, T.-C.; Liu, C.-S.; Hou, Y.-G.; Huang, R.; Wu, S.-H. Naturally Occurring Flavonoids and Isoflavonoids and Their Microbial Transformation: A Review. Molecules 2020, 25, 5112. [Google Scholar] [CrossRef] [PubMed]
  43. Parshikov, I.A.; Sutherland, J.B. Biotransformation of Steroids and Flavonoids by Cultures of Aspergillus niger. Appl. Biochem. Biotechnol. 2015, 176, 903–923. [Google Scholar] [CrossRef] [PubMed]
  44. Nguyen, W.; Grigori, L.; Just, E.; Santos, C.; Seleem, D. The in vivo anti-Candida albicans activity of flavonoids. J. Oral Biosci. 2021, 63, 120–128. [Google Scholar] [CrossRef]
  45. Gontijo, V.S.; dos Santos, M.H.; Viegas, C., Jr. Biological and Chemical Aspects of Natural Biflavonoids from Plants: A Brief Review. Mini-Rev. Med. Chem. 2017, 17, 834–862. [Google Scholar] [CrossRef] [PubMed]
  46. Menezes, J.C.; Campos, V.R. Natural biflavonoids as potential therapeutic agents against microbial diseases. Sci. Total Environ. 2021, 769, 145168. [Google Scholar] [CrossRef]
  47. Oliveira, T.T.D.; Silva, R.R.D.; Dornas, W.C.A.; Nagem, T.J. Flavonóides e Aterosclerose. Biologia 2010, 42, 49–54. [Google Scholar]
  48. Canard, B.; Figadère, B.; Guillemot, J.; Claude Nour, M. Biflavonoids of Dacrydium balansae with potent inhibitory activity on dengue 2 NS5 polymerase. Planta Medica 2012, 78, 672–677. [Google Scholar]
  49. Konziase, B. Protective activity of biflavanones from Garcinia kola against Plasmodium infection. J. Ethnopharmacol. 2015, 172, 214–218. [Google Scholar] [CrossRef]
  50. Lopes Andrade, A.W.; Dias Ribeiro Figueiredo, D.; TorequlIslam, M.; Viana Nunes, A.M.; da Conceição Machado, K.; da Conceição Machado, K.; Uddin, S.J.; Shilpi, J.A.; Rouf, R.; Carvalho Melo-Cavalcante, A.A.; et al. Toxicological evaluation of the biflavonoid, agathisflavone in albino Swiss mice. Biomed. Pharmacother. 2019, 110, 68–73. [Google Scholar] [CrossRef]
  51. Antia, B.S.; Pansanit, A.; Ekpa, O.D.; Ekpe, U.J.; Mahidol, C.; Kittakoop, P. α-Glucosidase inhibitory, aromatase inhibitory, and antiplasmodial activities of a biflavonoid gb1 from garcinia kola stem bark. Planta Medica 2010, 76, 276–277. [Google Scholar] [CrossRef]
  52. Wang, T.Y.; Li, Q.; Bi, K.S. Bioactive flavonoids in medicinal plants: Structure, activity and biological fate. Asian J. Pharm. Sci. 2018, 13, 12–23. [Google Scholar] [CrossRef]
  53. Kumar, S.; Pandey, A.K. Chemistry and Biological Activities of Flavonoids: An Overview. Sci. World J. 2013, 2013, 162750. [Google Scholar] [CrossRef] [PubMed]
  54. Lee, H.J.; Kerr, R.A.; Korshavn, K.J.; Lee, J.; Kang, J.; Ramamoorthy, A.; Ruotolo, B.T.; Lim, M.H. Effects of hydroxyl group variations on a flavonoid backbone toward modulation of metal-free and metal-induced amyloid-β aggregation. Inorg. Chem. Front. 2016, 3, 381–392. [Google Scholar] [CrossRef]
  55. Sagrera, G.; Bertucci, A.; Vazquez, A.; Seoane, G. Synthesis and antifungal activities of natural and synthetic biflavonoids. Bioorganic Med. Chem. 2011, 19, 3060–3073. [Google Scholar] [CrossRef] [PubMed]
  56. Kpabi, I.; Munsch, T.; Agban, A.; Théry-Koné, I.; Dorat, J.; Boudesocque-Delaye, L.; Delaye, P.-O.; Neveu, C.; Lanoue, A.; Enguehard-Gueiffier, C. Cassia sieberiana root bark used in traditional medicine in Togo: Anthelmintic property against Haemonchus contortus and tannins composition. S. Afr. J. Bot. 2022, 151, 549–558. [Google Scholar] [CrossRef]
  57. Chaves, O.A.; Lima, C.R.; Fintelman-Rodrigues, N.; Sacramento, C.Q.; de Freitas, C.S.; Vazquez, L.; Temerozo, J.R.; Rocha, M.E.N.; Dias, S.S.G.; Carels, N.; et al. Agathisflavone, a natural biflavonoid that inhibits SARS-CoV-2 replication by targeting its proteases. Int. J. Biol. Macromol. 2022, 222, 1015–1026. [Google Scholar] [CrossRef] [PubMed]
  58. Šamec, D.; Karalija, E.; Dahija, S.; Hassan, S.T.S. Biflavonoids: Important Contributions to the Health Benefits of Ginkgo (Ginkgo biloba L.). Plants 2022, 11, 1381. [Google Scholar] [CrossRef]
  59. Nazerian, Y.; Ghasemi, M.; Yassaghi, Y.; Nazerian, A. Role of SARS-CoV-2-induced cytokine storm in multi-organ failure: Molecular pathways and potential therapeutic options. Int. Immunopharmacol. 2022, 113, 109428. [Google Scholar] [CrossRef] [PubMed]
  60. Bhalerao, A.; Raut, S.; Noorani, B.; Mancuso, S.; Cucullo, L. Molecular mechanisms of multi-organ failure in COVID-19 and potential of stem cell therapy. Cells 2021, 10, 2878. [Google Scholar] [CrossRef]
  61. Hartini, Y.; Saputra, B.; Wahono, B.; Auw, Z.; Indayani, F.; Adelya, L.; Namba, G.; Hariono, M. Biflavonoid as potential 3-chymotrypsin-like protease (3CLpro) inhibitor of SARS-Coronavirus. Results Chem. 2021, 3, 100087. [Google Scholar] [CrossRef]
  62. Miki, K.; Nagai, T.; Suzuki, K.; Tsujimura, R.; Koyama, K.; Kinoshita, K.; Furuhata, K.; Yamada, H.; Takahashi, K. Anti-influenza virus activity of biflavonoids. Bioorganic Med. Chem. Lett. 2007, 17, 772–775. [Google Scholar] [CrossRef]
  63. Miki, K.; Nagai, T.; Nakamura, T.; Tuji, M.; Koyama, K.; Kinoshita, K.; Furuhata, K.; Yamada, H.; Takahashi, K. Synthesis and Evaluation of Influenza Virus Sialidase Inhibitory Activity of Hinokiflavone-Sialic Acid Conjugates. Heterocycles 2008, 75, 879. [Google Scholar] [CrossRef]
  64. De Freitas, C.S.; Rocha, M.E.N.; Sacramento, C.Q.; Marttorelli, A.; Ferreira, A.C.; Rocha, N.; de Oliveira, A.C.; de Gomes, A.M.O.; dos Santos, P.S.; de Silva, E.D.O.; et al. Agathisflavone, a Biflavonoid from Anacardium occidentale L., Inhibits Influenza Virus Neuraminidase. Curr. Top. Med. Chem. 2020, 20, 111–120. [Google Scholar] [CrossRef]
  65. Lin, Y.-M.; Anderson, H.; Flavin, M.T.; Pai, Y.-H.S.; Mata-Greenwood, E.; Pengsuparp, T.; Pezzuto, J.M.; Schinazi, R.F.; Hughes, S.H.; Chen, F.-C. In Vitro Anti-HIV Activity of Biflavonoids Isolated from Rhus succedanea and Garcinia multiflora. J. Nat. Prod. 1997, 60, 884–888. [Google Scholar] [CrossRef] [PubMed]
  66. Ito, C.; Itoigawa, M.; Miyamoto, Y.; Rao, K.S.; Takayasu, J.; Okuda, Y.; Mukainaka, T.; Tokuda, H.; Nishino, H.; Furukawa, H. A new biflavonoid from Calophyllum panciflorum with antitumor-promoting activity. J. Nat. Prod. 1999, 62, 1668–1671. [Google Scholar] [CrossRef] [PubMed]
  67. Zakaryan, H.; Arabyan, E.; Oo, A.; Zandi, K. Flavonoids: Promising natural compounds against viral infections. Arch. Virol. 2017, 162, 2539–2551. [Google Scholar] [CrossRef] [PubMed]
  68. Coulerie, P.; Nour, M.; Maciuk, A.; Eydoux, C.; Guillemot, J.C.; Lebouvier, N.; Hnawia, E.; Leblanc, K.; Lewin, G.; Canard, B.; et al. Structure-Activity Relationship Study of Biflavonoids on the Dengue Virus Polymerase DENV-NS5 RdRp. Planta Medica 2013, 79, 1313–1318. [Google Scholar] [CrossRef] [PubMed]
  69. De Sousa, L.R.F.; Wu, H.; Nebo, L.; Fernandes, J.B.; das Graças Fernandes da Silva, M.F.; Kiefer, W.; Kanitz, M.; Bodem, J.; Diederich, W.E.; Schirmeister, T.; et al. Flavonoids as noncompetitive inhibitors of Dengue virus NS2B-NS3 protease: Inhibition kinetics and docking studies. Bioorg. Med. Chem. 2015, 23, 466–470. [Google Scholar] [CrossRef] [PubMed]
  70. Lin, Y.-M.; Zembower, D.E.; Flavin, M.T.; Schure, R.M.; Anderson, H.M.; Korba, B.E.; Chen, F.-C. Robustaflavone, a naturally occurring biflavanoid, is a potent non-nucleoside inhibitor of hepatitis B virus replication in vitro. Bioorganic Med. Chem. Lett. 1997, 7, 2325–2328. [Google Scholar] [CrossRef]
  71. Yang, G.; Chen, D. Biflavanones, flavonoids, and coumarins from the roots of Stellera chamaejasme and their antiviral effect on hepatits B virus. Chem. Biodivers. 2008, 5, 1419–1424. [Google Scholar] [CrossRef] [PubMed]
  72. Lee, W.-P.; Liao, S.-X.; Huang, Y.-H.; Hou, M.-C.; Lan, K.-H. Inhibitory Effects of Amentoflavone and Orobol on Daclatasvir-Induced Resistance-Associated Variants of Hepatitis C Virus. Am. J. Chin. Med. 2018, 46, 835–852. [Google Scholar] [CrossRef] [PubMed]
  73. Li, F.; Song, X.; Su, G.; Wang, Y.; Wang, Z.; Jia, J.; Qing, S.; Huang, L.; Wang, Y.; Zheng, K.; et al. Amentoflavone Inhibits HSV-1 and ACV-Resistant Strain Infection by Suppressing Viral Early Infection. Viruses 2019, 11, 466. [Google Scholar] [CrossRef]
  74. Fidelis, Q.C.; Ribeiro, T.A.; Araújo, M.F.; de Carvalho, M.G. Ouratea genus: Chemical and pharmacological aspects. Rev. Bras. Farm. 2014, 24, 1–19. [Google Scholar] [CrossRef]
  75. Boff, L.; Silva, I.T.; Argenta, D.F.; Farias, L.M.; Alvarenga, L.F.; Pádua, R.M.; Braga, F.C.; Leite, J.P.V.; Kratz, J.M.; Simões, C.M.O. Strychnos pseudoquina A. St. Hil.: A Brazilian medicinal plant with promising in vitro antiherpes activity. J. Appl. Microbiol. 2016, 121, 1519–1529. [Google Scholar] [CrossRef] [PubMed]
  76. Hayashi, K.; Hayashi, T.; Morita, N. Mechanism of action of the antiherpesvirus biflavone ginkgetin. Antimicrob. Agents Chemother. 1992, 36, 1890–1893. [Google Scholar] [CrossRef]
  77. Wilsky, S.; Sobotta, K.; Wiesener, N.; Pilas, J.; Althof, N.; Munder, T.; Wutzler, P.; Henke, A. Inhibition of fatty acid synthase by amentoflavone reduces coxsackievirus B3 replication. Arch. Virol. 2012, 157, 259–269. [Google Scholar] [CrossRef]
  78. Salas, M.P.; Céliz, G.; Geronazzo, H.; Daz, M.; Resnik, S.L. Antifungal activity of natural and enzymatically-modified flavonoids isolated from citrus species. Food Chem. 2011, 124, 1411. [Google Scholar] [CrossRef]
  79. Arif, T.; Bhosale, J.; Kumar, N.; Mandal, T.; Bendre, R.; Lavekar, G.; Dabur, R. Natural products antifungal agents derived from plants. J. Asian Nat. Prod. Res. 2009, 11, 621–638. [Google Scholar] [CrossRef]
  80. Bagla, V.P.; McGaw, L.J.; Elgorashi, E.E.; Eloff, J.N. Antimicrobial activity, toxicity and selectivity index of two biflavonoids and a flavone isolated from Podocarpus henkelii (Podocarpaceae) leaves. BMC Complement. Altern. Med. 2014, 14, 383. [Google Scholar] [CrossRef]
  81. Jung, H.J.; Park, K.; Lee, I.-S.; Kim, H.S.; Yeo, S.H.; Woo, E.R.; Lee, D.G. S-Phase Accumulation of Candida albicans by anticandidal effect of Amentoflavone Isolated from Selaginella tamariscina. Biol. Pharm. Bull. 2007, 30, 1969–1971. [Google Scholar] [CrossRef]
  82. Krauze-Baranowska, M.; Wiwart, M. Antifungal Activity of Biflavones from Taxus baccata and Ginkgo biloba. Z. Naturforschung Sect. C J. Biosci. 2003, 58, 65–69. [Google Scholar] [CrossRef] [PubMed]
  83. Gao, M.; Wang, H.; Zhu, L. Quercetin Assists Fluconazole to Inhibit Biofilm Formations of Fluconazole-Resistant Candida Albicans in In Vitro and In Vivo Antifungal Managements of Vulvovaginal Candidiasis. Cell. Physiol. Biochem. 2016, 40, 727–742. [Google Scholar] [CrossRef] [PubMed]
  84. Gonçalez, E.; Felicio, J.; Pinto, M. Biflavonoids inhibit the production of aflatoxin by Aspergillus flavus. Braz. J. Med. Biol. Res. 2001, 34, 1453–1456. [Google Scholar] [CrossRef] [PubMed]
  85. de Freitas, A.L.D.; Kaplum, V.; Rossi, D.C.P.; da Silva, L.B.R.; Melhem, M.d.S.C.; 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] [PubMed]
  86. Chow, E.W.L.; Pang, L.M.; Wang, Y. Impact of the host microbiota on fungal infections: New possibilities for intervention? Adv. Drug Deliv. Rev. 2023, 198, 114896. [Google Scholar] [CrossRef] [PubMed]
  87. Hyun, J.J.; Woo, S.S.; Yeo, S.H.; Hyun, S.K.; Lee, I.S.; Woo, E.R.; Dong, G.L. Antifungal effect of amentoflavone derived from Selaginella tamariscina. Arch. Pharm. Res. 2006, 29, 746–751. [Google Scholar]
  88. Martino, R.; Canale, F.; Sülsen, V.; Alonso, R.; Davicino, R.; Mattar, A.; Anesini, C.; Micalizzi, B. A Fraction containing kaempferol-3,4-dimethylether from Larrea divaricata Cav. Induces macrophage activation on mice infected with Candida albicans. Phytother. Res. 2014, 28, 917–924. [Google Scholar] [CrossRef]
  89. Gazoni, V.F.; Balogun, S.O.; Arunacham, K.; Oliveira, D.M.; Filho, V.C.; Lima, S.R.; Colodel, E.M.; Soares, I.M.; Ascêncio, S.D.; de Martins, D.T.O. Assessment of toxicity and differential antimicrobial activity of methanol extract of rhizome of Simaba ferruginea A. St.-Hil. and its isolate canthin-6-one. J. Ethnopharmacol. 2018, 223, 122–134. [Google Scholar] [CrossRef]
  90. Prevention C for DC and Parasites. About Parasites. 2023. Available online: https://www.cdc.gov/parasites/ (accessed on 13 June 2023).
  91. Gontijo, V.S.; Judice, W.A.; Codonho, B.; Pereira, I.O.; Assis, D.M.; Januário, J.P.; Caroselli, E.E.; Juliano, M.A.; Dosatti, A.d.C.; Marques, M.J.; et al. Leishmanicidal, antiproteolytic and antioxidant evaluation of natural biflavonoids isolated from Garcinia brasiliensis and their semisynthetic derivatives. Eur. J. Med. Chem. 2012, 58, 613–623. [Google Scholar] [CrossRef]
  92. Mercado-Camargo, J.; Cervantes-Ceballos, L.; Vivas-Reyes, R.; Pedretti, A.; Serrano-García, M.L.; Gómez-Estrada, H. Homology Modeling of Leishmanolysin (gp63) from Leishmania panamensis and Molecular Docking of Flavonoids. ACS Omega 2020, 5, 14741–14749. [Google Scholar] [CrossRef]
  93. Muhammad, A.; Anis, I.; Ali, Z.; Awadelkarim, S.; Khan, A.; Khalid, A.; Shah, M.R.; Galal, M.; Khan, I.A.; Iqbal Choudhary, M. Methylenebissantin: A rare methylene-bridged bisflavonoid from Dodonaea viscosa which inhibits Plasmodium falciparum enoyl-ACP reductase. Bioorg. Med. Chem. Lett. 2012, 22, 610–612. [Google Scholar] [CrossRef]
  94. Weniger, B.; Vonthron-Sénécheau, C.; Kaiser, M.; Brun, R.; Anton, R. Comparative antiplasmodial, leishmanicidal and antitrypa-nosomal activities of several biflavonoids. Phytomedicine 2006, 13, 176–180. [Google Scholar] [CrossRef]
  95. Kunert, O.; Swamy, R.C.; Kaiser, M.; Presser, A.; Buzzi, S.; Rao, A.A.; Schühly, W. Antiplasmodial and leishmanicidal activity of biflavonoids from Indian Selaginella bryopteris. Phytochem. Lett. 2008, 1, 171–174. [Google Scholar] [CrossRef]
  96. Florencio, M.; Tomás Nery, E.; Rosa, D.; Auxiliadora Nascimento Ribeiro, T.; de Brito Braz Moraes, J.; Araujo Zuma, A.; da Silva Trindade, J.D.A.; Dutra Barbosa da Rocha, R.F.; Decote-Ricardo, D.; Pinto-da-Silva, L.H.; et al. The effect of the biflavonoid 2″,3″-dihydroochnaflavone on Trypanosoma cruzi Y strain. Parasitol. Int. 2020, 79, 102180. [Google Scholar] [CrossRef] [PubMed]
  97. da Rocha, C.Q.; Queiroz, E.F.; Meira, C.S.; Moreira, D.R.M.; Soares, M.B.P.; Marcourt, L.; Vilegas, W.; Wolfender, J.-L. Dimeric Flavonoids from Arrabidaea brachypoda and Assessment of Their Anti-Trypanosoma cruzi Activity. J. Nat. Prod. 2014, 77, 1345–1350. [Google Scholar] [CrossRef] [PubMed]
  98. Boniface, P.K.; Ferreira, E.I. Flavonoids as efficient scaffolds: Recent trends for malaria, leishmaniasis, Chagas disease, and dengue. Phytother. Res. 2019, 33, 2473–2517. [Google Scholar] [CrossRef] [PubMed]
  99. Ichino, C.; Kiyohara, H.; Soonthornchareonnon, N.; Chuakul, W.; Ishiyama, A.; Sekiguchi, H.; Namatame, M.; Otoguro, K.; Omura, S.; Yamada, H. Antimalarial Activity of Biflavonoids from Ochna integerrima. Planta Medica 2006, 72, 611–614. [Google Scholar] [CrossRef] [PubMed]
  100. Lage, P.S.; Chávez-Fumagalli, M.A.; Mesquita, J.T.; Mata, L.M.; Fernandes, S.O.A.; Cardoso, V.N.; Soto, M.; Tavares, C.A.P.; Leite, J.P.V.; Tempone, A.G.; et al. Antileishmanial activity and evaluation of the mechanism of action of strychnobiflavone flavonoid isolated from Strychnos pseudoquina against Leishmania infantum. Parasitol. Res. 2015, 114, 4625–4635. [Google Scholar] [CrossRef] [PubMed]
  101. Rizk, Y.S.; Fischer, A.; Cunha, M.d.C.; Rodrigues, P.O.; Marques, M.C.S.; Matos, M.d.F.C.; Kadri, M.C.T.; Carollo, C.A.; de Arruda, C.C.P. In vitro activity of the hydroethanolic extract and biflavonoids isolated from Selaginella sellowii on Leishmania (Leishmania) amazonensis. Mem. Inst. Oswaldo Cruz. 2014, 109, 1050–1056. [Google Scholar] [CrossRef] [PubMed]
  102. Rocha, V.P.C.; da Rocha, C.Q.; Queiroz, E.F.; Marcourt, L.; Vilegas, W.; Grimaldi, G.B.; Furrer, P.; Allémann, E.; Wolfender, J.-L.; Soares, M.B.P. Antileishmanial activity of dimeric flavonoids isolated from Arrabidaea brachypoda. Molecules 2018, 24, 1. [Google Scholar] [CrossRef] [PubMed]
  103. Zheng, W.; Sun, W.; Simeonov, A. Drug repurposing screens and synergistic drug-combinations for infectious diseases. Br. J. Pharmacol. 2017, 175, 181–191. [Google Scholar] [CrossRef]
  104. Thévenin, M.; Mouray, E.; Grellier, P.; Dubois, J. Facile formation of methylenebis(chalcone)s through unprecedented methylenation reaction. Application to antiparasitic and natural product synthesis. Eur. J. Org. Chem. 2014, 2014, 2986–2992. [Google Scholar] [CrossRef]
  105. Cane, H.P.C.A.; Saidi, N.; Yahya, M.; Darusman, D.; Erlidawati, E.; Safrida, S.; Musman, M. Macrophylloflavone: A New Biflavonoid from Garcinia macrophylla Mart. (Clusiaceae) for Antibacterial, Antioxidant, and Anti-Type 2 Diabetes Mellitus Activities. Sci. World J. 2020, 2020, 2983129. [Google Scholar] [CrossRef]
  106. Lee, J.; Kim, M.; Jeong, S.E.; Park, H.Y.; Jeon, C.O.; Park, W. Amentoflavone, a novel cyanobacterial killing agent from Selaginella tamariscina. J. Hazard. Mater. 2020, 384, 121312. [Google Scholar] [CrossRef]
  107. Xu, H.-X.; Mughal, S.; Taiwo, O.; Lee, S.F. Isolation and characterization of an antibacterial biflavonoid from an African chewing stick Garcinia kola Heckel (Clusiaceae). J. Ethnopharmacol. 2013, 147, 497–502. [Google Scholar] [CrossRef] [PubMed]
  108. Linden, M.; Brinckmann, C.; Feuereisen, M.M.; Schieber, A. Effects of structural differences on the antibacterial activity of biflavonoids from fruits of the Brazilian peppertree (Schinus terebinthifolius Raddi). Food Res. Int. 2020, 133, 109134. [Google Scholar] [CrossRef] [PubMed]
  109. Nandu, T.G.; Subramenium, G.A.; Shiburaj, S.; Viszwapriya, D.; Iyer, P.M.; Balamurugan, K.; Rameshkumar, K.B.; Pandian, S.K. Fukugiside, a biflavonoid from Garcinia travancorica inhibits biofilm formation of Streptococcus pyogenes and its associated virulence factors. J. Med. Microbiol. 2018, 67, 1391–1401. [Google Scholar] [CrossRef]
  110. Mkounga, P.; Fomum, Z.T.; Meyer, M.; Bodo, B.; Nkengfack, A.E. Globulixanthone F. a new polyoxygenated xanthone with an isoprenoid group and two antimicrobial biflavonoids from the stem bark of Symphonia globulifera. Nat. Prod. Commun. 2009, 4, 803–808. [Google Scholar] [CrossRef]
  111. Bitchagno, G.T.M.; Tankeo, S.B.; Tsopmo, A.; Mpetga, J.D.S.; Tchinda, A.T.; Fobofou, S.A.T.; Nkuete, A.H.L.; Wessjohann, L.A.; Kuete, V.; Tane, P. Ericoside, a new antibacterial biflavonoid from Erica mannii (Ericaceae). Fitoterapia 2016, 109, 206–211. [Google Scholar] [CrossRef]
  112. Makhafola, T.J.; Samuel, B.B.; Elgorashi, E.E.; Eloff, J.N. Ochnaflavone and Ochnaflavone 7-O-Methyl Ether two Antibacterial Biflavonoids from Ochna pretoriensis (Ochnaceae). Nat. Prod. Commun. 2012, 7, 1601–1604. [Google Scholar] [CrossRef] [PubMed]
  113. Araújo, M.G.d.F.; Pacífico, M.; Vilegas, W.; Dos Santos, L.C.; Icely, P.A.; Miró, M.S.; Scarpa, M.V.C.; Bauab, T.M.; Sotomayor, C.E. Evaluation of Syngonanthus nitens (Bong.) Ruhl. extract as antifungal and in treatment of vulvovaginal candidiasis. Med. Mycol. 2013, 51, 673–682. [Google Scholar] [CrossRef] [PubMed]
  114. Simonetti, G.; Santamaria, A.R.; D’Auria, F.D.; Mulinacci, N.; Innocenti, M.; Cecchini, F.; Pericolini, E.; Gabrielli, E.; Panella, S.; Antonacci, D.; et al. Evaluation of anti-Candida activity of Vitis vinifera L. seed extracts obtained from wine and table cultivars. BioMed Res. Int. 2014, 2014, 127021. [Google Scholar] [CrossRef] [PubMed]
  115. Seleem, D.; Benso, B.; Noguti, J.; Pardi, V.; Murata, R.M. In vitro and in vivo antifungal activity of lichochalcone-A against candida albicans biofilms. PLoS ONE 2016, 11, e0157188. [Google Scholar] [CrossRef]
  116. Marín, C.; Ramírez-Macías, I.; López-Céspedes, A.; Olmo, F.; Villegas, N.; Díaz, J.G.; Rosales, M.J.; Gutiérrez-Sánchez, R.; Sánchez-Moreno, M. In Vitro and in Vivo Trypanocidal Activity of Flavonoids from Delphinium staphisagria against Chagas Disease. J. Nat. Prod. 2011, 74, 744–750. [Google Scholar] [CrossRef]
  117. Gervazoni, L.F.O.; Gonçalves-Ozório, G.; Almeida-Amaral, E.E. 2′-Hydroxyflavanone activity in vitro and in vivo against wild-type and antimony-resistant Leishmania amazonensis. PLoS Neglected Trop. Dis. 2018, 12, e0006930. [Google Scholar] [CrossRef]
  118. Pereira, V.R.D.; Junior, I.J.A.; da Silveira, L.S.; Geraldo, R.B.; Pinto, P.d.F.; Teixeira, F.S.; Salvadori, M.C.; Silva, M.P.; Alves, L.A.; Capriles, P.V.S.Z. In Vitro and in Vivo Antischistosomal Activities of Chalcones. Chem. Biodivers. 2018, 15, e1800398. [Google Scholar] [CrossRef]
  119. Zhi, H.-J.; Zhu, H.-Y.; Zhang, Y.-Y.; Lu, Y.; Li, H.; Chen, D.-F. In vivo effect of quantified flavonoids-enriched extract of Scutellaria baicalensis root on acute lung injury induced by influenza A virus. Phytomedicine 2019, 57, 105–116. [Google Scholar] [CrossRef]
  120. Ma, Y.; Wang, L.; Lu, A.; Xue, W. Synthesis and Biological Activity of Novel Oxazinyl Flavonoids as Antiviral and Anti-Phytopathogenic Fungus Agents. Molecules 2022, 27, 6875. [Google Scholar] [CrossRef]
  121. Ferrazzano, G.F.; Cantile, T.; Roberto, L.; Ingenito, A.; Catania, M.R.; Roscetto, E.; Palumbo, G.; Zarrelli, A.; Pollio, A. Determination of the in vitro and in vivo antimicrobial activity on salivary streptococci and lactobacilli and chemical characterisation of the phenolic content of a plantago lanceolata infusion. BioMed Res. Int. 2015, 2015, 286817. [Google Scholar] [CrossRef]
  122. Machado, G.H.A.; Marques, T.R.; de Carvalho, T.C.L.; Duarte, A.C.; de Oliveira, F.C.; Gonçalves, M.C.; Piccoli, R.H.; Corrêa, A.D. Antibacterial activity and in vivo wound healing potential of phenolic extracts from jaboticaba skin. Chem. Biol. Drug Des. 2018, 92, 1333–1343. [Google Scholar] [CrossRef] [PubMed]
  123. Wink, M. Plant Secondary Metabolism: Diversity, Function and its Evolution. Nat. Prod. Commun. 2010, 1, 9–12. [Google Scholar] [CrossRef]
Figure 1. Main targets of dimeric flavonoids against fungi [46].
Figure 1. Main targets of dimeric flavonoids against fungi [46].
Compounds 04 00011 g001
Figure 2. Main targets of dimeric flavonoids against bacteria [46].
Figure 2. Main targets of dimeric flavonoids against bacteria [46].
Compounds 04 00011 g002
Table 1. Antiviral activity in vitro of dimeric flavonoids.
Table 1. Antiviral activity in vitro of dimeric flavonoids.
VirusCompoundConclusionsLiterature
Reference
SARS-CoV-2AgathisflavoneReplication block by Mpro protease inhibition.
Reduces TNF-α (tumor necrosis factor-alpha) levels in infected cells.
[57]
5,6,7-trihydroxy-2-phenyl-4H-chromen-4-oneInterruption of viral RNA replication by blocking the Chymotrypsinlike protease[61]
InfluenzaGinkgetinInhibits sialidase activity.[62]
Hinokiflavone [63]
AgathisflavoneInhibits neuraminidase activity.[64]
Human
immunodeficiency
virus (HIV)
Robustaflavone and hinokiflavoneBlocks reverse transcriptase activity.[65]
MorelloflavoneActivity against HIV-1.
Epstein−Barr
virus (EBV)
Garcinianin and talbotaflavoneInhibit the 12-O-tetradecanoylphorbol-13-acetate-(TPA)-induced Epstein−Barr virus early antigen (EBV-EA) activation in Raji cells.[66]
3′, 4′, 5, 7-tetrahydroxyflavoneInhibit the reactivation of EBV through early genes (Zta and Rta) block and interfering with binding of transcription factor Sp1.[67]
Dengue virusHinokiflavone and AmentoflavoneInhibits RNA-dependant RNA polymerase (DV-NS5 RdRp).[48,68]
Sotetsuflavone and
robustaflavone
Inhibits dengue virus NS5 RNA dependent RNA polymerase.[68]
AgathisflavoneInhibits NS2B-NS3 protease.[69]
Podocarpusflavone AInhibits the DV-NS5.[48]
Hepatitis B virus
(HBV)
RobustaflavoneInhibits the DNA polymerase.[70]
Sikokianin AReduces HBsAg secretion.[71]
Hepatitis C virus
(HBC)
AmentoflavoneDeregulates all the virus life cycle, including viral entry, replication, and translation. Inhibitor of NS5A.[72]
Herpes simplex
virus (HSV)
AmentoflavoneAffects the expression of UL52 (early gene), UL54 (immediate-early gene) and UL27 (late gene). More active against HSV-1.[73]
AgathisflavoneActive against HSV-1 and HSV-2.[74]
StrychnobiflavoneInterferes with the initial stages of viral infection and reduces HSV-1 protein expression.[75]
GinkgetinActivity against HSV-1 and HSV-2. Inhibits the transcription step in the protein synthesis of HSV-infected cells.[76]
Coxsackievirus B3AmentoflavoneInhibits fatty acid synthase.[77]
Abbreviations: TNF-α, tumor necrosis factor-α; RNA, ribonucleic acid; HIV, Human immunodeficiency virus; DNA, deoxyribonucleic acidHBV, Hepatitis B virus; HBC, Hepatitis C virus; HSV, Herpes simplex virus.
Table 2. Dimeric flavonoid antifungal activity.
Table 2. Dimeric flavonoid antifungal activity.
FungusCompoundConclusionsLiterature
Reference
Candida
albicans
AmentoflavoneFungistatic. Affects cell cycle progress during S-phase.[81]
Interrupts dimorphic transition.
Enhances the intracellular trehalose level, which induces a stress response in fungal cells.
[87]
QuercetinInhibits fungal adherence and biofilm formation.[83]
Kaempferol-3,40-dimethyletherActivates macrophages and increases lysosomal activity.[88]
Kaempferol, canthin-
6-one, and morin)
Cell membrane damage.[89]
ProanthocyanidinInhibits proliferation and dispersion cells from pre-formed biofilms.[85]
Aspergillus
flavus
Amentoflavone, 7,7″-
Dimethoxyagastisflavone, 6,6″-bigenkwanin, and tetramethoxy-6,6″-bigenkwanin
Reduces the production of aflatoxin B1 (AFB1) and B2 (AFB2).[84]
Aspergillus
fumigatus
IsoginkgetinGrowth inhibition.[80]
Cryptococcus neoformasIsoginkgetinGrowth inhibition.[80]
Podocarpusflavone
Fusarium
culmorum
BilobetinInhibits the growth of germinating tubes.[82]
Cladosporium
oxysporum
BilobetinInhibits the growth of germinating tubes.[82]
Alternaria
alternata
Ginkgetin and 7-O-methylamentoflavoneInhibits the growth of fungal spores.
Small changes in the cell wall.
[82]
Abbreviations: AFB1, Aflatoxin B1; AFB2, Aflatoxin B2.
Table 3. Activity of dimeric flavonoids against protozoa.
Table 3. Activity of dimeric flavonoids against protozoa.
ProtozoaCompoundConclusionsLiterature Reference
Plasmodium falciparum3″,4′,4‴,5,5″,7,7″-heptahydroxy-3,8-biflavanoneInhibition of α-glucosidase and aromatase.[51]
LanaroflavoneMechanism of action unknown.[94]
7,4′,7″-tri-O-methylamentoflavoneMechanism of action unknown.[95]
MethylenebissantinInhibits enoyl-ACP reductase.[93]
3,3″-di(7,4″-dihydroxyflavanone-3-yl)Mechanism of action unknown.[98,99]
Leishmania panamensisLanaroflavone
Podocarpusflavone A
Podocarpusflavone B
Amentoflavona
Interact with Glycoprotein 63.[92]
Leishmania
infantum
StrychnobiflavoneCauses depolarization of parasitic mitochondria.[100]
AmentoflavoneActivity against intracellular amastigotes.[95]
Leishmania donovani2,3-DihydrohinokiflavoneTested on axenic amastigotes.[91]
Leishmania mexicanaMorelloflavone and
Acetate
Interact with recombinant cysteine protease type 2.8[91]
Leishmania amazonensisAmentoflavone and
robustaflavone
Effective antioxidant activity by increasing nitric oxide (NO) production in macrophages. Strong activity against promastigote and amastigote forms.[91]
7-O-methyl ochnaflavoneActivity against promastigote forms.[101]
BrachydinReduces the number of amastigotes and infected macrophages. Presents a synergic effect with amphotericin B. Also showed ability to induce damage in Golgi apparatus by accumulation of vesicles.[102]
Trypanosoma cruzi2″,3″-DihydroochnaflavoneKills approximately 62% of amastigote forms and 100% of trypomastigotes in infected murine macrophages. The mechanism is unknown. It is also able to inhibit topoisomerase I and topoisomerase II-α, which may be the cause of mitochondrial alterations in the parasitic form.[96]
Brachydin B and CInhibits the parasite invasion and its intracellular multiplication in host cells, reducing parasitemia.[97]
Abbreviations: NO, nitric oxide.
Table 4. Activity of dimeric flavonoids in bacteria.
Table 4. Activity of dimeric flavonoids in bacteria.
BacteriaCompoundConclusionsLiterature Reference
Staphylococcus
aureus
7, 4′, 7″, 4‴-Tetramethoxy amenthoflavoneThe lipophilic nature of the molecules and the external porous peptide cell wall structure of Gram-positive bacteria determined their effect. In Gram-negative bacteria, growth inhibition is lower.[80]
Macrophylloflavone 18Inhibits nucleic acid synthesis, cytoplasmic membrane function, energy metabolism, and porins in cell membranes.[105]
IsoginkgetinGrowth inhibition.[80]
Podocarpusflavone—AMechanism of action unknown.[80]
ManniflavanoneMechanism of action unknown.[110]
Escherichia coliMacrophylloflavone 18Inhibits nucleic acid synthesis, cytoplasmic membrane function, energy metabolism, and porins in cell membranes.[105]
IsoginkgetinGrowth inhibition.[80]
EricosideMechanism of action unknown.[111]
Bacillus subtilis
and
Staphylococcus
carnosus
Agatisflavone 2, amentoflavone 1, and
Tetrahydroamentoflavone (THAF)
Inhibition of biofilm formation. Dimerization and a reduced C ring contribute to greater activity of the compounds.[108]
Streptococcus
pyogenes
FukugisideExhibited concentration-dependent biofilm inhibition by destabilizing the biofilm matrix and by inhibiting M proteins.[109]
Pseudomonas
aeruginosa
Ochnaflavone and ochnaflavone 7-O-methylether 15cMechanism of action unknown.[112]
Microcystis
aeruginosa
AmentoflavoneBacteria lose their round shape and eventually succumb completely. Affects the peptidoglycan layer and reduces pressure, which ends with the leaking of cell contents. Effects are dose-dependent.[106]
Enterococcus faecalisPodocarpusflavone—AMechanism of action unknown.[80]
ManniflavanoneMechanism of action unknown.[88]
IsoginkgetinGrowth inhibition.[80]
Ochnaflavone and ochnaflavone 7-O-methylether 15cMechanism of action unknown.[112]
Actinomyces
naeslundii, Porphyromonas gingivalis,
Streptococcus
mutans,
Streptococcus
mitis and
Streptococcus downeii
3″,4′,4‴,5,5″,7,7″-heptahydoxy-3-8″-biflavoneInhibition of glucan synthesis, glucose uptake and metabolism.
Induces bacterial aggregation.
[107]
Klebsiella
pneumoniae
EricosideMechanism of action unknown.[111]
Abbreviations: THAF, Tetrahydroamentoflavone.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lopes, I.; Campos, C.; Medeiros, R.; Cerqueira, F. Antimicrobial Activity of Dimeric Flavonoids. Compounds 2024, 4, 214-229. https://doi.org/10.3390/compounds4020011

AMA Style

Lopes I, Campos C, Medeiros R, Cerqueira F. Antimicrobial Activity of Dimeric Flavonoids. Compounds. 2024; 4(2):214-229. https://doi.org/10.3390/compounds4020011

Chicago/Turabian Style

Lopes, Inês, Carla Campos, Rui Medeiros, and Fátima Cerqueira. 2024. "Antimicrobial Activity of Dimeric Flavonoids" Compounds 4, no. 2: 214-229. https://doi.org/10.3390/compounds4020011

Article Metrics

Back to TopTop