Next Article in Journal
Impact of Various Essential Oils on the Development of Pathogens of the Fusarium Genus and on Health and Germination Parameters of Winter Wheat and Maize
Previous Article in Journal
Urethane Synthesis in the Presence of Organic Acid Catalysts—A Computational Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Product-Derived Phytochemicals for Influenza A Virus (H1N1) Prevention and Treatment

1
College of Pharmacy, Henan University of Chinese Medicine, Zhengzhou 450003, China
2
Foreign Language Education Department, Zhengzhou Shuqing Medical College, Zhengzhou 450064, China
3
Co-Construction Collaborative Innovation Center for Chinese Medicine and Respiratory Diseases by Henan & Education Ministry of China, Zhengzhou 450003, China
4
Academy of Chinese Medical Sciences, Henan University of Chinese Medicine, Zhengzhou, 450003, China
5
Department of Pharmacy, Henan Province Hospital of Traditional Chinese Medicine, Zhengzhou 450046, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(10), 2371; https://doi.org/10.3390/molecules29102371
Submission received: 8 March 2024 / Revised: 1 May 2024 / Accepted: 2 May 2024 / Published: 17 May 2024

Abstract

:
Influenza A (H1N1) viruses are prone to antigenic mutations and are more variable than other influenza viruses. Therefore, they have caused continuous harm to human public health since the pandemic in 2009 and in recent times. Influenza A (H1N1) can be prevented and treated in various ways, such as direct inhibition of the virus and regulation of human immunity. Among antiviral drugs, the use of natural products in treating influenza has a long history, and natural medicine has been widely considered the focus of development programs for new, safe anti-influenza drugs. In this paper, we focus on influenza A (H1N1) and summarize the natural product-derived phytochemicals for influenza A virus (H1N1) prevention and treatment, including marine natural products, flavonoids, alkaloids, terpenoids and their derivatives, phenols and their derivatives, polysaccharides, and derivatives of natural products for prevention and treatment of influenza A (H1N1) virus. We further discuss the toxicity and antiviral mechanism against influenza A (H1N1) as well as the druggability of natural products. We hope that this review will facilitate the study of the role of natural products against influenza A (H1N1) activity and provide a promising alternative for further anti-influenza A drug development.

1. Introduction

Influenza is a viral respiratory infection that causes acute febrile illness with associated myalgia, headache, and cough. It can result in significantly increased morbidity and mortality during an epidemic [1]. Highly pathogenic strains of influenza A virus have emerged unpredictably but repeatedly in recent history. Since 1918, the year of the “Spanish” influenza outbreak, this virus has caused 50 million deaths in this century [2]. Since then, the introduction of influenza A viruses from poultry or swine herds has led to four world pandemics, all causing high mortality, global health threats, and large economic losses [3,4]. Due to the rapid variation of influenza and the low vaccine penetration rate, antiviral drug therapy remains the core treatment for influenza. Anti-influenza virus drugs approved by the U.S. Food and Drug Administration (FDA) include M2 inhibitors, adamantanes (rimantadine and amantadine), NA inhibitors (NAI; peramivir, zanamivir, and oseltamivir), and, more recently, the cap-dependent endonuclease inhibitor targeting the PA polymerase subunit (baloxavir) [5,6,7]. However, adverse drug reactions and the emergence of resistant strains of the virus have made the development of safe and effective new antiviral drugs increasingly urgent for both therapeutic or prophylactic purposes.
Natural products refer to compounds or substances derived from natural resources such as animals, plants, microorganisms, minerals, etc., which are not artificially synthesized or modified. [8]. The use of natural medicine has a long history, and it has been widely considered the focus of development programs for new, safe anti-influenza drugs [9]. In this review, we aimed to provide a new perspective for the development of novel anti-H1N1 drugs. We conclude the review of newly discovered natural products with anti-influenza A (H1N1) activity with a call for more researchers to focus on natural products as potential drugs for the treatment of influenza A (H1N1).

2. Treatment Strategies against Influenza A (H1N1)

The influenza virus is a negative, single-stranded RNA virus belonging to the Orthomyxoviridae family [10]. Its genome is about 13.6 kb and consists of eight independent RNA fragments of different sizes, encoding 10 proteins, respectively: HA, NA, PA (RNA polymerase subunit PA), PB1 (RNA polymerase subunit PB1), PB2 (RNA polymerase subunit PB2), nuclear protein (NP), M (membrane proteins including M1 and M2, encoded by the same RNA fragment), and NS (non-structural proteins including N1 and N2, encoded by the same RNA fragment) [11] (Figure 1). According to the NP and membrane protein and their genetic characteristics, influenza virus can be divided into A, B, and C types [11]. Among them, influenza A viruses are more prone to antigenic mutations and are more variable than the other influenza viruses. Therefore, it is not only the most common and lethal type of influenza virus but also the main virus that causes seasonal or pandemic influenza. Furthermore, 18 different hemagglutinin (HA) subtypes (H1–H18) and 11 different neuraminidase (NA) subtypes (N1–N11) of influenza A viruses currently exist, which together define the influenza A virus subtype [12,13,14].
The process of influenza virus infection and proliferation mainly includes the following steps. (1) Adsorption: In the process of influenza virus infection, the virus first recognizes and adsorbs to the sialic acid (SA) receptor on the surface of the host cell glycoprotein through its surface HA and then enters the host cell to form intracellular bodies (endosomes) through receptor-mediated endocytosis. (2) Integration: The conformational change of the endosome mediates the fusion of the viral envelope with the endosome membrane and releases the ribonucleoprotein (VRNP) into the cytoplasm of the host cell. (3) Nuclear import and replication: The endosome activates the ion channel M2, which generates an inward proton flow that directs the transport of VRNP into the nucleus for viral RNA replication. Polymerase proteins (PB2, PB1, and PA) and NP play important roles in the transcription and replication of influenza viruses. (4) Assembly and release: The newly generated progeny viral RNA is transported out of the nucleus into the cytoplasm and assembled into the mature virus. The new viral particles are still attached to the outer membrane of the host cell through HA (Figure 2). Thus, blocking any of these four steps could prevent influenza virus infection, and key parts or key proteins of these steps would be possible targets for anti-influenza virus drugs. Antiviral small-molecule compounds currently on the market or in clinical use mainly include specific influenza virus inhibitors (M2 ion channel blockers, NA inhibitors, PA inhibitors, and PB2 inhibition agents) and some broad-spectrum antiviral drugs [15,16,17].
Influenza virus hemagglutinin (HA) binds to sialic acid-presenting receptors on the surface of host cell membranes. Virus particles enter the host cells to form endosomes through receptor-mediated cellular endocytosis. (2) Endosome acidification promotes conformational changes of HA, resulting in the uncoating of the virus and release of the vRNP into the cytosol of the host cell with further transportation to the nucleus. (3) vRNPs enter the nucleus to initiate the viral mRNA. HA, NA, and M2 are processed at the ER apparatus and Golgi before transport to the cell surface. Influenza virus polymerase can synthesize viral mRNA and vRNA. The vRNA is first converted into plus-strand cRNA; then the new vRNA is synthesized using cRNA as template. (4) Viral proteins and vRNA are transported to the cell surface to assemble progeny viruses and initiate the virus budding process. The progeny virus is then released from the surface of the infected cells and seeks new host cells to infect.

2.1. M2 Inhibitors

The main function of the M2 ion channel is to transport protons and induce fusion of the viral membrane with the endosomal membrane. M2 ion channel blockers act by binding to the interior of the ion channel, blocking the influx of protons and thereby preventing hemagglutinin-mediated membrane fusion [18]. The first antiviral drugs approved by the FDA to treat influenza A virus infection were adamantanes, including amantadine and rimantadine. They can block the M2 membrane protein ion channel, thereby preventing the dissociation of M1 protein and ribonucleoside protein and further blocking the initiation of the replication process of influenza A virus RNA to achieve therapeutic effect [19,20]. The M2 protein is only present in the membrane protein of influenza A virus. Therefore, these drugs belong to specific inhibitors of influenza A viruses and are ineffective against influenza B viruses. M2 ion channel blockers have been implicated in the virulence occurring in the digestive and autonomic nervous systems [21]. Forty years of long-term, widespread, and extensive use of these drugs has caused most influenza A viruses to be severely resistant to them [22,23].

2.2. NA Inhibitors

Neuraminidase (NA), also known as sialidase, plays a key role in the life cycle of the influenza virus. Currently, four drugs have been approved for clinical treatment; they include zanamivir (zanamivir, ZAN; Relenza®), oseltamivir phosphate (oseltamivir phosphate; Tamiflu®), peramivir (peramivir1; Rapivab®), and ranamivir (laninamiviroctanoata; Inavir®). In 1974, Palese et al. [24] discovered Neu5Ac2en, a sialic acid-analogue NA inhibitor. Zanamivir was successfully developed [25] in 1991 and approved by the FDA in July 1999. In 1996, oseltamivir was successfully developed and approved by the FDA in October 1999 [26]. Peramivir was approved for marketing in 2009 [27]. Ranimivir octanoate, a prodrug of ranimivir, was marketed in Japan in 2010 [26,27]. Currently, NA inhibitors are the research hotspot for anti-influenza virus drugs. However, existing NA inhibitors are not perfect drugs against influenza A viruses. For example, zanamivir has a high antiviral activity but low bioavailability and is rapidly excreted in the kidneys; oseltamivir often has side effects such as nausea and vomiting in adults [7]. Moreover, due to the widespread clinical application of existing NA inhibitors, influenza viruses mutate and develop different degrees of resistance to these drugs through the change in shape of the NA catalytic position to reduce the sensitivity of the virus to the NA inhibitor [7,28]. The catalytic site of NA consists of eight functional residues, and the surrounding eleven framework residues ensure the stability of the active site structure. Due to the hydrophobic bond of oseltamivir, NA must undergo rearrangement to adapt to drug binding. Any mutation that affects the rearrangement may reduce the binding affinity of oseltamivir [29]. In addition to H275Y mutation, which can enhance the drug resistance of oseltamivir, D199G, S247N, and I223M mutations can reduce the sensitivity of oseltamivir; I223R mutation can reduce the sensitivity of oseltamivir and zanamivir; and Q136K mutation can reduce the sensitivity of zanamivir [30].

2.3. Viral Polymerase Complex Inhibitor

Influenza virus polymerase is composed of alkaline PB1, PB2, and polymerase acidic (PA) [31]. The PB2 subunit binds to the cap of the host cell pre-messenger RNA and is subsequently cleaved by cap-dependent endonucleases in the PA subunit. This “cap-capturing” process provides RNA primers for the transcription of viral MRNAs through the RNA-dependent RNA polymerase function of PB1 [32]. These include the PB1 inhibitor, favipiravir (Avigan), which was approved for influenza treatment in Japan in 2014 [33]. Pimodivir, also known as JNJ-63623872 and VX-787, is a PB2 inhibitor with selective activity against influenza A viruses and is suitable for oral administration [34]. VX-787 has performed well in clinical studies. Baloxavirmarboxil (44, S-033188), a PA inhibitor, was marketed in Japan in February 2018 and in the United States in October 2018 as Xofluza for the treatment of influenza A and B, respectively [35].

2.4. NP Inhibitors

NP is one of the most abundant viral proteins produced during viral replication. During the viral life cycle, NP participates in the formation of viral ribonucleoprotein complexes (vRNPs) by binding to influenza virus RNA and polymerase subunits (PB 1, PB 2, and PA) and also participates in the nuclear import, replication, and export of vRNPs [36,37]. In recent years, NP has become a hotspot in antiviral drug development because of its important role in virus replication. Naproxen was identified by computer screening as a dual inhibitor of NP and COX2, a type of cyclooxygenase [38]. Naproxen can inhibit viral replication by targeting the RNA groove of NP and preventing NP from interacting with viral RNA. Naproxen showed good antiviral effect against H1N1 and H3N2 viruses both in vivo and in vitro [39]. Due to the strong serum variability of NP antigen, many researchers once thought that NP was not suitable as a drug target against influenza viruses. However, some NP inhibitors that are effective against both influenza A and B viruses have recently emerged. Kao et al. [40] confirmed that NP can serve as a target for influenza treatment drugs by forward chemical genetic technology and found a small-molecule compound named nucleozin, which targets NP. Nucleozin can induce NP aggregation into large NP complexes, completely antagonizing their entry and accumulation in the nucleus and inhibiting virus replication. Animal research results indicate that nucleozin can effectively treat highly pathogenic H5N1 influenza virus-infected mice [40]. Liu et al. [41] found that the groove surface between the head and body domains of influenza A NP are covered by a large number of conserved residues, playing an important role in influenza virus RNA binding. To explore the mechanism of NP binding to RNA, they performed [41] a series of directed induced mutations in the RNA binding slots and characterized the interaction between RNA and NP by surface plasmon resonance (SPR). Liu et al. [41] identified an influenza virus NP inhibitor, (E, E) -1,7-bis (4-hydroxy-3-methoxyphenyl) -1,6-heptadiene-3,5-dione. This inhibitor reduces the RNA binding affinity of NP and hinders viral replication. NP is highly conserved among influenza A virus strains from different species, indicating that influenza viruses are less likely to develop resistance to NP inhibitors. The multiple drug binding sites and high sequence conservation of NP makes it a highly anticipated drug target [42].

3. Natural Compounds That Exert Anti-Influenza A Effects

3.1. Marine Natural Products

The ocean is the largest treasure house of resources on the earth, with a huge diversity of species. Marine natural products have unique chemical structure and function, high biological activity, and research and development potential [43]. As a benefit from the unique aquatic environment and rich species in the sea, the active metabolites produced by marine microorganism are more novel in structure and more unique in function than terrestrial microorganisms [44]. Azaphilones are reported to be a class of fungal metabolites with antibacterial, antiviral, anti-inflammatory, antioxidant, nematocidal, and cytotoxic biological activities [45,46]. Currently, more than 430 kinds of azaphilones have been extracted from marine and terrestrial fungi. Among them, sclerotiorin E, (+) sclerotiorin, TL-1-monoactate, ochrephilone, 8-acetyldechloroisochromophilone III, scleratioramine, and isochromophilone IX (17), isolated form Penicillium sclerotiorum [47,48] (Figure 3), showed a comparable protective effect of canine kidney cells from H1N1 infection [49,50] (seeing Table 1). Data has shown that the dimer azaphilones had a stronger antiviral activity than the monomer, and this antiviral activity has no relationship with the substitution of chlorine atoms on C-5 or the oxygen atom connected to C-1 [51]. He et al. isolated a new antiviral cyclic tetrapeptide containing a rare 3-OH-N-CH3-Phe residue named asperterrestide A (8) from the fermentation broth of the marine-derived fungus Aspergillus terreus SCSGAF0162 [52] (Figure 3). CPE was performed for the influenza virus strains A/WSN/33 (H1N1) and strain A/Hong Kong/8/68(H3N2) to determine the inhibition of viral replication in cellular MDCK [52]. A cytotoxicity assay showed that asperterrestide A exhibited cytotoxicity against human cancer cell lines U937 and MOLT4, with IC50 values of 6.4 μM and 6.2 μM, respectively. Asperterrestide A showed inhibitory effects on A/WSN/33 (H1N1) and A/Hong Kong/8/68 (H3N2), with half-maximal inhibitory concentration (IC50) values of 15 and 8 μM, respectively (Table 1) [52]. Two new rubrolides were isolated from the fermentation broth of the marine-derived fungus Aspergillus terreus OUCMDZ-1925 [52]. Rubrolides S (9) (Figure 3) showed comparable or superior anti-influenza A (H1N1) virus activity to that of ribavirin, with an IC50 value of 87 µM (Table 1) in CPE inhibition assay [53]. A cytotoxicity assay showed that rubrolides S (9) was non-cytotoxic to A549, HL-60, HeLa, and HCT-116 cell lines (IC50 = 430 mM). In a study by Tian et al. [54], a new asteltoxin from Aspergillus sp. SCSIO XWS02F40, asteltoxin E (10) (Figure 3), exhibited an inhibitory activity against H1N1, with an IC50 value of 4 µM in a CPE inhibition assay [54] (Table 1). Five new phenolic polyketides and two known compounds were isolated and identified by Zhu et al. [55] from the fermentation broth of Streptomyces sp. OUCMDZ-3434 associated with the marine green algae Enteromorpha prolifera. The antiviral activity of seven compounds against influenza A H1N1 virus was evaluated using the CPE inhibition assay [49,50]. Wailupemycin J (11), R-wailupemycin K (12), and 5-deoxyenterocin (13) (seeing Figure 3) showed anti-H1N1 virus activity, with 47.8%, 42.5%, and 60.6% inhibitory activities, respectively, at 50 mg/mL (ribavirin, 45.3% inhibition) [55] (Table 1). The cytotoxicity assay of these compounds on HeLa cells showed that none of them exhibited cytotoxicity (IC50 > 50 μM). Pyropheophoride a (PPa) (14) (Figure 3) is a porphyrin derivative isolated from the seafood Musculus senhousei (M. senhousei), which exhibits promising anti-influenza activity in vitro. PPa showed inhibitory effects on A/Puerto Rico/8/34(H1N1), with an IC50 value of 0.17 μg/mL (Table 1). Studies have shown that PPa may interact with viral envelope lipids. Thus, the fusion of virus and cell membrane was blocked, and anti-H1N1 virus activity was achieved [56]. Antiviral activity of the South China Sea soft coral Sarcophyton sp. and all isolated compounds against H1N1 virus by CPE, namely (24R)-methylcholest-7-en-3β,5α,6β-triol (15) (Figure 3) and (24S)-ergost-3β,5α,6β, 11α-tetraol (16) (Figure 3), showed relatively strong inhibitory effects on A/Puerto Rico/8/34(H1N1), with IC50 values of 19.6 and 36.7 μg/mL (Table 1) [57].

3.2. Flavonoids

Flavonoids are one of the largest classes of secondary metabolites in plants. They are bio-synthesized through the pathways of shikimic acid/phenylpropane and acetic acid/malonic acid. Flavonoids contain a common C6-C3-C6 skeleton, which has two aromatic rings that are connected by a three-carbon bridge, typically forming a benzene chromone arrangement. According to degree of unsaturation and substitution mode, flavonoids can be classified into sub-classes: flavones, flavonols, flavanone, flavan-3-ol, anthocyanins, dihydroflavonols, and isoflavones as well as the form of biogenetic intermediate chalcone. In recent studies, flavonoids have been widely recognized as NA inhibitors.
Cleistocalyx operculatus (Roxb.) Merr. and Perry (Myrtaceae), widely used in traditional medicine in China, Vietnam, and other tropical countries, has diverse biological activities [93]. It is utilized to treat various conditions, including fever and bacterial dysentery, and their main chemical components, flavonoids, exhibit antioxidant, anti-hyperglycemic, anti-influenza, and cholinesterase-inhibitory properties [93,94,95,96]. Therefore, to further investigate the inhibitory activity of their extracts against influenza virus NA, Ha et al. [58] repeatedly isolated nine compounds from rhizomes of Pentarhizidium orientale through a series of chromatographic procedures. By a CPE inhibition assay, 2′,4′dihydroxy-6′-methoxy-3′,5′-dimethylchalcone (17) and 3′,5′-dimethylether 3-O-β-D-galactopyranoside (18) (Figure 4) not only had inhibitory activity against the influenza virus H1N1 A/PR/8/34 NA but also showed inhibitory activity against wild-type novel swine influenza (WT) and oseltamivir-resistant (H274Y mutation) viral NA. The IC50 values were 8 and 9 μM, respectively (seeing Table 1). The inhibition patterns of 2′,4′dihydroxy-6′-methoxy-3′,5′-dimethylchalcone (17) and 3′,5′-dimethylether 3-O-β-D-galactopyranoside (18) (see Figure 4) were further investigated using the double-reciprocal Lineweaver–Burk plot, which showed that the two compounds exhibit a noncompetitive inhibition pattern against influenza NA [58].
In one study, C-methylated flavonoids were reported to exhibit anti-influenza virus (H1N1)-inhibitory effects [97]. Pentarhizidium orientale (Hook.) Hayata (Onocleaceae) (syn. Matteuccia orientalis) is a perennial pteridophyte that is distributed mainly in East Asia and the temperate regions of the northern hemisphere [59,97]. As part of the present investigation on the bioactive compounds from a Korean medicinal plant [98], the phytochemicals in an 80% aqueous MeOH extract of the P. orientale rhizome were purified [59]. The NA-inhibitory activity of several isolated compounds from P. orientale against the H1N1 influenza virus was tested using the CPE inhibition assay [59]. The results indicated that demethoxymatteucinol (19), matteucinol (20), matteucin (21), methoxymatteucin (22), and 3′-hydroxy-5′-methoxy-6,8dimethylhuazhongilexone (23) (see Figure 4) exhibited NA-inhibitory activities, with IC50 values ranging from 24 to 30 μM (Table 1). These compounds were compared with oseltamivir [59]. Demethoxymatteucinol (19) exhibited cytotoxicity against MDCK cells (CC50, 77.6 μM), while other compounds did not show significant cytotoxicity.
Bee pollen is a combination of plant pollen and bee secretions and nectar. Bee pollen contains bioactive compounds including proteins, amino acids, lipids, carbohydrates, minerals, vitamins, and polyphenols [60]. kaempferol-3-sophoroside (24), kaempferol-3-neohesperidoside (25), kaempferol-3-sambubioside (26), kaempferol-3-glucoside (27), quercetin-3-sophoroside (28), luteolin (29) (Figure 4), and chelianthifoline (30) (Figure 5) A-inhibitory activities of seven compounds were evaluated against the recombinant influenza viral subtypes, namely H1N1, H3N2, and H5N1, with zanamivir as a positive control. All compounds inhibited NAs in a dose-dependent manner. The IC50 values of seven compounds on the NAs from influenza H1N1 ranged from 11 to 101 µM (see Table 1). The modes of inhibition of these compounds were further investigated using Dixon plots, which showed that six compounds other than quercetin-3-sophoroside (28) exhibited non-competitive inhibitory activities. The authors further analyzed the structure–activity relationship and concluded that the bulky sugar moiety in compounds 2428 causes a decrease in activity. The cytotoxicity assessment revealed that kaempferol-3-sophoroside (24), kaempferol-3-neohesperidoside (25), kaempferol-3-sambubioside (26), kaempferol-3-glucoside (27), quercetin-3-sophoroside (28), luteolin (29), and chelianthifoline (30) exhibited no toxicity to MDCK cells at a concentration of 100 µM, with a cell viability of 100% [60].
Rhodiola rosea L. belongs to the plant family Crassulaceae, which is widely distributed in the world. It is a traditional Chinese medicine with various biological activities, such as anti-diabetes, anti-cancer, anti-inflammatory, anti-aging, and anti-depression [99,100,101]. Five flavonols, i.e., kaempferol (32), herbacetin (34), rhodiolinin (38), rhodionin (39), and rhodiosin (40) (Figure 4), were isolated from Rhodiola rosea by Lee et al. [61] and compared with the commercially available flavonoids apigenin (31), luteolin (29), quercetin (33), gossypetin (35), cosmosiin (36), astragalin (37), linocinamarin (41), rutin (42), and nicotiflorin (43) (Figure 4) to facilitate analysis of their structure–activity relationship. Fourteen compounds showed HIN1-inhibitory activities, with IC50 values ranging from 2 to 57 µM (Table 1). Exploring the structure–activity relationship between kaempferol (32) and four hydroxyls, the authors further investigated the optimal position and number of hydroxyl groups on the flavonoid skeleton and showed that quercetin (33) activity with more hydroxyl substitutions was comparable to herbacetin activity with a 7,8-dihydroxyl group. Luteolin (29) with 3′ and 4′ dihydroxy groups has higher activity than kaempferol (32) with 3,4′-dihydroxy groups, whereas apigenin (31) with one less hydroxyl group is relatively less active. Gossypetin (35) showed 10- or 4-fold increased potency against both Clostridium perfringens and rvH1N1 NAs but showed similar activity against rvH1N1 NA to quercetin (33). Contrary to the conclusion of Du et al. [102], Lee et al. suggested [61] that the hydroxyl groups on the A or B rings are necessary for the inhibitory effect on rvH1N1 NAs. Additionally, the authors compared the activities of glycoside flavonoids and glycoside flavonoids. Astragalin (37) and nicotiflorin (43) obtained from replacing the hydroxyl group of kaempferol (32) showed reduced anti-fctivity. Therefore, Lee et al. [61] concluded that glycosides can cause reduced anti-influenza activity (The binding site with the target enzyme activity may be interrupted by a bulky sugar moiety).
Elderberry (Sambucus nigra L.) is a European native plant commonly known as black elderberry, European elderberry, European elderberry, and European black elderberry. It is rich in phenolic compounds, including phenolic acids, flavonoids, catechins, and proanthocyanidins [103]. In folk medicine, its flowers and berries are used to treat fever, cough, nasal congestion, and influenza in addition to being widely used as an anti-inflammatory, analgesic, and diuretic agent [104,105]. Zakay-Rones et al. [106] showed that elderberry extract exhibited anti-influenza activity in human clinical trials [106]. The IC50 of the extract for H1N1 was 252 µg/mL, while 100% inhibition of H1N1 infection was achieved at 1000 µg/mL. Roschek et al. [62] confirmed by direct binding assays that the flavonoids in elderberry extracts can bind to H1N1 influenza virus and block the ability of the virus to infect host cells upon binding. To verify the anti-influenza mode of action and in vitro anti-influenza activity of the extract, the authors synthesized 5,7,3′,4′-tetra-O-methylquercetin (44) and dihydromyricetin (45) (Figure 4). Usually, the in vitro anti-H1N1 mechanisms of polyphenols prevent intranuclear acidification [107], inhibit membrane fusion [108], inhibit the release of offspring virions [109], inhibit NA activity [109,110], and inhibit intercell replication [111]. Roschek et al. further verified that the synthetic compounds and elderberry extracts use viral envelope binding to inhibit H1N1 infection and may involve the HA domain in host cell binding and recognition. 5,7,3′,4′-tetra-O-methylquercetin (44) gave an IC50 of 0.1 µg/mL (0.4 µM) for H1N1 infection inhibition, whereas dihydromyricetin (45) achieved an IC50 of 3 µg/mL (9 µM) (Table 1) [62].
Honeysuckle (HS; Lonicera japonica) has been used as both medicine and food. In recent years, research has explored its various pharmacological effects, including anti-inflammatory, antibacterial, antioxidant, etc. It has significant applications and research implications in health care and disease treatment [112]. Honeysuckle has been used to treat the flu for thousands of years [113]. To clarify the main antiviral components of honeysuckle and the underlying mechanisms of its action, Li et al. [64] assessed the inhibitory activity of the total extract, organic acids extract, flavonoids extract, and acid–flavonoid mixture of honeysuckle against IAVs (H1N1 and H3N2) in vitro and in vivo. The total extract induced the lowest cytotoxicity, with a CC50 value of 350 µg/mL in MDCK cells. The acid–flavonoid mixture showed the most effective antiviral activity against H1N1, with an EC50 value of 4 µg/mL [64]. Unlike other studies, Li et al. [64] demonstrated the anti-influenza efficacy in vivo. Experiments have shown that oral administration of organic acid extracts can effectively reduce the mortality of mice infected with H1N1 virus. Oral administration of the acid extract at a dosage of 600 mg/kg/d significantly alleviated influenza virus-induced acute lung injury, improved the lung parameters, and improved the survival rate of the mice by 30%. To explore the anti-influenza mechanism of the four honeysuckle extracts, the author found through the time of drug addition (TOA) experiment that the honeysuckle extract mainly inhibited the replication and release of influenza virus into the host cells rather than viral adsorption or penetration. The influenza virus NA plays an important role by hydrolyzing sialic acid residues in progeny viruses and promoting the release of progeny virus particles. Therefore, using MUNANA, a special fluorescent substrate, they detected that four honeysuckle extracts significantly inhibited the NA activity of various influenza viruses in a dose-dependent manner. The authors found that the extract had broad-spectrum activity not only against H1N1, H3N2, H5N1, and H7N9 viral NA but also on oseltamivir-resistant mutant strains.
Salvia plebeia R. Br., a globally distributed edible plant, is commonly utilized in countries like India, China, Japan, and Korea as a folk remedy for various ailments, including the common cold, flu, cough, hepatitis, and hemorrhoids. [114,115,116]. To identify new antiviral lead compounds, the phytochemistry of conifer was studied. The methanolic extracts from the above-ground fraction of S. plebeia were extracted with CHCl3, EtOAc, and n-BuOH. The EtOAc and CHCl3 fractions were analyzed by continuous chromatography, which led to the isolation of 14 compounds. Chemical studies of the methanolic extract of S. plebeia isolated two new benzoylated monoterpene glycosides, polyglycosides A and B, and twelve known compounds, four flavonoids, two sesquiterpenes, four phenolics, one steroid, and one triterpene. Hispidulin (46), nepetin (47) (Figure 4), rosmarinic acid methyl ester (48), and luteolin (29) (Figure 4) exhibited moderate NA-inhibitory activity, with an IC50 ranging from 11 to 20 µM (seeing Table 1), compared with the inhibitory effects of the positive control (oseltamivir) (IC50 = 0.1 µM). Additionally, Bang et al. [63] confirmed their inhibitory effect on the virus through CPE experiments. Cell survival increased significantly at hispidulin (46), nepetin (47), and rosmarinic acid methyl ester (48) concentrations of 40 µM. Moreover, hispidulin (46) restored the chromosome condensation caused by H1N1 virus infection of MDCK cells. Among them, hispidulin (46), nepetin (47), and luteolin (29) had a flavonoid skeleton, hydroxyl groups on C-7 and C-4′, and α- and β-unsaturated groups on C-2, C-3, and C-4, which satisfy the structure–activity relationship of flavonoids on NA inhibition [58]. Rosmarinic acid methyl ester (48) is a caffeic acid ester of salvianic acid A (3,4-dihydroxyphenyllactic acid). Caffeic acid derivatives have been reported as novel influenza NA inhibitors [117].
In conclusion, the above flavonoids extracted from natural products have some in vitro activities against influenza virus, and they all have anti-influenza virus activity by inhibiting NA activity (Figure 2). Therefore, the analysis of the above studies can summarize the structure–activity relationship of flavonoids on NA inhibition to elucidate the structural characteristics of flavonoids on NA inhibition, which was summarized in Figure 6. According to Liu et al. [102], in the structure–activity relationship study of flavonoid anti-influenza virus NA inhibitors, for flavonoids to have better anti-influenza virus activity, OH groups on C-7 and C-4′, a double bond between C-2 and C-3, and a carbanyl group at the C-4 position must be present. The structure of the 3′,5′-dimethylether 3-O-β-D-galactopyranoside (18) (Figure 4) exhibiting significant NA-inhibitory activity fully meets the above requirement. However, the IC50 values of demethoxymatteucinol (19), matteucinol (20), matteucin (21) (Figure 4), and other flavonoids with double bonds between C-2 and C-3 did not show significant changes in activity.
In exploring a new compound from the leaves of Cleistocalyx operculatus and its inhibitory activity against influenza A neuraminidase, Ha et al. [58] proposed that OH at the C-3′ site was replaced by methoxy, which significantly improved the inhibitory activity of flavonoids against NA. In investigating favone with six different modification groups and comparing their anti-influenza activities, Morimoto et al. [118] concluded that the modification of the B ring from the 3′ to 5′ position is important, and the hydroxyl group should preferably be at 3′ and 4′ positions rather than the 5′ position. Liu et al. [102] suggested that the increase in the number of OH groups on the B ring reduced its inhibitory effect. In 5,7,3′,4′-tetra-O-methylquercetin (44) and dihydromyricetin (45) (Figure 4), dihydromyricetin (45) with three hydroxyl groups on the B ring showed reduced activity against influenza virus compared with 5,7,3′,4′-tetra-O-methylquercetin (44). Moreover, Morimoto and Liu et al. [118,119] believed that the inhibitory effect of glycosylated flavonoid compounds would be weakened. This finding may be due to the large volume of glycosylation leading to steric hindrance or the interruption of binding sites with target enzyme activity. Therefore, we concluded that the C-7, OH, C-4′ positions, and C-4 carbon groups of flavonoid compounds are necessary to obtain good NA inhibition.
Excessive OH modification from the 3′ to 5′ sites on the B ring reduces activity. Moreover, any glycosylation group at any location will greatly reduce its activity. However, the IC50 values of demethoxymatteucinol (19), matteucinol (20), matteucin (21) (Figure 4), and other flavonoids with double bonds between C-2 and C-3 did not show significant changes in activity. In exploring the new compound of atresia leaf and its inhibitory activity against influenza A neuraminidase, Ha et al. [97] proposed that OH at the C-3′ site be replaced by methoxy, which significantly improved the inhibitory activity of flavonoids against NA. Similarly, Morimoto et al. [118], in investigating flavone with six different modification groups and comparing their anti-influenza activity, concluded that the modification of the B ring from 3′ to 5′ position is important, and the hydroxyl group should preferably be at 3′ and 4′ position rather than the 5′ position.

3.3. Alkaloids

Alkaloids refer to a naturally occurring class of nitrogen-containing organic compounds, excluding low-molecular-weight amines, amino acids, peptides, and proteins. They often have complex nitrogen heterocyclic structures, most of which are alkaline and can combine with acids to form salts. Alkaloids are a class of natural compounds usually with strong biological activity. They have been extensively studied for their broad-spectrum antiviral activities against different DNA and RNA viruses [119].
Commelina communis L. (also known as dayflower), a globally distributed herb, has been used in traditional Chinese medicine for the treatment of non-infectious fever, edema, cutin, diabetes mellitus, and other ailments [120]. Chemical components such as flavonoids, alkaloids, polysaccharides, terpenoids, and sterols have been isolated from this plant [121]. Through spectroscopic analysis, the chemical structures of the three compounds were determined as harman, homonojirimycin (HNJ) (49) (Figure 5), and 2,5-dihydroxymethyl-3,4dihydroxypyrrolidine. The results show that HNJ has strong antiviral activity against the influenza virus A/PR/8/34 (H1N1), with an EC50 value of 10 μg/mL (Table 1) and an SI value of 18. These features are comparable to those of the approved antiviral drug, ribavirin [65].
Mangroves are a unique forest ecosystem, mainly distributed in tropical and subtropical intertidal zones, with rich biodiversity and abundant actinomycetes [70,122]. To search for bioactive products from mangrove actinomycetes, a new endophytic actinomycete was isolated from the roots of Xylocarpus granatum (Meliaceae) and identified as Jishengella endophytica 161,111 [123]. Wang et al. [66] chemically studied the EtOAc extract of strain 161,111 and identified 13 compounds. These 13 compounds were tested for their antivirus effects on H1N1 using the CPE inhibition assay, and ribavirin was used as the positive control, with an IC50 value of 23 µg/mL. The results showed that perlolyrine (50), 1-hydroxy-β-carboline (51), lumichrome (52), and 1H-indole-3-carboxaldehyde (53) (Figure 5) have moderate anti-H1N1 activity with semi-inhibitory concentrations of 38, 25, 40, and 46 μg/mL, respectively (seeing Table 1). The cytotoxicity assay revealed that perlolyrine (50), 1-hydroxy-β-carboline (51), and 1H-indole-3-carboxaldehyde (53) exhibited mild toxicity towards MDCK normal cells, with CC50 values of 116.3 ± 12.1, 403.2 ± 31.4, and 522.5 ± 24.5 μg/mL, respectively.
Since the 1980s, studies have been conducted on marine sponges belonging to the class Calcarea, producing large amounts of bioactive alkaloids containing imidazole heterocyclary substitutions [67]. This alkaloid has been reported to have cytotoxic [124,125,126], antimicrobial [127], and antifungal [128] properties. Additionally, they possess leukotriene B4 receptor [129] and epidermal growth factor receptor [130] antagonist activities. According to a chemical study by Gong et al. [67] on sponge-heterologous organisms collected from the South China Sea, a new imidazole alkaloid named naamidine J—along with four known ones: naamidine H, pyronaamidine, leucettaamine B, and leucettamine C (54)—was identified (Figure 5) [131,132]. Of these five alkaloids, leucettamine C (54) exhibited weak anti-H1N1 activity with an inhibition ratio of 33% (Table 1). The cytotoxicity assay revealed that leucettamine C (54) did not exhibit cytotoxicity on human leukemia (K562), acute myeloid leukemia (HL-60), cervical cancer (HeLa), and lung adenocarcinoma (A549) cells.
Peganum harmala L. (family Zygophyllaceae) is a perennial, hairless plant widely grown in China, the Middle East, India, and South America [133,134,135]. Moradi et al. [68] found that the crude extract inhibited the replication of type A PR8 virus in this cell line, with an IC50 value of about 10 (95% CI: 7–11) µg/mL (Table 1). The total alkaloids of this extract had antiviral activity, with an IC50 value of about 6 (95% CI: 4–9) µg/mL (Table 1), which was better than that of the crude extract. To explore the mode of action of the extract against influenza virus replication, Moradi et al. [68] studied the mode of action of viricidal activity against influenza virus replication using tests, including a coagulation inhibition test, time of addition test, RNA replication, Western blot analysis, and RNA polymerase blocking assay. The results suggested that harmala seed extracts may reduce NP levels and viral polymerase activity, thereby affecting the RNP complex activity and subsequently inhibiting viral RNA transcription and replication.
Papaverine (55) (Figure 5) is a non-narcotic opiate alkaloid. Medicinal papaverine is used as a smooth muscle relaxant for the treatment of vasospasm and erectile dysfunction, and its mechanism of action is the inhibition of phosphodiesterase 10A [136,137,138]. Papaverine has antiviral activity, with an IC50 value of about 17 µM (Table 1). Papaverine exhibited a dose-dependent inhibition of influenza virus strains used in this study (A/WSN/33, A/Udorn/72, A/Eq/2/Miami/1/63, B/Lee/40, and B/MD/59) [69]. Aggarwal et al. [69] identified the inhibition step in the viral life cycle through the time of addition (TOA) and time of elimination (TOE) experiments. The results collectively suggest that the inhibitory effect of papaverine occurs late in the influenza virus infection cycle.
Aggarwal et al. [69] evaluated the effects of papaverine on the activities of HA and NA proteins and found that papaverine had no effect on the activities of HA or NA surface glycoproteins, proving that papaverine did not interfere with the entry or release steps of the virus life cycle. Viral RNA synthesis was analyzed by semiquantitative reverse transcriptase PCR (RT-PCR). The viral RNA grew in the presence of a specified concentration of papaverine, and the results showed that papaverine did not affect the synthesis of influenza viral RNA. HEK293T cells were infected with A/WSN/33 virus and treated with papaverine or DMSO. Western blotting analysis showed that the presence of papaverine reduced the phosphorylation of MEK and ERK (Figure 2). However, the amount of MEK and ERK remained the same. Therefore, the results suggest that papaverine alters the activation of MEK/ERK pathway in 293T cells. Papaverine was found to affect the morphology of influenza virus by inhibiting nuclear export of the viral genome (Figure 2).
Isatis indigotica Fort. (Cruciferae) is a widely grown medicinal plant. The dried leaves and roots of the plant, known in China as “da qin ye” and “ban lan gen”, are used in traditional Chinese medicine to treat influenza and other infections [139]. Water extracts from I. indigotica leaves and roots were studied. Fifty-seven new alkaloids were isolated from Radix isatidis, of which 22 alkaloids were indoles and bisindole alkalosides. Seven indole alkaloid glycosides containing the 1′-(4′-hydroxy-3′,5-dimethoxyphenyl) ethyl unit were isolated from the I. indigotica leaf (da qing ye) decoction by Guo et al. [140]. Isatidifoliumosides (56)/epiisatidifoliumosides C (57) (Figure 5) in the 3:2 ratio exhibited antiviral activity against influenza virus H1N1 PR8, with an IC50 of 65 µmol/L (seeing Table 1) (the positive control ribavirin, IC50: 54 µmol/L).

3.4. Terpenoid Derivatives

Terpenoids are compounds derived from meglutaric acid and their derivatives whose molecular framework is based on isoprene unit (C5 unit). Terpenoids have anti-tumor, anti-inflammatory, antibacterial, and antiviral properties. They prevent cardiovascular and cerebrovascular diseases and are one of the most abundant compounds in natural products [141].
The genus Sonneratia, consisting of nine mangrove plant species in the family Sonneratiaceae, is widely distributed in tropical and subtropical regions. These plants exhibit diverse biological activities and have been traditionally employed for treating ailments including asthma, ulcers, hepatitis, hemorrhoids, sprains, and bleeding [142]. Chemical studies of the aerial parts of the mangrove plant, Sonneratia paracaseolaris, yielded five new triterpene paraproteins: proteins A–E and 12 known analogues. Additionally, the CPE test was used to evaluate the antiviral activity of all isolates against the influenza A H1N1 virus (IAV) [142]. Only paracaseolins A (58) (Figure 7) showed significant anti-H1N1 viral activity, with an IC50 value of 28 µg/mL, which was close to that of the ribavirin positive control, with an IC50 value of 25 µg/mL (Table 1). The other compounds showed no activity with inhibition rates of <50% at 50 µg/mL [71].
Ganoderma lingzhi (formerly called Ganoderma lucidum [143,144]), an oriental fungus, has been used in traditional Chinese medicine for thousands of years [145]. The lanostane-type triterpenoids are the main bioactive components of G. lingzhi, and they have been reported to have various physiological activities, including anticancer, immunomodulatory, antihypertensive, antiandrogenic, antidiabetic, and antiviral properties [146,147]. The direct effect of the hot water extract of G. lingzhi on NA was evaluated by Zhu et al. using an in vitro NA-inhibition assay with four different influenza A virus subtypes [72]. The extract strongly inhibited the activities of NAs derived from the influenza A virus subtype H1N1, with an IC50 value of 15 µg/mL. Moreover, Zhu et al. [72] studied the anti-influenza effects of G. lucidum hot water extract in two ways, i.e., intranasally and orally, in a convalescent mouse model. Their results showed that intranasal injection of G. lingzhi hot water extract had a direct inhibitory effect on influenza A virus and could effectively alleviate weight loss during influenza infection in mice. In further NA-inhibition experiments, the extracts showed strong inhibitory effects on NA activity of influenza A virus H1N1 subtype (IC50 = 15 µg/mL) and H5N1 subtype (IC50 = 2 µg/mL) (Table 1). However, no inhibitory effect was observed on NA activity of the other two influenza A viruses, H3N2 and H7N9 subtypes (IC50 >900 µg/mL).
In the process of screening for novel anti-influenza virus agents from microbial metabolites, two novel diterpene compounds, wickerols A (59) and B (60) (Figure 7), were isolated from the culture broth of a fungus, Trichoderma atroviride FKI-3849, by Yamamoto et al. [73]. Wickerol showed a strong antiviral activity against the A/H1N1 flu virus (A/PR/8/34 and A/WSN/33 strains), with an IC50 value of 0.1 mg/mL (Table 1), but it was not active against the A/H3N2 virus [73].
Located in the tropics and subtropics, the South China Sea has a vast sea area and rich marine biological resources. Since the 1980s, many compounds with novel structures, including alkaloids and terpenoids, have been isolated from various corals collected in the South China Sea, and the isolated chemical components have been studied [57,148,149,150,151,152]. Lemnalia sp. (No. xssc2011907) is a soft coral collected from the coast of the Xisha Islands in the South China Sea. A chemical study of acetone extract by Yan et al. [74] revealed thirteen structurally different terpenoids, six new diterpenoids, and three known related compounds. Four sesquiterpenes showed anti-H1N1 viral activity at 30 µM with 77.6–100% inhibition, whereas the novel sesquiterpene lineolemnenes F (61) (Figure 7) showed activity with an IC50 of 6 µM (Table 1).
Celastrus aculeatus Merr. is an evergreen vine widely distributed in southern China [151]. In folk culture, the roots and stems of this plant are often used in traditional Chinese medicine to treat diseases such as rheumatoid arthritis, gout, cholecystitis, nephritis, and hypertension, and they have many biological activities [152]. Previous studies have demonstrated the inhibitory effect of triterpenes and diterpenes (including phenolic and indexing diterpenoids) on H1N1 virus A [153]. The phytochemical exploration of C. aculeatus was performed to identify new antiviral lead compounds by Chen et al. [75]. Antiviral activities of the compounds were evaluated. The EtOAc-soluble fraction of the methanolic extract of the whole plants of C. aculeatus was subjected to silica gel chromatography and then purified by semipreparative HPLC to yield seven diterpenoids. Antiviral activities of the compounds were preliminary evaluated on the A/PR/8/34 (H1N1) strain using oseltamivir as a positive control. The antiviral activity assay showed that aculeatusane A (62) (Figure 7) is active against A/GZ/GIRD07/09 (H1N1), with an IC50 of 23 µM and SI of 7 µM (Table 1).

3.5. Phenol Derivatives

Phenolic compounds are compounds formed by the combination of one or more aromatic rings with one or more hydroxyl groups. Phenolic compounds are secondary metabolites of plants, which have been confirmed to have various significant biological activities [154], including anti-cancer, analgesic, anti-inflammatory, and antibacterial effects. Additionally, they have biological activities against diabetes mellitus, cardiovascular and cerebrovascular diseases, and influenza A virus.
Calotropis gigantea (Asclepiadaceae) is a shrub found in East and Southeast Asia, and its bark and leaves have been traditionally used in Chinese folk medicine [154]. A series of bioactive secondary metabolites, such as cardiosteroids, triterpene alcohols, alkaloids, and flavonoids, have been isolated from different parts and have shown pharmacological activities such as analgesic, sedative, anti-inflammatory, and antidiarrheal effects [155,156,157,158]. A new lignan glycoside, (+)-pinoresinol 4-O-[6″-O-vanilloyl]-β-D-glucopyranoside (63) (Figure 8), was isolated from the latex of Calotropis gigantea (Asclepiadaceae) by Parhira et al. [76]. Three isolates and one authentic compound were screened for A/PR/8/34 (H1N1)-inhibitory activity by CPE inhibition assay on MDCK cells. (+)-pinoresinol 4-O-[6″-O-vanilloyl]-β-D-glucopyranoside (63) showed an inhibitory activity against A/PR/8/34 (H1N1). An antiviral activity assay showed that (+)-pinoresinol 4-O-[6″-O-vanilloyl]-β-D-glucopyranoside is active against A/GZ/GIRD07/09 (H1N1), with an IC50 of 25 µM (Table 1). Parhira et al. [158] tested the antiviral mechanism of (+)-pinoresinol 4-O-[6″-O-vanilloyl]-β-D-glucopyranoside through TOA. The results showed that, unlike oseltamivir, it blocked the release of progeny virions and blocked the later stage of viral replication. The antiviral mechanism of (+)-pinoresinol 4-O-[6″-O-vanilloyl]-β-D-glucopyranoside is by inhibiting the early stages of influenza virus replication. Furthermore, it inhibited virus-induced NF-kB activation in a dose-dependent manner [159]. These results suggest that (+)-pinoresinol 4-O-[6″-O-vanilloyl]-β-D-glucopyranoside prevents influenza virus replication by inhibiting NF-kB activation, leading to impaired nuclear RNP output.
Tea polyphenols are a general group of natural polyhydroxyphenolic compounds extracted from tea, which have a variety of physiological functions such as antioxidant, anti-cancer, anti-radiation, anti-aging, cardiovascular disease prevention, and antiviral activity [160]. The in vitro antiviral activity of 13 tea polyphenols against influenza A and B viruses was comprehensively investigated. The results showed that five compounds, namely (−)-epigallocatechin (EGC) (64), (−)-epigallocatechingallate (EGCG) (65), procyanidin B-2 (66), procyanidin B-2 3,3′-di-O-gallate (67), and theaflavin (68) (Figure 8), had anti-influenza A virus activity [77]. Based on these results, Yang et al. [77] found that dimeric flavonoid 3-ols without gallic groups, such as theaflavin (68) and procyanidin B-2 (66), had a broader spectrum of anti-influenza virus activity (Table 1). The dimer molecule showed stronger activity against influenza virus than catechin monomer. Additionally, compared with methylated EGC, the EGC phenol hydroxyl group (64) on the B ring played an important role in anti-influenza A virus activity. The antiviral effects of tea catechin and its dimer on influenza A and B viruses may have the following two mechanisms: (1) blockage of the binding of virus to cell receptors in the early stage of viral infection and (2) inhibition of viral replication after the entry of virus [77]. Theaflavin (68) and procyanidin B-2 (66), two dimers of flavan-3-ols, both inhibit influenza A and B viruses, possibly due to the combination of the two mechanisms mentioned above.

3.6. Polysaccharides

Polysaccharides are carbohydrate substances with complex and large molecular structures, formed by condensation of several monosaccharide molecules and loss of water. Polysaccharides are natural macromolecular substances composed of ten or more monosaccharides bonded by glycosidic bonds, which are the material basis of active ingredients in traditional Chinese medicine. Natural polysaccharides are widely found in nature and have various pharmacological effects, such as anti-tumor and immune regulatory effects. The following discussion focuses on some polysaccharides isolated from natural products with inhibitory activity against influenza A viruses.
Naturally occurring sulfated exopolysaccharides have exhibited varying degrees of inhibitory activity, probably depending on their molecular weight and degree of sulfation [161,162,163]. p-KG03, is found in the marine dinoflagellate Gyrodinium impudium [164], and the results from virus-infected cells showed that influenza A viruses were sensitive to p-KG03, with EC50 values of 0.5 and 0.2 µg/mL against PR8 and TW, respectively (Table 1). Kim et al. [78] demonstrated that p-KG03 has anti-H1N1 influenza virus activity in MDCK cells, with an EC50 of 0.2–0.5 µg/mL (IS > 200). In the mechanism study, the inhibition of influenza virus replication was greatest within 0–6 h after infection, indicating that the compound mainly targets the adsorption and internalization steps in the early stages of the viral replication cycle (Figure 2). The binding experiments and fluorescence microscopy showed that the antiviral activity of p-KG03 was directly related to the interaction between p-KG03 and viral particles.
Jiao et al. [79] extracted polysaccharides from two types of brown algae (A. nodosum and F. Vesiculosus) using a sequential extraction method. Methanol was used for the initial extraction to remove non-polysaccharide components soluble in methanol. Subsequently, cold water, hot water, and alkaline solution were used for separate extractions. The antiviral activity of polysaccharides isolated from various Atlantic seaweed species against influenza was then evaluated. The results demonstrated that all extract of polysaccharides from A. nodosum and F. vesiculosus exhibited significant inhibitory effects on the A/PR/8/34 (H1N1) influenza virus and displayed high or medium antiviral activities against the influenza A/PR/8/34 virus (Table 1). In another study, Yu et al. [80] extracted polysaccharides in three conditions, namely 85% ethanol at 80℃, 20 volumes of distilled water at 80℃, and 10 volumes of 4% NaOH at 60℃, from the red algae Eucheuma denticulatum and evaluated the anti-influenza A (H1N1) viral activity of these three extracts using the Madin–Darby canine kidney cell model. Among them, ethanol extract had a good activity against H1N1 virus, with a semi-inhibitory concentration of 277 μg/mL and 52% inhibition against H1N1 virus at 250 μg/mL (Table 1). The semi-inhibitory concentration of ι -carrageenan water extract was 366 μg/mL, while NaOH extract had less anti-H1N1 virus activity (IC50 > 430 μg/mL) (Table 1). The available data suggest that hybrid carrageenan ethanol extract obtained in the future could be used as a potential anti-H1N1 viral inhibitor. The difference in the effectiveness of ethanol extract and water extract against the H1N1 virus may be due to the different levels of acidification and viscosity.
The antiviral activity of polysaccharides has been reported to be related to monosaccharide composition, molecular weight, and sulfation level [165,166,167,168]. Ray et al. [165] found that the degree of sulfation (DS) of the polymer was an important parameter for the antiviral activity of polysaccharides, and the higher the DS, the better the antiviral effect [169]. From a structural point of view, highly charged molecules effectively interfere with electrostatic interactions between the positively charged regions of viral glycoproteins and the negatively charged heparan sulfate (HS) chains of glycoprotein receptors on the cell surface. Sulfated polysaccharides with sulfate content higher than 20 (mol%) showed obvious antiviral activity. A higher average molecular weight (MW) indicated a higher antiviral activity. High-molecular-weight polysaccharides can significantly inhibit the binding and entry of viral receptors, but low-molecular-weight polysaccharides can better reduce the intercellular transmission of viruses [165].

3.7. Miscellaneous Compounds

In addition to the natural compounds mentioned above, a number of coumarins, steroidal saponins, and other compounds have been found to have anti-influenza activities. The roots of Ilex asprella (Hook. et Arn.) Champ. ex Benth. (Aquifoliaceae) are widely used in Chinese medicine to treat diseases such as influenza, amygdalitis, and pertussis [81]. A novel saponin containing sulphonic groups, namely asprellcoside A (69) and 3,4,5-trimethoxyphe β-D-5-O-caffeoyl-apiofuranosyl-(16) -β-D-Glucopyranoside (70) (Figure 9), was isolated from the roots of Ilex asprella and inhibited influenza virus strain A/PuertoRico/8/1934 (H1N1) strongly, with EC50 values of 4.1 and 1.7 mM (Table 1), respectively. Moreover, both compounds inhibited the secretion of IP-10, with EC50 values of 7 and 3 mM, respectively. In ref. [81], asprellcoside A (69) and 3,4,5-trimethoxyphe β-D-5-O-caffeoyl-apiofuranosyl-(16) -β-D-Glucopyranoside (70) showed strong anti-inflammatory effects on influenza-related tests and reduced levels of the chemokine IP-10 in H1N1-infected cells (Figure 9). However, further studies are needed on their antiviral mechanisms, effects on the virus life cycle, and targets. Toxicity assays showed that asprellcoside A (69) and 3,4,5-trimethoxyphe β-D-5-O-caffeoyl-apiofuranosyl-(16) -β-D-Glucopyranoside (70) showed no cytotoxicity (CC50 > 100 μM).
Lee et al. [82] identified four active furanocoumarins from 70% ethanol extract of Angelica dahurica (A. dahurica) root through the bioactivity-guided isolation: isoimperatorin (71), oxypeucedanin (72), oxypeucedanin hydrate (73), and imperatorin (74) (Figure 9). Among them, oxypeucedanin showed a significant CPE-inhibitory effect, which was stronger than that of the positive control, ribavirin, against H1N1, with an EC50 of 6 μM (Table 1). Oxypeucedanin inhibited the synthesis of NA and NP in a dose-dependent manner. Lee et al. [82] further studied the mechanism of oxypeucedanin and found that it interfered with the synthesis of NP and NA at the early stage of viral replication; it exerted anti-influenza activity but did not affect the virus’ entry into host cells, emergence, and release. The molecular docking analysis predicted the role of oxypeucedanin in polymerase acidic protein, which inhibits viral mRNA transcription and thus viral protein synthesis. Additionally, oxypeucedanin reduced H1N1-induced apoptosis by inhibiting the Bax/casepase-3 pathway (Figure 2).

3.8. Derivatives of Natural Products

Natural products have long been an important source of many marketed drugs, with studies showing that 10% of drugs on the market are unmodified natural products, and 29% are derivatives (hemicompounds) [170]. The natural product gentiopicroside (GPS) is an annular ether terpenoid, which is widely found in gentian (Gentiana manshurica Kitag.), which is one of its active components. Although it has been shown to have anti-inflammatory activity [171], it is still very far from clinical use. To improve its bioavailability and lipid solubility, Wu et al. [83] designed and synthesized a series of new gentiside derivatives, followed by in vitro inhibition of influenza virus biological evaluation of all synthesized compounds. 2′,3′,6′-Tri-O-benzoyl-4′- O-methylsulfonyl gentiopicroside (75), 4′-fluoro-4′-deoxy gentiopicroside (76), and 2′,3′,6′-Tri-O-benzoyl-4′,5′-olefin gentiopicroside (77) (Figure 10) showed significant activity against influenza viruses, with IC50 values of 39.5 μM, 45.2 μM, and 44.0 μM, respectively (Table 1) [83]. The authors further evaluated their cytotoxicity in MDCK cells, with results indicating no cytotoxic effects on uninfected MDCK cells at a concentration of 50µM. Triptolide (TP) is one of the main active substances of mine, a commonly used Chinese medicine in China. Although recent studies have demonstrated that TP has antiviral effects [172], TP has poor water solubility and rapid elimination in vivo, and high concentrations of TP can cause multiple organ toxicity, which hinders the development of its clinical application. The TP derivatives were designed and synthesized, and their activity against influenza A viruses was evaluated by Jiang et al. [84]; the results show that 4-(((5bS, 6aS, 7aR, 8R, 8aS, 9aS, 9bS, 10aS, 10bS)-8a-isopropyl-10 bmethyl-3-oxo 1, 2,3, 5,5b, 6,6a, 8,8a, 9a, 9b, 10b dodecahydrotris(oxireno) [2′, 3′:4b, 5;2″, 3″:6, 7;2, 3‴:8a, 9] phenanthrol [1, 2-c] furan-8-yl)oxy)-2, 2-dimethyl-4 oxobutanoic acid (TPDMSA) (78) (Figure 10) is an anti-influenza virus agent mainly by inhibiting the nuclear export of the influenza virus vRNP, acting at the late stage of the H1N1 viral replication cycle, with significant inhibition of NP activity late in the H1N1 viral replication cycle by binding to the NP tail loop active site. The EC50 value of TPDMSA (78) for A/WSN/33 (H1N1) was 3.24 μM (Table 1). Anastasiya S. Sokolova et al. found high anti-influenza activity of camphor imine derivatives in a previous study [173,174,175]. A series of compounds containing a 1,7,7-trimethylbicyclo [2.2.1] heptane fragment was then studied. The results showed that N, N, N-Trimethyl-2-oxo-2-((1S,2R,4S)-1,7,7-trimethylbicy-clo [2.2.1] heptan-2-yloxy) ethanaminium iodide (79) (Figure 10) exhibited potent anti-influenza activity against influenza A (H1N1) virus in vitro, with an IC50 = 2.4µM, and showed low toxicity (CC50 = 1311µM) (Table 1) [85]. Chen et al. [86] synthesized a series of andrographolide derivatives and investigated their anti-influenza activity, which showed that AL-1 (14-alpha-lipoyl andrographolide) (80) (Figure 10) showed potent anti-influenza activity against H1N1 virus in vitro, with EC50 = 7.2 µM. Li et al.’s [87] evaluation of the anti-influenza virus activity of the 50 synthesized resveratrol derivatives showed that R42 (81) (Figure 10) not only had high inhibitory activity against NA, with an IC50 = 3.56 µM (Table 1), but also against influenza virus in MDCK cells, implying that the mechanism of anti-influenza virus activity may be through NA inhibition. Lin et al. [88] designed 23 pterodontic acid derivatives for the structural modification of pterodontic acid derivatives isolated from Laggera pterodonta (DC.) Benth. The results showed that compound 15 (82) (Figure 10) had a better activity against H1N1 virus (IC50 = 9.92 µM). Xue et al. [89] synthesized a series of substituted arylglycosides as analogues of gastrodin and evaluated their anti-influenza activity. The most effective compound was methyl 4-fluoro-3-((2S,3R,4S,5R,6R)3,4,5-triacetoxy-6-(acetoxymethyl)-tetrahydro-2H-pyran-2-yloxy) benzoate (83)] (Figure 10), which showed significant inhibitory activity against influenza A virus A/FM/1/47(H1N1), with an IC50 = 34.4 μM. Tret‘yakova et al. [90] synthesized a series of compounds of the terpene series by introducing heterocyclic fragments into the diterterpene skeleton. Among them, compound 12 (84) (Figure 10), containing pyrrolidine fragments, showed the least toxicity to influenza A virus. The antiviral activity was the highest (IC50 = 5.0μM, CC50 = 643μM) (Table 1). Chukicheva et al. [91] performed a preliminary evaluation of the antiviral activity of terpenophenols and their N or O-containing derivatives, in which 2-(1,7,7-Trimethylbicyclo [2.2.1]hept-exo-2-yl)cyclohexa-2,5-dien-1,4-dione (85) (Figure 10) was the most active influenza NA inhibitor, with an IC50 = 0.5 µM (Table 1). Cui et al. [92] designed and synthesized four series of ferulic acid derivatives, among which the most effective compound was E)-3-(4-Hydroxy-3-methoxyphenyl)-1-(4-methylpiperazin-1-yl)- prop-2-en-1-one (86) (Figure 10), with an IC50 = 12.77 ± 0.47 μg/mL (Table 1).

4. Druggability

Natural products have long been recognized as important sources of clinically innovative drugs and valuable sources of drug design not only in the anti-influenza field but also in areas such as anti-tumor research [176]. Although natural products contain large amounts of biologically active groups, there are many difficulties in developing them into clinical drugs [177]. Compared with synthetic drugs, natural products usually have medicinal disadvantages such as poor water solubility, poor bioavailability, and weak biological relative activity. Therefore, through the traditional Chinese medicine systems pharmacology database and analysis platform (TCMSP) website, we summarized the ADME-related properties of some natural compounds to visually assess their medicinal properties (Table 2).
Lipinski’s five rules posit that molecular weight (MW < 500), lipophilicity (AlogP < 5), hydrogen bond donor count (Hdon < 5), hydrogen bond acceptor count (Hacc < 10), and the number of rotatable bonds (RBN < 10) are essential criteria for assessing drug likeness [178,179]. Excessive molecular weight can impact drug absorption in the intestine. With the exception of kaempferol-3-sophoroside, quercetin-3-sophoroside, rutin, nicotiflorin, and procyanidin B-2, the molecular weights (MW) of other compounds are all below 500 Daltons (Da). Compounds in Table 2 exhibit favorable lipophilicity, with the AlogP of all < 5. Perlolyrine, 1H-indole-3-carboxaldehyde, hispidulin, and others meet Lipinski’s criteria by having favorable hydrogen bond donor count (Hdon < 5) and hydrogen bond acceptor count (Hacc < 10) as well as and RBN < 10. OB is one of the most critical pharmacokinetic parameters, serving as a key indicator of the effectiveness of drug entry into the human circulatory system. Adequate oral bioavailability is fundamental for compounds to exhibit pharmacological activity. Luteolin (36.16%), perlolyrine (65.95%), papaverine (64.04%), and others in Table 2 demonstrate relatively high oral bioavailability. Moreover, except for 1H-indole-3-carboxaldehyde, DL (drug-likeness) is above 0.18, indicating good drug-likeness.
However, this is only a small part of the natural products against H1N1, and more of these demonstrate less druggability. Nonetheless, natural products have received much attention due to their abundant sources and potential anti-H1N1 applications. Therefore, researchers still need to overcome the problems of low solubility, poor stability, and low bioavailability. Nanomedicine applications can well solve these problems. Nanoparticle technology helps to improve the solubility, stability, bioavailability, target specificity, and bioactivity of natural products through different encapsulation techniques [180,181].

5. Conclusions and Outlooks

The emergence of influenza A virus has caused severe harm to individuals, societies, and countries. Currently, there are two strategies to combat influenza viruses: vaccine and antiviral drugs. To date, WHO believes that influenza virus vaccination remains the best way to control the spread of influenza in humans [182]. However, use of the vaccine has been limited by side effects and difficulties in transportation and storage [183]. Therefore, drug therapy is the most effective means to control the spread of influenza virus. Because influenza A virus is prone to antigenic mutations, inhibitors targeting virus surface proteins (such as NA inhibitors and M2 ion channel inhibitors) are prone to developing resistance despite the emergence of new antiviral drugs. The active ingredients in natural products and natural medicine have always been a hotspot of new drug development. They can play a role in preventing and treating influenza A by improving immune function in a number of ways. Therefore, natural products have unique advantages and broad prospects in the fight against influenza A.
In this review, different types of natural anti-influenza A products were classified. They included marine fungal metabolites, flavonoids, alkaloids, terpenoids, phenols, and polysaccharides, and their anti-influenza A activities were discussed. The inhibitory effect of flavonoids on viral NA causes flavonoid compounds to exhibit strong anti-influenza A activity, and the structure–activity relationship was reviewed clearly. Alkaloids have been reported as inhibitors of various viral protein synthesis, primarily targeting protein synthesis during the virus’ life cycle. In studies on the anti-influenza activity of papaverine, it was found to inhibit viral infection later in the life cycle of influenza virus. However, the reason underlying the changes in morphology of influenza virus is still not exactly understood. Interestingly, to find highly effective antiviral active ingredients from natural products, several marine natural products have gained attention, providing new ideas for the development of novel anti-influenza A virus drugs.
In conclusion, natural products are an important source for the discovery of lead compounds and drug candidates against influenza A virus. However, the current studies on the anti-influenza A (H1N1) virus activity of natural products mainly focus on the screening of active ingredients and in vitro studies. Therefore, we believe that the mechanism of anti-influenza A (H1N1) requires more in-depth study, and more in vivo and clinical studies are needed to demonstrate anti-influenza A activity to further study effective and safe anti-influenza A (H1N1) drugs. Moreover, many natural products still need to overcome the problems of low solubility, poor stability, low bioavailability, and other drug-formation properties. Although natural products combined with nanoparticle technology can solve these problems, they are mostly in the preclinical research stage and require further clinical research.

Author Contributions

Writing—original draft preparation, R.L., Q.H. and X.L. (Xiaokun Li); conceptualization and writing—review and editing, X.L. (Xinguang Liu); conceptualization and supervision and writing—review, W.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Scientific research project of Traditional Chinese Medicine of Henan Province (No. 20-21ZY3012), Traditional Chinese Medicine Research Project for COVID-19 of Henan Province (No. 2022ZYFY06), and National Natural Science Foundation of China (No. 82274274).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Smith, G.J.; Vijaykrishna, D.; Bahl, J.; Lycett, S.J.; Worobey, M.; Pybus, O.G.; Ma, S.K.; Cheung, C.L.; Raghwani, J.; Bhatt, S.; et al. Origins and evolutionary genomics of the 2009 swine-origin H1N1 influenza A epidemic. Nature 2009, 459, 1122–1125. [Google Scholar] [CrossRef] [PubMed]
  2. Taubenberger, J.K.; Morens, D.M. 1918 Influenza: The mother of all pandemics. Emerg. Infect. Dis. 2006, 12, 15–22. [Google Scholar] [CrossRef]
  3. Sharon, B.; Schleiss, M.R. Pandemic Influenza A (H1N1). Pediatr. Res. 2009, 66, 599. [Google Scholar] [CrossRef] [PubMed]
  4. Enserink, M.; Cohen, J. Virus of the year. The novel H1N1 influenza. Science 2009, 326, 1607. [Google Scholar] [CrossRef] [PubMed]
  5. Yin, H.; Jiang, N.; Shi, W.; Chi, X.; Liu, S.; Chen, J.L.; Wang, S. Development and Effects of Influenza Antiviral Drugs. Molecules 2021, 26, 810. [Google Scholar] [CrossRef]
  6. O’Hanlon, R.; Shaw, M.L. Baloxavir marboxil: The new influenza drug on the market. Curr. Opin. Virol. 2019, 35, 14–18. [Google Scholar] [CrossRef] [PubMed]
  7. De Clercq, E. Antiviral agents active against influenza A viruses. Nat. Rev. Drug Discov. 2006, 5, 1015–1025. [Google Scholar] [CrossRef] [PubMed]
  8. Ntie-Kang, F.; Svozil, D. Volume 2 Advanced Concepts and Applications. In An Enumeration of Natural Products from Microbial, Marine and Terrestrial Sources; Ntie-Kang, F., Ed.; De Gruyter: Berlin, Germany, 2022; pp. 1–30. [Google Scholar]
  9. Abdal Dayem, A.; Choi, H.Y.; Kim, Y.B.; Cho, S.G. Antiviral effect of methylated flavonol isorhamnetin against influenza. PLoS ONE 2015, 10, e0121610. [Google Scholar] [CrossRef]
  10. Paules, C.; Subbarao, K. Influenza. Lancet 2017, 390, 697–708. [Google Scholar] [CrossRef]
  11. Pleschka, S. Overview of influenza viruses. Curr. Top. Microbiol. Immunol. 2013, 370, 1–20. [Google Scholar] [CrossRef]
  12. Durães-Carvalho, R.; Salemi, M. In-depth phylodynamics, evolutionary analysis and in silico predictions of universal epitopes of Influenza A subtypes and Influenza B viruses. Mol. Phylogenet. Evol. 2018, 121, 174–182. [Google Scholar] [CrossRef]
  13. Huang, S.S.H.; Banner, D.; Paquette, S.G.; Leon, A.J.; Kelvin, A.A.; Kelvin, D.J. Pathogenic influenza B virus in the ferret model establishes lower respiratory tract infection. J. Gen. Virol. 2014, 95, 2127–2139. [Google Scholar] [CrossRef] [PubMed]
  14. Bodewes, R.; Morick, D.; de Mutsert, G.; Osinga, N.; Bestebroer, T.; van der Vliet, S.; Smits, S.L.; Kuiken, T.; Rimmelzwaan, G.F.; Fouchier, R.A.; et al. Recurring influenza B virus infections in seals. Emerg. Infect. Dis. 2013, 19, 511–512. [Google Scholar] [CrossRef] [PubMed]
  15. Goldhill, D.H.; Te Velthuis, A.J.W.; Fletcher, R.A.; Langat, P.; Zambon, M.; Lackenby, A.; Barclay, W.S. The mechanism of resistance to favipiravir in influenza. Proc. Natl. Acad. Sci. USA 2018, 115, 11613–11618. [Google Scholar] [CrossRef]
  16. Leneva, I.A.; Burtseva, E.I.; Yatsyshina, S.B.; Fedyakina, I.T.; Kirillova, E.S.; Selkova, E.P.; Osipova, E.; Maleev, V.V. Virus susceptibility and clinical effectiveness of anti-influenza drugs during the 2010–2011 influenza season in Russia. Int. J. Infect. Dis. 2016, 43, 77–84. [Google Scholar] [CrossRef]
  17. Haffizulla, J.; Hartman, A.; Hoppers, M.; Resnick, H.; Samudrala, S.; Ginocchio, C.; Bardin, M.; Rossignol, J.F. Effect of nitazoxanide in adults and adolescents with acute uncomplicated influenza: A double-blind, randomised, placebo-controlled, phase 2b/3 trial. Lancet Infect. Dis. 2014, 14, 609–618. [Google Scholar] [CrossRef]
  18. Pinto, L.H.; Holsinger, L.J.; Lamb, R.A. Influenza virus M2 protein has ion channel activity. Cell 1992, 69, 517–528. [Google Scholar] [CrossRef] [PubMed]
  19. Zhou, Z.; Liu, T.; Zhang, J.; Zhan, P.; Liu, X. Influenza A virus polymerase: An attractive target for next-generation anti-influenza therapeutics. Drug Discov. Today 2018, 23, 503–518. [Google Scholar] [CrossRef]
  20. Kumar, B.; Asha, K.; Khanna, M.; Ronsard, L.; Meseko, C.A.; Sanicas, M. The emerging influenza virus threat: Status and new prospects for its therapy and control. Arch. Virol. 2018, 163, 831–844. [Google Scholar] [CrossRef]
  21. Bantia, S.; Arnold, C.S.; Parker, C.D.; Upshaw, R.; Chand, P. Anti-influenza virus activity of peramivir in mice with single intramuscular injection. Antivir. Res. 2006, 69, 39–45. [Google Scholar] [CrossRef]
  22. Bright, R.A.; Medina, M.J.; Xu, X.; Perez-Oronoz, G.; Wallis, T.R.; Davis, X.M.; Povinelli, L.; Cox, N.J.; Klimov, A.I. Incidence of adamantane resistance among influenza A (H3N2) viruses isolated worldwide from 1994 to 2005: A cause for concern. Lancet 2005, 366, 1175–1181. [Google Scholar] [CrossRef] [PubMed]
  23. Bright, R.A.; Shay, D.K.; Shu, B.; Cox, N.J.; Klimov, A.I. Adamantane resistance among influenza A viruses isolated early during the 2005–2006 influenza season in the United States. JAMA 2006, 295, 891–894. [Google Scholar] [CrossRef] [PubMed]
  24. Palese, P.; Schulman, J.L.; Bodo, G.; Meindl, P. Inhibition of influenza and parainfluenza virus replication in tissue culture by 2-deoxy-2,3-dehydro-N-trifluoroacetylneuraminic acid (FANA). Virology 1974, 59, 490–498. [Google Scholar] [CrossRef] [PubMed]
  25. McClellan, K.; Perry, C.M. Oseltamivir: A review of its use in influenza. Drugs 2001, 61, 263–283. [Google Scholar] [CrossRef] [PubMed]
  26. Anuwongcharoen, N.; Shoombuatong, W.; Tantimongcolwat, T.; Prachayasittikul, V.; Nantasenamat, C. Exploring the chemical space of influenza neuraminidase inhibitors. PeerJ 2016, 4, e1958. [Google Scholar] [CrossRef] [PubMed]
  27. Cheer, S.M.; Wagstaff, A.J. Zanamivir: An update of its use in influenza. Drugs 2002, 62, 71–106. [Google Scholar] [CrossRef] [PubMed]
  28. Samson, M.; Pizzorno, A.; Abed, Y.; Boivin, G. Influenza virus resistance to neuraminidase inhibitors. Antivir. Res. 2013, 98, 174–185. [Google Scholar] [CrossRef] [PubMed]
  29. Reece, P.A. Neuraminidase inhibitor resistance in influenza viruses. J. Med. Virol. 2007, 79, 1577–1586. [Google Scholar] [CrossRef]
  30. Hussain, M.; Galvin, H.D.; Haw, T.Y.; Nutsford, A.N.; Husain, M. Drug resistance in influenza A virus: The epidemiology and management. Infect. Drug Resist. 2017, 10, 121–134. [Google Scholar] [CrossRef]
  31. Pflug, A.; Guilligay, D.; Reich, S.; Cusack, S. Structure of influenza A polymerase bound to the viral RNA promoter. Nature 2014, 516, 355–360. [Google Scholar] [CrossRef]
  32. Yuan, P.; Bartlam, M.; Lou, Z.; Chen, S.; Zhou, J.; He, X.; Lv, Z.; Ge, R.; Li, X.; Deng, T.; et al. Crystal structure of an avian influenza polymerase PAN reveals an endonuclease active site. Nature 2009, 458, 909–913. [Google Scholar] [CrossRef] [PubMed]
  33. Tarbet, E.B.; Maekawa, M.; Furuta, Y.; Babu, Y.S.; Morrey, J.D.; Smee, D.F. Combinations of favipiravir and peramivir for the treatment of pandemic influenza A/California/04/2009 (H1N1) virus infections in mice. Antivir. Res. 2012, 94, 103–110. [Google Scholar] [CrossRef] [PubMed]
  34. Byrn, R.A.; Jones, S.M.; Bennett, H.B.; Bral, C.; Clark, M.P.; Jacobs, M.D.; Kwong, A.D.; Ledeboer, M.W.; Leeman, J.R.; McNeil, C.F.; et al. Preclinical activity of VX-787, a first-in-class, orally bioavailable inhibitor of the influenza virus polymerase PB2 subunit. Antimicrob. Agents Chemother. 2015, 59, 1569–1582. [Google Scholar] [CrossRef] [PubMed]
  35. Noshi, T.; Kitano, M.; Taniguchi, K.; Yamamoto, A.; Omoto, S.; Baba, K.; Hashimoto, T.; Ishida, K.; Kushima, Y.; Hattori, K.; et al. In vitro characterization of baloxavir acid, a first-in-class cap-dependent endonuclease inhibitor of the influenza virus polymerase PA subunit. Antivir. Res. 2018, 160, 109–117. [Google Scholar] [CrossRef] [PubMed]
  36. Amorim, M.J.; Read, E.K.; Dalton, R.M.; Medcalf, L.; Digard, P. Nuclear export of influenza A virus mRNAs requires ongoing RNA polymerase II activity. Traffic 2007, 8, 1–11. [Google Scholar] [CrossRef] [PubMed]
  37. Cianci, C.; Gerritz, S.W.; Deminie, C.; Krystal, M. Influenza nucleoprotein: Promising target for antiviral chemotherapy. Antivir. Chem. Chemother. 2012, 23, 77–91. [Google Scholar] [CrossRef] [PubMed]
  38. Dilly, S.; Fotso Fotso, A.; Lejal, N.; Zedda, G.; Chebbo, M.; Rahman, F.; Companys, S.; Bertrand, H.C.; Vidic, J.; Noiray, M.; et al. From Naproxen Repurposing to Naproxen Analogues and Their Antiviral Activity against Influenza A Virus. J. Med. Chem. 2018, 61, 7202–7217. [Google Scholar] [CrossRef] [PubMed]
  39. Zheng, W.; Fan, W.; Zhang, S.; Jiao, P.; Shang, Y.; Cui, L.; Mahesutihan, M.; Li, J.; Wang, D.; Gao, G.F.; et al. Naproxen Exhibits Broad Anti-influenza Virus Activity in Mice by Impeding Viral Nucleoprotein Nuclear Export. Cell. Rep. 2019, 27, 1875–1885.e1875. [Google Scholar] [CrossRef] [PubMed]
  40. Kao, R.Y.; Yang, D.; Lau, L.S.; Tsui, W.H.; Hu, L.; Dai, J.; Chan, M.P.; Chan, C.M.; Wang, P.; Zheng, B.J.; et al. Identification of influenza A nucleoprotein as an antiviral target. Nat. Biotechnol. 2010, 28, 600–605. [Google Scholar] [CrossRef]
  41. Liu, C.L.; Hung, H.C.; Lo, S.C.; Chiang, C.H.; Chen, I.J.; Hsu, J.T.; Hou, M.H. Using mutagenesis to explore conserved residues in the RNA-binding groove of influenza A virus nucleoprotein for antiviral drug development. Sci. Rep. 2016, 6, 21662. [Google Scholar] [CrossRef]
  42. Hu, Y.; Sneyd, H.; Dekant, R.; Wang, J. Influenza A Virus Nucleoprotein: A Highly Conserved Multi-Functional Viral Protein as a Hot Antiviral Drug Target. Curr. Top. Med. Chem. 2017, 17, 2271–2285. [Google Scholar] [CrossRef] [PubMed]
  43. Blunt, J.W.; Carroll, A.R.; Copp, B.R.; Davis, R.A.; Keyzers, R.A.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2018, 35, 8–53. [Google Scholar] [CrossRef] [PubMed]
  44. Teng, Y.F.; Xu, L.; Wei, M.Y.; Wang, C.Y.; Gu, Y.C.; Shao, C.L. Recent progresses in marine microbial-derived antiviral natural products. Arch. Pharm. Res. 2020, 43, 1215–1229. [Google Scholar] [CrossRef] [PubMed]
  45. Osmanova, N.; Schultze, W.; Ayoub, N. Azaphilones: A class of fungal metabolites with diverse biological activities. Phytochem. Rev. 2010, 9, 315–342. [Google Scholar] [CrossRef]
  46. Gao, J.M.; Yang, S.X.; Qin, J.C. Azaphilones: Chemistry and biology. Chem. Rev. 2013, 113, 4755–4811. [Google Scholar] [CrossRef]
  47. Eade, R.A.; Page, H.; Robertson, A.; Turner, K.; Whalley, W.B. 986. The chemistry of fungi. Part XXVIII. Sclerotiorin and its hydrogenation products. J. Chem. Soc. Resumed 1957, 4913–4924. [Google Scholar] [CrossRef]
  48. Seto, H.; Tanabe, M. Utilization of 13C-13C coupling in structural and biosynthetic studies. III. Ochrephilone—A new fungal metabolite. Tetrahedron Lett. 1974, 15, 651–654. [Google Scholar] [CrossRef]
  49. Grassauer, A.; Weinmuellner, R.; Meier, C.; Pretsch, A.; Prieschl-Grassauer, E.; Unger, H. Iota-Carrageenan is a potent inhibitor of rhinovirus infection. Virol. J. 2008, 5, 107. [Google Scholar] [CrossRef] [PubMed]
  50. Hung, H.-C.; Tseng, C.-P.; Yang, J.-M.; Ju, Y.-W.; Tseng, S.-N.; Chen, Y.-F.; Chao, Y.-S.; Hsieh, H.-P.; Shih, S.-R.; Hsu, J.T.A. Aurintricarboxylic acid inhibits influenza virus neuraminidase. Antivir. Res. 2009, 81, 123–131. [Google Scholar] [CrossRef]
  51. Jia, Q.; Du, Y.; Wang, C.; Wang, Y.; Zhu, T.; Zhu, W. Azaphilones from the Marine Sponge-Derived Fungus Penicillium sclerotiorum OUCMDZ-3839. Mar. Drugs 2019, 17, 260. [Google Scholar] [CrossRef]
  52. He, F.; Bao, J.; Zhang, X.Y.; Tu, Z.C.; Shi, Y.M.; Qi, S.H. Asperterrestide A, a cytotoxic cyclic tetrapeptide from the marine-derived fungus Aspergillus terreus SCSGAF0162. J. Nat. Prod. 2013, 76, 1182–1186. [Google Scholar] [CrossRef] [PubMed]
  53. Zhu, T.; Chen, Z.; Liu, P.; Wang, Y.; Xin, Z.; Zhu, W. New rubrolides from the marine-derived fungus Aspergillus terreus OUCMDZ-1925. J. Antibiot. 2014, 67, 315–318. [Google Scholar] [CrossRef] [PubMed]
  54. Tian, Y.Q.; Lin, X.P.; Wang, Z.; Zhou, X.F.; Qin, X.C.; Kaliyaperumal, K.; Zhang, T.Y.; Tu, Z.C.; Liu, Y. Asteltoxins with Antiviral Activities from the Marine Sponge-Derived Fungus Aspergillus sp. SCSIO XWS02F40. Molecules 2015, 21, 34. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, H.; Chen, Z.; Zhu, G.; Wang, L.; Du, Y.; Wang, Y.; Zhu, W. Phenolic polyketides from the marine alga-derived Streptomyces sp. OUCMDZ-3434. Tetrahedron 2017, 73, 5451–5455. [Google Scholar] [CrossRef]
  56. Chen, D.; Lu, S.; Yang, G.; Pan, X.; Fan, S.; Xie, X.; Chen, Q.; Li, F.; Li, Z.; Wu, S.; et al. The seafood Musculus senhousei shows anti-influenza A virus activity by targeting virion envelope lipids. Biochem. Pharmacol. 2020, 177, 113982. [Google Scholar] [CrossRef] [PubMed]
  57. Gong, K.K.; Tang, X.L.; Zhang, G.; Cheng, C.L.; Zhang, X.W.; Li, P.L.; Li, G.Q. Polyhydroxylated steroids from the South China Sea soft coral Sarcophyton sp. and their cytotoxic and antiviral activities. Mar. Drugs 2013, 11, 4788–4798. [Google Scholar] [CrossRef]
  58. Ha, T.K.; Dao, T.T.; Nguyen, N.H.; Kim, J.; Kim, E.; Cho, T.O.; Oh, W.K. Antiviral phenolics from the leaves of Cleistocalyx operculatus. Fitoterapia 2016, 110, 135–141. [Google Scholar] [CrossRef]
  59. Huh, J.; Ha, T.K.Q.; Kang, K.B.; Kim, K.H.; Oh, W.K.; Kim, J.; Sung, S.H. C-Methylated Flavonoid Glycosides from Pentarhizidium orientale Rhizomes and Their Inhibitory Effects on the H1N1 Influenza Virus. J. Nat. Prod. 2017, 80, 2818–2824. [Google Scholar] [CrossRef]
  60. Lee, I.K.; Hwang, B.S.; Kim, D.W.; Kim, J.Y.; Woo, E.E.; Lee, Y.J.; Choi, H.J.; Yun, B.S. Characterization of Neuraminidase Inhibitors in Korean Papaver rhoeas Bee Pollen Contributing to Anti-Influenza Activities In Vitro. Planta Med. 2016, 82, 524–529. [Google Scholar] [CrossRef]
  61. Jeong, H.J.; Ryu, Y.B.; Park, S.J.; Kim, J.H.; Kwon, H.J.; Kim, J.H.; Park, K.H.; Rho, M.C.; Lee, W.S. Neuraminidase inhibitory activities of flavonols isolated from Rhodiola rosea roots and their in vitro anti-influenza viral activities. Bioorg. Med. Chem. 2009, 17, 6816–6823. [Google Scholar] [CrossRef]
  62. Roschek, B., Jr.; Fink, R.C.; McMichael, M.D.; Li, D.; Alberte, R.S. Elderberry flavonoids bind to and prevent H1N1 infection in vitro. Phytochemistry 2009, 70, 1255–1261. [Google Scholar] [CrossRef] [PubMed]
  63. Bang, S.; Quy Ha, T.K.; Lee, C.; Li, W.; Oh, W.K.; Shim, S.H. Antiviral activities of compounds from aerial parts of Salvia plebeia R. Br. J. Ethnopharmacol. 2016, 192, 398–405. [Google Scholar] [CrossRef] [PubMed]
  64. Li, M.; Wang, Y.; Jin, J.; Dou, J.; Guo, Q.; Ke, X.; Zhou, C.; Guo, M. Inhibitory Activity of Honeysuckle Extracts against Influenza A Virus In Vitro and In Vivo. Virol. Sin. 2021, 36, 490–500. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, G.; Zhang, B.; Zhang, X.; Bing, F. Homonojirimycin, an alkaloid from dayflower inhibits the growth of influenza A virus in vitro. Acta Virol. 2013, 57, 85–86. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, P.; Kong, F.; Wei, J.; Wang, Y.; Wang, W.; Hong, K.; Zhu, W. Alkaloids from the mangrove-derived actinomycete Jishengella endophytica 161111. Mar. Drugs 2014, 12, 477–490. [Google Scholar] [CrossRef] [PubMed]
  67. Gong, K.K.; Tang, X.L.; Liu, Y.S.; Li, P.L.; Li, G.Q. Imidazole Alkaloids from the South China Sea Sponge Pericharax heteroraphis and Their Cytotoxic and Antiviral Activities. Molecules 2016, 21, 150. [Google Scholar] [CrossRef] [PubMed]
  68. Moradi, M.T.; Karimi, A.; Rafieian-Kopaei, M.; Fotouhi, F. In vitro antiviral effects of Peganum harmala seed extract and its total alkaloids against Influenza virus. Microb. Pathog. 2017, 110, 42–49. [Google Scholar] [CrossRef] [PubMed]
  69. Aggarwal, M.; Leser, G.P.; Lamb, R.A. Repurposing Papaverine as an Antiviral Agent against Influenza Viruses and Paramyxoviruses. J. Virol. 2020, 94, e01888-19. [Google Scholar] [CrossRef] [PubMed]
  70. Bandaranayake, W.M. Bioactivities, bioactive compounds and chemical constituents of mangrove plants. Wetl. Ecol. Manag. 2002, 10, 421–452. [Google Scholar] [CrossRef]
  71. Gong, K.K.; Li, P.L.; Qiao, D.; Zhang, X.W.; Chu, M.J.; Qin, G.F.; Tang, X.L.; Li, G.Q. Cytotoxic and Antiviral Triterpenoids from the Mangrove Plant Sonneratia paracaseolaris. Molecules 2017, 22, 1319. [Google Scholar] [CrossRef]
  72. Zhu, Q.; Amen, Y.M.; Ohnuki, K.; Shimizu, K. Anti-influenza effects of Ganoderma lingzhi: An animal study. J. Funct. Foods 2017, 34, 224–228. [Google Scholar] [CrossRef]
  73. Yamamoto, T.; Izumi, N.; Ui, H.; Sueki, A.; Masuma, R.; Nonaka, K.; Hirose, T.; Sunazuka, T.; Nagai, T.; Yamada, H.; et al. Wickerols A and B: Novel anti-influenza virus diterpenes produced by Trichoderma atroviride FKI-3849. Tetrahedron 2012, 68, 9267–9271. [Google Scholar] [CrossRef]
  74. Yan, X.; Ouyang, H.; Wang, W.; Liu, J.; Li, T.; Wu, B.; Yan, X.; He, S. Antimicrobial Terpenoids from South China Sea Soft Coral Lemnalia sp. Mar. Drugs 2021, 19, 294. [Google Scholar] [CrossRef] [PubMed]
  75. Chen, C.-Y.; Yau, L.-F.; Mai, Z.-T.; Li, R.-F.; Yang, Z.-F.; Zhu, G.-Y.; Jiang, Z.-H.; Wang, J.-R. Aculeatusane A: A new diterpenoid from the whole plants of Celastrus Aculeatus Merr. Phytochem. Lett. 2020, 40, 72–75. [Google Scholar] [CrossRef]
  76. Parhira, S.; Yang, Z.F.; Zhu, G.Y.; Chen, Q.L.; Zhou, B.X.; Wang, Y.T.; Liu, L.; Bai, L.P.; Jiang, Z.H. In vitro anti-influenza virus activities of a new lignan glycoside from the latex of Calotropis gigantea. PLoS ONE 2014, 9, e104544. [Google Scholar] [CrossRef]
  77. Yang, Z.F.; Bai, L.P.; Huang, W.B.; Li, X.Z.; Zhao, S.S.; Zhong, N.S.; Jiang, Z.H. Comparison of in vitro antiviral activity of tea polyphenols against influenza A and B viruses and structure-activity relationship analysis. Fitoterapia 2014, 93, 47–53. [Google Scholar] [CrossRef] [PubMed]
  78. Kim, M.; Yim, J.H.; Kim, S.Y.; Kim, H.S.; Lee, W.G.; Kim, S.J.; Kang, P.S.; Lee, C.K. In vitro inhibition of influenza A virus infection by marine microalga-derived sulfated polysaccharide p-KG03. Antivir. Res. 2012, 93, 253–259. [Google Scholar] [CrossRef] [PubMed]
  79. Jiao, G.; Yu, G.; Wang, W.; Zhao, X.; Zhang, J.; Ewart, S.H. Properties of polysaccharides in several seaweeds from Atlantic Canada and their potential anti-influenza viral activities. J. Ocean. Univ. China 2012, 11, 205–212. [Google Scholar] [CrossRef]
  80. Yu, G.; Li, M.; Wang, W.; Liu, X.; Zhao, X.; Lv, Y.; Li, G.; Jiao, G.; Zhao, X. Structure and anti-influenza A (H1N1) virus activity of three polysaccharides from Eucheuma denticulatum. J. Ocean. Univ. China 2012, 11, 527–532. [Google Scholar] [CrossRef]
  81. Peng, M.H.; Dai, W.P.; Liu, S.J.; Yu, L.W.; Wu, Y.N.; Liu, R.; Chen, X.L.; Lai, X.P.; Li, X.; Zhao, Z.X.; et al. Bioactive glycosides from the roots of Ilex asprella. Pharm. Biol. 2016, 54, 2127–2134. [Google Scholar] [CrossRef]
  82. Lee, B.W.; Ha, T.K.Q.; Cho, H.M.; An, J.P.; Kim, S.K.; Kim, C.S.; Kim, E.; Oh, W.K. Antiviral activity of furanocoumarins isolated from Angelica dahurica against influenza a viruses H1N1 and H9N2. J. Ethnopharmacol. 2020, 259, 112945. [Google Scholar] [CrossRef] [PubMed]
  83. Wu, S.; Yang, L.; Sun, W.; Si, L.; Xiao, S.; Wang, Q.; Dechoux, L.; Thorimbert, S.; Sollogoub, M.; Zhou, D.; et al. Design, synthesis and biological evaluation of gentiopicroside derivatives as potential antiviral inhibitors. Eur. J. Med. Chem. 2017, 130, 308–319. [Google Scholar] [CrossRef] [PubMed]
  84. Jiang, N.; Quan, L.; Zhou, Y.; Cheng, Y.; Li, H.; Chen, X.; Li, R.; Liu, D. Exploring the anti-influenza virus activity of novel triptolide derivatives targeting nucleoproteins. Bioorg. Chem. 2022, 129, 106118. [Google Scholar] [CrossRef] [PubMed]
  85. Sokolova, A.S.; Yarovaya, O.I.; Baranova, D.V.; Galochkina, A.V.; Shtro, A.A.; Kireeva, M.V.; Borisevich, S.S.; Gatilov, Y.V.; Zarubaev, V.V.; Salakhutdinov, N.F. Quaternary ammonium salts based on (-)-borneol as effective inhibitors of influenza virus. Arch. Virol. 2021, 166, 1965–1976. [Google Scholar] [CrossRef] [PubMed]
  86. Chen, J.X.; Xue, H.J.; Ye, W.C.; Fang, B.H.; Liu, Y.H.; Yuan, S.H.; Yu, P.; Wang, Y.Q. Activity of andrographolide and its derivatives against influenza virus in vivo and in vitro. Biol. Pharm. Bull. 2009, 32, 1385–1391. [Google Scholar] [CrossRef] [PubMed]
  87. Li, C.; Fang, J.S.; Lian, W.W.; Pang, X.C.; Liu, A.L.; Du, G.H. In vitro antiviral effects and 3D QSAR study of resveratrol derivatives as potent inhibitors of influenza H1N1 neuraminidase. Chem. Biol. Drug Des. 2015, 85, 427–438. [Google Scholar] [CrossRef] [PubMed]
  88. Lin, Y.J.; Huang, B.L.; Dai, S.Y.; Song, L.D.; Yu, H.F.; Yu, X.L.; Zhang, R.P. Synthesis of pterodontic acid derivatives and the study of their anti-influenza A virus (H1N1) activity. Fitoterapia 2021, 152, 104942. [Google Scholar] [CrossRef]
  89. Xue, S.T.; He, W.Y.; Ma, L.L.; Wang, H.Q.; Wang, B.; Zheng, G.H.; Ji, X.Y.; Zhang, T.; Li, Y.H.; Jiang, J.D.; et al. Synthesis and anti-influenza virus activities of a novel class of gastrodin derivatives. Molecules 2013, 18, 3789–3805. [Google Scholar] [CrossRef] [PubMed]
  90. Tret’yakova, E.V.; Ma, X.; Kazakova, O.B.; Shtro, A.A.; Petukhova, G.D.; Klabukov, A.M.; Dyatlov, D.S.; Smirnova, A.A.; Xu, H.; Xiao, S. Synthesis and evaluation of diterpenic Mannich bases as antiviral agents against influenza A and SARS-CoV-2. Phytochem. Lett. 2022, 51, 91–96. [Google Scholar] [CrossRef]
  91. Chukicheva, I.Y.; Buravlev, E.V.; Dvornikova, I.A.; Fedorova, I.V.; Zarubaev, V.V.; Slita, A.V.; Esaulkova, Y.L.; Kutchin, A.V. Evaluation of antiviral activity of terpenophenols and some of their N- and O-derivatives. Russ. Chem. Bull. 2022, 71, 2473–2481. [Google Scholar] [CrossRef]
  92. Cui, M.Y.; Xiao, M.W.; Xu, L.J.; Chen, Y.; Liu, A.L.; Ye, J.; Hu, A.X. Bioassay of ferulic acid derivatives as influenza neuraminidase inhibitors. Arch. Pharm. 2020, 353, e1900174. [Google Scholar] [CrossRef] [PubMed]
  93. Mai, T.T.; Chuyen, N.V. Anti-Hyperglycemic Activity of an Aqueous Extract from Flower Buds of Cleistocalyx operculatus (Roxb.) Merr and Perry. Biosci. Biotechnol. Biochem. 2007, 71, 69–76. [Google Scholar] [CrossRef] [PubMed]
  94. Dao, T.T.; Tung, B.T.; Nguyen, P.H.; Thuong, P.T.; Yoo, S.S.; Kim, E.H.; Kim, S.K.; Oh, W.K. C-Methylated Flavonoids from Cleistocalyx operculatus and Their Inhibitory Effects on Novel Influenza A (H1N1) Neuraminidase. J. Nat. Prod. 2010, 73, 1636–1642. [Google Scholar] [CrossRef]
  95. Min, B.S.; Thu, C.V.; Dat, N.T.; Dang, N.H.; Jang, H.S.; Hung, T.M. Antioxidative flavonoids from Cleistocalyx operculatus buds. Chem. Pharm. Bull. 2008, 56, 1725–1728. [Google Scholar] [CrossRef]
  96. Min, B.S.; Cuong, T.D.; Lee, J.S.; Shin, B.S.; Woo, M.H.; Hung, T.M. Cholinesterase inhibitors from Cleistocalyx operculatus buds. Arch. Pharm. Res. 2010, 33, 1665–1670. [Google Scholar] [CrossRef]
  97. Li, B.; Ni, Y.; Zhu, L.J.; Wu, F.B.; Yan, F.; Zhang, X.; Yao, X.S. Flavonoids from Matteuccia struthiopteris and Their Anti-influenza Virus (H1N1) Activity. J. Nat. Prod. 2015, 78, 987–995. [Google Scholar] [CrossRef]
  98. Basnet, P.; Kadota, S.; Shimizu, M.; Xu, H.X.; Namba, T. 2′-Hydroxymatteucinol, a new C-methyl flavanone derivative from Matteccia orientalis; potent hypoglycemic activity in streptozotocin (STZ)-induced diabetic rat. Chem. Pharm. Bull. 1993, 41, 1790–1795. [Google Scholar] [CrossRef]
  99. Pu, W.-L.; Zhang, M.-Y.; Bai, R.-Y.; Sun, L.-K.; Li, W.-H.; Yu, Y.-L.; Zhang, Y.; Song, L.; Wang, Z.-X.; Peng, Y.-F.; et al. Anti-inflammatory effects of Rhodiola rosea L.: A review. Biomed. Pharmacother. 2020, 121, 109552. [Google Scholar] [CrossRef] [PubMed]
  100. Chen, Y.; Tang, M.; Yuan, S.; Fu, S.; Li, Y.; Li, Y.; Wang, Q.; Cao, Y.; Liu, L.; Zhang, Q. Rhodiola rosea: A Therapeutic Candidate on Cardiovascular Diseases. Oxidative Med. Cell. Longev. 2022, 2022, 1348795. [Google Scholar] [CrossRef]
  101. Amsterdam, J.D.; Panossian, A.G. Rhodiola rosea L. as a putative botanical antidepressant. Phytomedicine 2016, 23, 770–783. [Google Scholar] [CrossRef]
  102. Liu, A.L.; Wang, H.D.; Lee, S.M.; Wang, Y.T.; Du, G.H. Structure-activity relationship of flavonoids as influenza virus neuraminidase inhibitors and their in vitro anti-viral activities. Bioorg. Med. Chem. 2008, 16, 7141–7147. [Google Scholar] [CrossRef] [PubMed]
  103. de Pascual-Teresa, S.; Santos-Buelga, C.; Rivas-Gonzalo, J.C. Quantitative analysis of flavan-3-ols in Spanish foodstuffs and beverages. J. Agric. Food Chem. 2000, 48, 5331–5337. [Google Scholar] [CrossRef] [PubMed]
  104. Barak, V.; Halperin, T.; Kalickman, I. The effect of Sambucol, a black elderberry-based, natural product, on the production of human cytokines: I. Inflammatory cytokines. Eur. Cytokine Netw. 2001, 12, 290–296. [Google Scholar] [PubMed]
  105. Santin, J.R.; Benvenutti, L.; Broering, M.F.; Nunes, R.; Goldoni, F.C.; Patel, Y.B.K.; de Souza, J.A.; Kopp, M.A.T.; de Souza, P.; da Silva, R.C.V.; et al. Sambucus nigra: A traditional medicine effective in reducing inflammation in mice. J. Ethnopharmacol. 2022, 283, 114736. [Google Scholar] [CrossRef] [PubMed]
  106. Zakay-Rones, Z.; Thom, E.; Wollan, T.; Wadstein, J. Randomized study of the efficacy and safety of oral elderberry extract in the treatment of influenza A and B virus infections. J. Int. Med. Res. 2004, 32, 132–140. [Google Scholar] [CrossRef]
  107. Imanishi, N.; Tuji, Y.; Katada, Y.; Maruhashi, M.; Konosu, S.; Mantani, N.; Terasawa, K.; Ochiai, H. Additional inhibitory effect of tea extract on the growth of influenza A and B viruses in MDCK cells. Microbiol. Immunol. 2002, 46, 491–494. [Google Scholar] [CrossRef]
  108. Nagai, T.; Moriguchi, R.; Suzuki, Y.; Tomimori, T.; Yamada, H. Mode of action of the anti-influenza virus activity of plant flavonoid, 5,7,4′-trihydroxy-8-methoxyflavone, from the roots of Scutellaria baicalensis. Antivir. Res. 1995, 26, 11–25. [Google Scholar] [CrossRef] [PubMed]
  109. Knox, Y.M.; Hayashi, K.; Suzutani, T.; Ogasawara, M.; Yoshida, I.; Shiina, R.; Tsukui, A.; Terahara, N.; Azuma, M. Activity of anthocyanins from fruit extract of Ribes nigrum L. against influenza A and B viruses. Acta Virol. 2001, 45, 209–215. [Google Scholar]
  110. Song, J.M.; Lee, K.H.; Seong, B.L. Antiviral effect of catechins in green tea on influenza virus. Antivir. Res. 2005, 68, 66–74. [Google Scholar] [CrossRef]
  111. Serkedjieva, J.; Manolova, N.; Zgórniak-Nowosielska, I.; Zawilińska, B.; Grzybek, J. Antiviral activity of the infusion (SHS-174) from flowers of Sambucus nigra L., aerial parts of Hypericum perforatum L., and roots of Saponaria officinalis L. against influenza and herpes simplex viruses. Phytother. Res. 1990, 4, 97–100. [Google Scholar] [CrossRef]
  112. Cao, Y.X.; Ji, P.; Wu, F.L.; Dong, J.Q.; Li, C.C.; Ma, T.; Yang, H.C.; Wei, Y.M.; Hua, Y.L. Lonicerae japonicae caulis: A review of its research progress of active metabolites and pharmacological effects. Front. Pharmacol. 2023, 14, 1277283. [Google Scholar] [CrossRef] [PubMed]
  113. Zhou, Z.; Li, X.; Liu, J.; Dong, L.; Chen, Q.; Liu, J.; Kong, H.; Zhang, Q.; Qi, X.; Hou, D.; et al. Honeysuckle-encoded atypical microRNA2911 directly targets influenza A viruses. Cell Res. 2015, 25, 39–49. [Google Scholar] [CrossRef] [PubMed]
  114. Afifi, F.U.; Kasabri, V.; Al-Jaber, H.I.; Abu-Irmaileh, B.E.; Al-Qudah, M.A.; Abazaa, I.F. Composition and Biological Activity of Volatile Oil from Salviajudaica and S. multicaulis from Jordan. Nat. Prod. Commun. 2016, 11, 535–538. [Google Scholar] [PubMed]
  115. Afonso, A.F.; Pereira, O.R.; Fernandes, Â.; Calhelha, R.C.; Silva, A.M.S.; Ferreira, I.; Cardoso, S.M. Phytochemical Composition and Bioactive Effects of Salvia africana, Salvia officinalisIcterina’ and Salvia mexicana Aqueous Extracts. Molecules 2019, 24, 4327. [Google Scholar] [CrossRef] [PubMed]
  116. Krol, A.; Kokotkiewicz, A.; Luczkiewicz, M. White Sage (Salvia apiana)—A Ritual and Medicinal Plant of the Chaparral: Plant Characteristics in Comparison with Other Salvia Species. Planta Med. 2022, 88, 604–627. [Google Scholar] [CrossRef] [PubMed]
  117. Xie, Y.; Huang, B.; Yu, K.; Shi, F.; Liu, T.; Xu, W. Caffeic acid derivatives: A new type of influenza neuraminidase inhibitors. Bioorg. Med. Chem. Lett. 2013, 23, 3556–3560. [Google Scholar] [CrossRef] [PubMed]
  118. Morimoto, R.; Hanada, A.; Matsubara, C.; Horio, Y.; Sumitani, H.; Ogata, T.; Isegawa, Y. Anti-influenza A virus activity of flavonoids in vitro: A structure-activity relationship. J. Nat. Med. 2023, 77, 219–227. [Google Scholar] [CrossRef] [PubMed]
  119. Abookleesh, F.L.; Al-Anzi, B.S.; Ullah, A. Potential Antiviral Action of Alkaloids. Molecules 2022, 27, 903. [Google Scholar] [CrossRef]
  120. Youn, J.Y.; Park, H.Y.; Cho, K.H. Anti-hyperglycemic activity of Commelina communis L.: Inhibition of alpha-glucosidase. Diabetes Res. Clin. Pract. 2004, 66 (Suppl. 1), S149–S155. [Google Scholar] [CrossRef]
  121. Yang, Q.; Ye, G.; Zhao, W.-M. Chemical Constituents of Commelina communis Linn. Biochem. Syst. Ecol. 2007, 35, 621–623. [Google Scholar] [CrossRef]
  122. Hong, K.; Gao, A.H.; Xie, Q.Y.; Gao, H.; Zhuang, L.; Lin, H.P.; Yu, H.P.; Li, J.; Yao, X.S.; Goodfellow, M.; et al. Actinomycetes for marine drug discovery isolated from mangrove soils and plants in China. Mar. Drugs 2009, 7, 24–44. [Google Scholar] [CrossRef] [PubMed]
  123. Xie, Q.Y.; Wang, C.; Wang, R.; Qu, Z.; Lin, H.P.; Goodfellow, M.; Hong, K. Jishengella endophytica gen. nov., sp. nov., a new member of the family Micromonosporaceae. Int. J. Syst. Evol. Microbiol. 2011, 61, 1153–1159. [Google Scholar] [CrossRef] [PubMed]
  124. Gross, H.; Kehraus, S.; König, G.M.; Woerheide, G.; Wright, A.D. New and biologically active imidazole alkaloids from two sponges of the genus Leucetta. J. Nat. Prod. 2002, 65, 1190–1193. [Google Scholar] [CrossRef] [PubMed]
  125. Plubrukarn, A.; Smith, D.W.; Cramer, R.E.; Davidson, B.S. (2E,9E)-pyronaamidine 9-(N-methylimine), a new imidazole alkaloid from the northern Mariana islands sponge Leucetta sp. cf. chagosensis. J. Nat. Prod. 1997, 60, 712–715. [Google Scholar] [CrossRef] [PubMed]
  126. Tsukamoto, S.; Kawabata, T.; Kato, H.; Ohta, T.; Rotinsulu, H.; Mangindaan, R.E.; van Soest, R.W.; Ukai, K.; Kobayashi, H.; Namikoshi, M. Naamidines H and I, cytotoxic imidazole alkaloids from the Indonesian marine sponge Leucetta chagosensis. J. Nat. Prod. 2007, 70, 1658–1660. [Google Scholar] [CrossRef] [PubMed]
  127. Ralifo, P.; Crews, P. A new structural theme in the imidazole-containing alkaloids from a calcareous Leucetta sponge. J. Org. Chem. 2004, 69, 9025–9029. [Google Scholar] [CrossRef] [PubMed]
  128. Fu, X.; Schmitz, F.J.; Tanner, R.S.; Kelly-Borges, M. New imidazole alkaloids and zinc complexes from the micronesian sponge Leucetta cf. chagosensis. J. Nat. Prod. 1998, 61, 384–386. [Google Scholar] [CrossRef] [PubMed]
  129. Chan, G.W.; Mong, S.; Hemling, M.E.; Freyer, A.J.; Offen, P.H.; DeBrosse, C.W.; Sarau, H.M.; Westley, J.W. New leukotriene B4 receptor antagonist: Leucettamine A and related imidazole alkaloids from the marine sponge Leucetta microraphis. J. Nat. Prod. 1993, 56, 116–121. [Google Scholar] [CrossRef]
  130. Copp, B.R.; Fairchild, C.R.; Cornell, L.; Casazza, A.M.; Robinson, S.; Ireland, C.M. Naamidine A is an antagonist of the epidermal growth factor receptor and an in vivo active antitumor agent. J. Med. Chem. 1998, 41, 3909–3911. [Google Scholar] [CrossRef]
  131. Akee, R.K.; Carroll, T.R.; Yoshida, W.Y.; Scheuer, P.J.; Stout, T.J.; Clardy, J. Two imidazole alkaloids from a sponge. J. Org. Chem. 1990, 55, 1944–1946. [Google Scholar] [CrossRef]
  132. Crews, P.; Clark, D.P.; Tenney, K. Variation in the Alkaloids among Indo-Pacific Leucetta Sponges. J. Nat. Prod. 2003, 66, 177–182. [Google Scholar] [CrossRef]
  133. Zhao, T.; Wang, Z.T.; Branford-White, C.J.; Xu, H.; Wang, C.H. Classification and differentiation of the genus Peganum indigenous to China based on chloroplast trnL-F and psbA-trnH sequences and seed coat morphology. Plant Biol. 2011, 13, 940–947. [Google Scholar] [CrossRef] [PubMed]
  134. Mina, C.N.; Farzaei, M.H.; Gholamreza, A. Medicinal properties of Peganum harmala L. in traditional Iranian medicine and modern phytotherapy: A review. J. Tradit. Chin. Med. 2015, 35, 104–109. [Google Scholar] [CrossRef]
  135. Moloudizargari, M.; Mikaili, P.; Aghajanshakeri, S.; Asghari, M.H.; Shayegh, J. Pharmacological and therapeutic effects of Peganum harmala and its main alkaloids. Pharmacogn. Rev. 2013, 7, 199–212. [Google Scholar] [CrossRef]
  136. Liu, H.M.; Tu, Y.K. The efficacy of papaverine administration by different routes for the treatment of experimental acute cerebral vasospasm. J. Clin. Neurosci. 2002, 9, 561–565. [Google Scholar] [CrossRef] [PubMed]
  137. Chappie, T.A.; Humphrey, J.M.; Allen, M.P.; Estep, K.G.; Fox, C.B.; Lebel, L.A.; Liras, S.; Marr, E.S.; Menniti, F.S.; Pandit, J.; et al. Discovery of a series of 6,7-dimethoxy-4-pyrrolidylquinazoline PDE10A inhibitors. J. Med. Chem. 2007, 50, 182–185. [Google Scholar] [CrossRef]
  138. Yildiz, N.; Gokkaya, N.K.; Koseoglu, F.; Gokkaya, S.; Comert, D. Efficacies of papaverine and sildenafil in the treatment of erectile dysfunction in early-stage paraplegic men. Int. J. Rehabil. Res. 2011, 34, 44–52. [Google Scholar] [CrossRef] [PubMed]
  139. Wu, X.; Liu, Y.; Sheng, W.; Sun, J.; Qin, G. Chemical constituents of Isatis indigotica. Planta Med. 1997, 63, 55–57. [Google Scholar] [CrossRef]
  140. Guo, Q.; Li, D.; Xu, C.; Zhu, C.; Guo, Y.; Yu, H.; Wang, X.; Shi, J. Indole alkaloid glycosides with a 1′-(phenyl)ethyl unit from Isatis indigotica leaves. Acta Pharm. Sin. B 2020, 10, 895–902. [Google Scholar] [CrossRef]
  141. Yang, W.; Chen, X.; Li, Y.; Guo, S.; Wang, Z.; Yu, X. Advances in Pharmacological Activities of Terpenoids. Nat. Prod. Commun. 2020, 15, 1934578X20903555. [Google Scholar] [CrossRef]
  142. Patra, A.; Chaudhuri, S.K.; Panda, S.K. Betulin-3-Caffeate from Quercus suber, 13C-nmr Spectra of Some Lupenes. J. Nat. Prod. 1988, 51, 217–220. [Google Scholar] [CrossRef]
  143. Bishop, K.S.; Kao, C.H.; Xu, Y.; Glucina, M.P.; Paterson, R.R.; Ferguson, L.R. From 2000 years of Ganoderma lucidum to recent developments in nutraceuticals. Phytochemistry 2015, 114, 56–65. [Google Scholar] [CrossRef]
  144. Cao, Y.; Wu, S.-H.; Dai, Y.-C. Species clarification of the prize medicinal Ganoderma mushroom “Lingzhi”. Fungal Divers. 2012, 56, 49–62. [Google Scholar] [CrossRef]
  145. Paterson, R.R. Ganoderma—A therapeutic fungal biofactory. Phytochemistry 2006, 67, 1985–2001. [Google Scholar] [CrossRef] [PubMed]
  146. Wu, G.S.; Guo, J.J.; Bao, J.L.; Li, X.W.; Chen, X.P.; Lu, J.J.; Wang, Y.T. Anti-cancer properties of triterpenoids isolated from Ganoderma lucidum—A review. Expert Opin. Investig. Drugs 2013, 22, 981–992. [Google Scholar] [CrossRef]
  147. Fatmawati, S.; Shimizu, K.; Kondo, R. Ganoderol B: A potent α-glucosidase inhibitor isolated from the fruiting body of Ganoderma lucidum. Phytomedicine 2011, 18, 1053–1055. [Google Scholar] [CrossRef]
  148. Liang, L.F.; Chen, W.T.; Mollo, E.; Yao, L.G.; Wang, H.Y.; Xiao, W.; Guo, Y.W. Sarcophytrols G–L, Novel Minor Metabolic Components from South China Sea Soft Coral Sarcophyton trocheliophorum Marenzeller. Chem. Biodivers. 2017, 14, e1700079. [Google Scholar] [CrossRef]
  149. Wang, Q.; Tang, X.; Liu, H.; Luo, X.; Sung, P.J.; Li, P.; Li, G. Clavukoellians G-K, New Nardosinane and Aristolane Sesquiterpenoids with Angiogenesis Promoting Activity from the Marine Soft Coral Lemnalia sp. Mar. Drugs 2020, 18, 171. [Google Scholar] [CrossRef] [PubMed]
  150. Wu, Q.; Sun, J.; Chen, J.; Zhang, H.; Guo, Y.W.; Wang, H. Terpenoids from Marine Soft Coral of the Genus Lemnalia: Chemistry and Biological Activities. Mar. Drugs 2018, 16, 320. [Google Scholar] [CrossRef]
  151. Tang, W.-H.; Bai, S.-T.; Tong, L.; Duan, W.-J.; Su, J.-W.; Chen, J.-X.; Xie, Y. Chemical constituents from Celastrus aculeatus Merr. Biochem. Syst. Ecol. 2014, 54, 78–82. [Google Scholar] [CrossRef]
  152. Tong, L.; Moudgil, K.D. Celastrus aculeatus Merr. suppresses the induction and progression of autoimmune arthritis by modulating immune response to heat-shock protein 65. Arthritis Res. Ther. 2007, 9, R70. [Google Scholar] [CrossRef] [PubMed]
  153. Dang, Z.; Jung, K.; Zhu, L.; Xie, H.; Lee, K.-H.; Chen, C.-H.; Huang, L. Phenolic Diterpenoid Derivatives as Anti-Influenza A Virus Agents. ACS Med. Chem. Lett. 2015, 6, 355–358. [Google Scholar] [CrossRef]
  154. Shahidi, F.; Yeo, J. Bioactivities of Phenolics by Focusing on Suppression of Chronic Diseases: A Review. Int. J. Mol. Sci. 2018, 19, 1573. [Google Scholar] [CrossRef]
  155. Pari, K.; Rao, P.J.; Devakumar, C.; Rastogi, J.N. A Novel Insect Antifeedant Nonprotein Amino Acid from Calotropis gigantea. J. Nat. Prod. 1998, 61, 102–104. [Google Scholar] [CrossRef]
  156. Lhinhatrakool, T.; Sutthivaiyakit, S. 19-Nor- and 18,20-epoxy-cardenolides from the leaves of Calotropis gigantea. J. Nat. Prod. 2006, 69, 1249–1251. [Google Scholar] [CrossRef]
  157. Silva, M.C.C.; da Silva, A.B.; Teixeira, F.M.; de Sousa, P.C.P.; Rondon, R.M.M.; Honório, J.E.R.; Sampaio, L.R.L.; Oliveira, S.L.; Holonda, A.N.M.; de Vasconcelos, S.M.M. Therapeutic and biological activities of Calotropis procera (Ait.) R. Br. Asian Pacific Journal of Tropical Medicine 2010, 3, 332–336. [Google Scholar] [CrossRef]
  158. Parhira, S.; Zhu, G.Y.; Jiang, R.W.; Liu, L.; Bai, L.P.; Jiang, Z.H. 2′-Epi-uscharin from the latex of Calotropis gigantea with HIF-1 inhibitory activity. Sci. Rep. 2014, 4, 4748. [Google Scholar] [CrossRef] [PubMed]
  159. Wurzer, W.J.; Ehrhardt, C.; Pleschka, S.; Berberich-Siebelt, F.; Wolff, T.; Walczak, H.; Planz, O.; Ludwig, S. NF-kappaB-dependent induction of tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) and Fas/FasL is crucial for efficient influenza virus propagation. J. Biol. Chem. 2004, 279, 30931–30937. [Google Scholar] [CrossRef]
  160. Chen, Y.; Cheng, S.; Dai, J.; Wang, L.; Xu, Y.; Peng, X.; Xie, X.; Peng, C. Molecular mechanisms and applications of tea polyphenols: A narrative review. J. Food Biochem. 2021, 45, e13910. [Google Scholar] [CrossRef]
  161. Talarico, L.B.; Pujol, C.A.; Zibetti, R.G.; Faría, P.C.; Noseda, M.D.; Duarte, M.E.; Damonte, E.B. The antiviral activity of sulfated polysaccharides against dengue virus is dependent on virus serotype and host cell. Antivir. Res. 2005, 66, 103–110. [Google Scholar] [CrossRef] [PubMed]
  162. Hasui, M.; Matsuda, M.; Okutani, K.; Shigeta, S. In vitro antiviral activities of sulfated polysaccharides from a marine microalga (Cochlodinium polykrikoides) against human immunodeficiency virus and other enveloped viruses. Int. J. Biol. Macromol. 1995, 17, 293–297. [Google Scholar] [CrossRef] [PubMed]
  163. Hosoya, M.; Balzarini, J.; Shigeta, S.; De Clercq, E. Differential inhibitory effects of sulfated polysaccharides and polymers on the replication of various myxoviruses and retroviruses, depending on the composition of the target amino acid sequences of the viral envelope glycoproteins. Antimicrob. Agents Chemother. 1991, 35, 2515–2520. [Google Scholar] [CrossRef] [PubMed]
  164. Yim, J.H.; Lee, H.K. Axenic culture of Gyrodinium impudicum strain KG03, a marine red-tide microalga that produces exopolysaccharide. J. Microbiol. 2004, 42, 305–314. [Google Scholar] [PubMed]
  165. Ghosh, T.; Chattopadhyay, K.; Marschall, M.; Karmakar, P.; Mandal, P.; Ray, B. Focus on antivirally active sulfated polysaccharides: From structure-activity analysis to clinical evaluation. Glycobiology 2009, 19, 2–15. [Google Scholar] [CrossRef] [PubMed]
  166. Damonte, E.B.; Matulewicz, M.C.; Cerezo, A.S. Sulfated seaweed polysaccharides as antiviral agents. Curr. Med. Chem. 2004, 11, 2399–2419. [Google Scholar] [CrossRef]
  167. Talarico, L.B.; Damonte, E.B. Interference in dengue virus adsorption and uncoating by carrageenans. Virology 2007, 363, 473–485. [Google Scholar] [CrossRef] [PubMed]
  168. Jiao, G.; Yu, G.; Zhang, J.; Ewart, H.S. Chemical structures and bioactivities of sulfated polysaccharides from marine algae. Mar. Drugs 2011, 9, 196–223. [Google Scholar] [CrossRef]
  169. Witvrouw, M.; De Clercq, E. Sulfated polysaccharides extracted from sea algae as potential antiviral drugs. Gen. Pharmacol. 1997, 29, 497–511. [Google Scholar] [CrossRef]
  170. Bade, R.; Chan, H.F.; Reynisson, J. Characteristics of known drug space. Natural products, their derivatives and synthetic drugs. Eur. J. Med. Chem. 2010, 45, 5646–5652. [Google Scholar] [CrossRef]
  171. Wang, Y.M.; Xu, M.; Wang, D.; Yang, C.R.; Zeng, Y.; Zhang, Y.J. Anti-inflammatory compounds of “Qin-Jiao”, the roots of Gentiana dahurica (Gentianaceae). J. Ethnopharmacol. 2013, 147, 341–348. [Google Scholar] [CrossRef]
  172. Jin, H.; Li, D.; Lin, M.H.; Li, L.; Harrich, D. Tat-Based Therapies as an Adjuvant for an HIV-1 Functional Cure. Viruses 2020, 12, 415. [Google Scholar] [CrossRef]
  173. Sokolova, A.S.; Yarovaya, O.I.; Shernyukov, A.V.; Gatilov, Y.V.; Razumova, Y.V.; Zarubaev, V.V.; Tretiak, T.S.; Pokrovsky, A.G.; Kiselev, O.I.; Salakhutdinov, N.F. Discovery of a new class of antiviral compounds: Camphor imine derivatives. Eur. J. Med. Chem. 2015, 105, 263–273. [Google Scholar] [CrossRef]
  174. Sokolova, A.S.; Yarovaya, О.I.; Baev, D.S.; Shernyukov, А.V.; Shtro, A.A.; Zarubaev, V.V.; Salakhutdinov, N.F. Aliphatic and alicyclic camphor imines as effective inhibitors of influenza virus H1N1. Eur. J. Med. Chem. 2017, 127, 661–670. [Google Scholar] [CrossRef] [PubMed]
  175. Sokolova, A.S.; Yarovaya, O.C.; Korchagina, D.V.; Zarubaev, V.V.; Tretiak, T.S.; Anfimov, P.M.; Kiselev, O.I.; Salakhutdinov, N.F. Camphor-based symmetric diimines as inhibitors of influenza virus reproduction. Bioorg. Med. Chem. 2014, 22, 2141–2148. [Google Scholar] [CrossRef] [PubMed]
  176. Xu, H.; Wu, T.; Huang, L. Therapeutic and delivery strategies of phytoconstituents for renal fibrosis. Adv. Drug Deliv. Rev. 2021, 177, 113911. [Google Scholar] [CrossRef]
  177. Rodrigues, T.; Reker, D.; Schneider, P.; Schneider, G. Counting on natural products for drug design. Nat. Chem. 2016, 8, 531–541. [Google Scholar] [CrossRef]
  178. Giménez, B.G.; Santos, M.S.; Ferrarini, M.; Fernandes, J.P. Evaluation of blockbuster drugs under the rule-of-five. Pharmazie 2010, 65, 148–152. [Google Scholar]
  179. Lipinski, C.A. Rule of five in 2015 and beyond: Target and ligand structural limitations, ligand chemistry structure and drug discovery project decisions. Adv. Drug Deliv. Rev. 2016, 101, 34–41. [Google Scholar] [CrossRef] [PubMed]
  180. Wang, S.; Su, R.; Nie, S.; Sun, M.; Zhang, J.; Wu, D.; Moustaid-Moussa, N. Application of nanotechnology in improving bioavailability and bioactivity of diet-derived phytochemicals. J. Nutr. Biochem. 2014, 25, 363–376. [Google Scholar] [CrossRef]
  181. Armendáriz-Barragán, B.; Zafar, N.; Badri, W.; Galindo-Rodríguez, S.A.; Kabbaj, D.; Fessi, H.; Elaissari, A. Plant extracts: From encapsulation to application. Expert Opin. Drug Deliv. 2016, 13, 1165–1175. [Google Scholar] [CrossRef]
  182. Nyhan, B.; Reifler, J. Does correcting myths about the flu vaccine work? An experimental evaluation of the effects of corrective information. Vaccine 2015, 33, 459–464. [Google Scholar] [CrossRef] [PubMed]
  183. Elbahesh, H.; Gerlach, T.; Saletti, G.; Rimmelzwaan, G.F. Response Modifiers: Tweaking the Immune Response against Influenza A Virus. Front. Immunol. 2019, 10, 809. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The structure of influenza A virus. PB1, polymerase basic protein 1; PB2, polymerase basic protein 2; PA, polymerase acidic protein; HA, hemagglutinin; NA, neuraminidase; NP, nucleoprotein; PA, polymerase acidic protein; NS, non-structural protein; M, matrix protein.
Figure 1. The structure of influenza A virus. PB1, polymerase basic protein 1; PB2, polymerase basic protein 2; PA, polymerase acidic protein; HA, hemagglutinin; NA, neuraminidase; NP, nucleoprotein; PA, polymerase acidic protein; NS, non-structural protein; M, matrix protein.
Molecules 29 02371 g001
Figure 2. The life cycle of the influenza A virus and the targets of natural compounds against influenza A viruses (H1N1). Blue rounded box, anti-H1N1 compound; red arrow, target. (1) Influenza virus hemagglutinin (HA) binds to sialic acid-presenting receptors on the surface of host cell membranes. Virus particles enter the host cells to form endosomes through receptor-mediated cellular endocytosis. (2) Endosome acidification promotes conformational changes of HA, resulting in the uncoating of the virus and release of the vRNP into the cytosol of the host cell, with further transportation to the nucleus. (3) vRNPs enter the nucleus to initiate the viral mRNA. HA, NA, and M2 are processed at the ER apparatus and Golgi before transport to the cell surface. Influenza virus polymerase can synthesize viral mRNA and vRNA. The vRNA is first converted into plus-strand cRNA; then the new vRNA is synthesized using cRNA as a template. (4) Viral proteins and vRNA are transported to the cell surface to assemble progeny viruses and initiate the virus budding process. The progeny virus is then released from the surface of the infected cells and seeks new host cells to infect. SA, sialic acid; HA, hemagglutinin; NA, neuraminidase; M2, ion channel protein; NP, nucleoprotein; PA, polymerase acidic protein; PB1, polymerase basic protein 1; PB2, influenza polymerase subunit protein; M1, matrix protein1; NS1, non-structural protein 1; NEP, nuclear export protein; vRNP, viral RNA ribonucleoprotein; NS2, non-structural protein 2; ER, endoplasmic reticulum.
Figure 2. The life cycle of the influenza A virus and the targets of natural compounds against influenza A viruses (H1N1). Blue rounded box, anti-H1N1 compound; red arrow, target. (1) Influenza virus hemagglutinin (HA) binds to sialic acid-presenting receptors on the surface of host cell membranes. Virus particles enter the host cells to form endosomes through receptor-mediated cellular endocytosis. (2) Endosome acidification promotes conformational changes of HA, resulting in the uncoating of the virus and release of the vRNP into the cytosol of the host cell, with further transportation to the nucleus. (3) vRNPs enter the nucleus to initiate the viral mRNA. HA, NA, and M2 are processed at the ER apparatus and Golgi before transport to the cell surface. Influenza virus polymerase can synthesize viral mRNA and vRNA. The vRNA is first converted into plus-strand cRNA; then the new vRNA is synthesized using cRNA as a template. (4) Viral proteins and vRNA are transported to the cell surface to assemble progeny viruses and initiate the virus budding process. The progeny virus is then released from the surface of the infected cells and seeks new host cells to infect. SA, sialic acid; HA, hemagglutinin; NA, neuraminidase; M2, ion channel protein; NP, nucleoprotein; PA, polymerase acidic protein; PB1, polymerase basic protein 1; PB2, influenza polymerase subunit protein; M1, matrix protein1; NS1, non-structural protein 1; NEP, nuclear export protein; vRNP, viral RNA ribonucleoprotein; NS2, non-structural protein 2; ER, endoplasmic reticulum.
Molecules 29 02371 g002
Figure 3. Marine fungi metabolites with activity against H1N1.
Figure 3. Marine fungi metabolites with activity against H1N1.
Molecules 29 02371 g003
Figure 4. Flavonoids with activity against H1N1.
Figure 4. Flavonoids with activity against H1N1.
Molecules 29 02371 g004
Figure 5. Alkaloids derivatives with activity against H1N1.
Figure 5. Alkaloids derivatives with activity against H1N1.
Molecules 29 02371 g005
Figure 6. Summary of the anti-H1N1structure–activity relationships of flavonoids. Orange ellipse, groups with increased anti-H1N1 activity; orange dashed line, promoting effect; green ellipse, group with reduced anti-H1N1 activity; green solid line, inhibiting effect.
Figure 6. Summary of the anti-H1N1structure–activity relationships of flavonoids. Orange ellipse, groups with increased anti-H1N1 activity; orange dashed line, promoting effect; green ellipse, group with reduced anti-H1N1 activity; green solid line, inhibiting effect.
Molecules 29 02371 g006
Figure 7. Terpenoid derivatives with activity against H1N1.
Figure 7. Terpenoid derivatives with activity against H1N1.
Molecules 29 02371 g007
Figure 8. Phenol compounds with activity against H1N1.
Figure 8. Phenol compounds with activity against H1N1.
Molecules 29 02371 g008
Figure 9. Phenol and miscellaneous compounds with activity against H1N1.
Figure 9. Phenol and miscellaneous compounds with activity against H1N1.
Molecules 29 02371 g009
Figure 10. Derivatives of natural products with activity against H1N1.
Figure 10. Derivatives of natural products with activity against H1N1.
Molecules 29 02371 g010
Table 1. Anti- H1N1 phytochemicals from natural resources.
Table 1. Anti- H1N1 phytochemicals from natural resources.
NoNatural ResourceActive CompoundStrainActivitiesRef.
1Paratetilla sp. sponge-derived fungus, Penicillium sclerotiorum OUCMDZ-3839Sclerotiorin EH1N1 A/PR/8/34IC50 = 79 μM[51]
2(+) SclerotiorinH1N1 A/PR/8/34IC50 = 129 μM
3TL-1-monoactateH1N1 A/PR/8/34IC50 = 115 μM
4OchrephiloneH1N1 A/PR/8/34IC50 = 151 μM
58-acetyldechloroisochromophilone IIIH1N1 A/PR/8/34IC50 = 91 μM
6ScleratioramineH1N1 A/PR/8/34IC50 = 134 μM
7Isochromophilone IXH1N1IC50 = 157 μM
8Marine-derived fungus Aspergillus terreus SCSGAF0162Asperterrestide AA/WSN/33 (H1N1)IC50 = 15 μM[52]
9Marine-derived fungus Aspergillus terreus OUCMDZ-1925Rubrolides SH1N1 A/PR/8/34IC50 = 87 µM[53]
10Aspergillus sp. SCSIO XWS02F40Asteltoxin E H1N1IC50 = 4 µM[54]
11Streptomyces sp. OUCMDZ-3434Wailupemycin JH1N147.8% inhibition at 50 μg/mL [55]
12R-wailupemycin KH1N142.5% inhibition at 50 μg/mL
135-deoxyenterocin H1N160.6% inhibition at 50 μg/mL
14M. senhouseiPyropheophoride aH1N1 A/PR/8/34IC50 = 0.17 µg/mL[56]
15The South China Sea soft coral Sarcophyton sp.(24R)-methylcholest-7-en-3β,5α,6β-triolH1N1 A/PR/8/34IC50 = 19.6 µg/mL[57]
16The South China Sea soft coral Sarcophyton sp.(24S)-ergost-3β,5α,6β, 11α-tetraolH1N1 A/PR/8/34IC50 = 36.7 µg/mL[57]
17Cleistocalyx operculatus leaves2′,4′dihydroxy-6′-methoxy-3′,5′-dimethylchalcone H1N1 A/PR/8/34IC50 = 8 μM[58]
18Myricetin-3′,5′-dimethylether 3-O-β-D-galactopyranoside H1N1 A/PR/8/34IC50 = 9 μM
19Pentarhizidium orientaleDemethoxymatteucinolH1N1 A/PR/8/34IC50 = 30 µM[59]
20MatteucinolH1N1 A/PR/8/34IC50 = 25 µM
21MatteucinH1N1 A/PR/8/34IC50 = 24 µM
22MethoxymatteucinH1N1 A/PR/8/34IC50 = 25 M
233′-hydroxy-5′-methoxy-6,8dimethylhuazhongilexoneH1N1 A/PR/8/34IC50 = 24 µM
24Bee pollenKaempferol-3-sophoroside H1N1IC50 = 86 µM[60]
25Kaempferol-3-neohesperidosideH1N1IC50 = 56 µM
26Kaempferol-3-sambubioside H1N1IC50 = 45 µM
27Kaempferol-3-glucosideH1N1IC50 = 36 µM
28Quercetin-3-sophoroside H1N1IC50 = 88 µM
29Bee pollen/Rhodiola rosea/Salvia plebeia R. Br.Luteolin H1N1IC50 = 11-18 µM
30Bee pollenChelianthifoline H1N1IC50 = 101 µM
31Rhodiola roseaApigenin rvH1N1IC50 = 33 µM[61]
32Kaempferol rvH1N1IC50 = 11 µM
33QuercetinrvH1N1IC50 = 2 µM
34Herbacetin rvH1N1IC50 = 9 µM
35GossypetinrvH1N1IC50 = 3 µM
36Cosmosiin rvH1N1IC50 = 47 µM
37Astragalin rvH1N1IC50 = 38 µM
38Rhodiolinin rvH1N1IC50 = 10 µM
39RhodioninrvH1N1IC50 = 32 µM
40RhodiosinrvH1N1IC50 = 57 µM
41LinocinamarinrvH1N1IC50 = 44 µM
42RutinrvH1N1IC50 = 34 µM
43Nicotiflorin rvH1N1IC50 = 32 µM
44Elderberries5,7,3′,4′-tetra-O-methylquercetin H1N1 A/PR/8/34IC50 = 0.4 µM[62]
45(±)-Dihydromyricetin H1N1 A/PR/8/34IC50 = 9 µM
46Salvia plebeia R. Br.HispidulinH1N1 A/PR/8/34IC50 = 20 µM[63]
47NepetinH1N1 A/PR/8/34IC50 = 11 µM
48Rosmarinic acid methyl esterH1N1 A/PR/8/34IC50 = 17 µM
-HoneysuckleAcids-flavonoidsH1N1EC50 = 4 µg/mL[64]
49Commelina communis L.HomonojirimycinA/PR/8/34 (H1N1)EC50 = 10 μg/mL [65]
50Jishengella endophytica 161111PerlolyrineH1N1 A/PR/8/34IC50 = 38 μg/mL[66]
511-hydroxy-β-carbolineH1N1 A/PR/8/34IC50 = 25 μg/mL
52LumichromeH1N1 A/PR/8/34IC50 = 40 μg/mL
531H-indole-3-carboxaldehydeH1N1 A/PR/8/34IC50 = 46 μg/mL
54Marine sponges Pericharax heteroraphisLeucettamine CH1N1 A/PR/8/3433% Inhibition at 50 μg/mL[67]
-Peganum harmala L.The crude extractH1N1 A/PR/8/34IC50 = 10 µg/mL[68]
-Total alkaloidH1N1 A/PR/8/34IC50 = 6 µg/mL
55Papaver somniferumPapaverineA/WSN/33 (H1N1) IC50 = 17 µM[69]
56I. indigotica leavesIsatidifoliumosidesIsatidifoliumosides/Epiisatidifoliumosides C (53/54) in the 3:2 ratioH1N1 A/PR/8/34IC50 = 65 µmol/L[70]
57Epiisatidifoliumosides C
58Sonneratia paracaseolarisParacaseolins AH1N1IC50 = 28 µg/mL[71]
-Ganoderma lingzhiHot water extract (A/California/04/2009/(H1N1))IC50 = 15 µg/mL[72]
59Trichoderma atroviride FKI-3849.Wickerol AA/PR/8/34 (H1N1)IC50 = 0.1 mg/mL[73]
60Wickerol BA/PR/8/34 (H1N1)IC50 = 5 mg/mL
61Lemnalia sp. (No. XSSC201907)Neolemnane sesquiterpene lineolemnenes FH1N1IC50 = 6 µM[74]
62Celastrus aculeatus Merr.Aculeatusane AA/GZ/GIRD07/09 (H1N1)IC50 = 23 µM[75]
63Calotropis gigantea (Asclepiadaceae)(+)-pinoresinol 4-O-[6″-O-vanilloyl]- β-D-glucopyranoside A/PR/8/34 (H1N1) IC50 = 25 µM[76]
64Tea polyphenols(-)-epigallocatechin (EGC)A/PR/8/34 (H1N1) IC50 = 31 µg/mL[77]
65(-)-epigallocatechingallate (EGCG)A/PR/8/34 (H1N1) IC50 = 56 µg/mL
66Procyanidin B-2A/PR/8/34 (H1N1) IC50 = 51 µg/mL
67Procyanidin B-23,3′-di-O-gallateA/PR/8/34 (H1N1) IC50 = 35 µg/mL
68TheaflavinA/PR/8/34 (H1N1)IC50 = 16 µg/mL
-Gyrodinium impudiumThe sulfated exopolysaccharide, p-KG03A/PR/8/34 (H1N1) EC50 = 0.5 µg/mL[78]
-wild-type (WT) H1N1EC50 = 0.2 µg/mL
-A. nodosumCold water extractA/PR/8/34 (H1N1) IC50 = 112 µg/mL[79]
-Hot water extractA/PR/8/34 (H1N1) IC50 = 101 µg/mL
-2% Na2CO3 extractA/PR/8/34 (H1N1) IC50 = 152 µg/mL
-0.5 mol−1NaOH extractA/PR/8/34 (H1N1) IC50 = 191 µg/mL
-F. vesiculosusCold water extractA/PR/8/34 (H1N1) IC50 = 75 µg/mL
-Hot water extractA/PR/8/34 (H1N1) IC50 = 181 µg/mL
-2% Na2CO3 extractA/PR/8/34 (H1N1) IC50 = 177 µg/mL
-0.5 mol−1 NaOH extractA/PR/8/34 (H1N1) IC50 = 125 µg/mL
-Red algae Eucheuma denticulatum85% ethanol extractH1N1IC50 = 277 µg/mL[80]
-Water extractH1N1IC50 = 366 µg/mL
-4% NaOH extractH1N1IC50 = 436 µg/mL
69Roots of Ilex asprellaAsprellcoside A A/PR/8/34 (H1N1) EC50 = 4 mM[81]
703,4,5-trimethoxyphe β-D-5-O-caffeoyl-Apiofuranosyl-(16) -β-D-GlucopyranosideA/PR/8/34 (H1N1) EC50 = 2 mM
71Angelica dahuricaIsoimperatorinA/PR/8/34 (H1N1) EC50 = 8 µM[82]
72OxypeucedaninA/PR/8/34 (H1N1) EC50 = 6 µM
73Oxypeucedanin hydrate A/PR/8/34 (H1N1) EC50 = 11 µM
74Imperatorin A/PR/8/34 (H1N1) EC50 = 11 µM
75Gentiopicroside derivatives2′,3′,6′-Tri-O-benzoyl-4′- O-methylsulfonyl gentiopicrosideA/WSN/33 (H1N1) IC50 = 39.5 µM[83]
76Gentiopicroside derivatives4′-fluoro-4′-deoxy gentiopicrosideA/WSN/33 (H1N1) IC50 = 45.2 µM[83]
77Gentiopicroside derivatives2′,3′,6′-Tri-O-benzoyl-4′,5′-olefin gentiopicrosideA/WSN/33 (H1N1) IC50 = 44.0 µM[83]
78Triptolide derivatives4-(((5bS, 6aS, 7aR, 8R, 8aS, 9aS, 9bS, 10aS, 10bS)-8a-isopropyl-10bmethyl-3-oxo 1, 2,3, 5,5b, 6,6a, 8,8a, 9a, 9b, 10b dodecahydrotris(oxireno) [2′, 3′:4b, 5;2″, 3″:6, 7;2‴, 3‴:8a, 9] phenanthro[1, 2-c] furan-8-yl)oxy)-2, 2-dimethyl-4 oxobutanoic acidA/WSN/33 (H1N1)EC50 = 3.24 µM[84]
79(-)-Borneol derivativesN,N,N-Trimethyl-2-oxo-2-((1S,2R,4S)-1,7,7-trimethylbicyclo [2.2.1]heptan-2-yloxy)ethanaminium iodideA/PR/8/34 (H1N1)IC50= 2.4 μM[85]
80Andrographolide derivatives14-alpha-lipoyl andrographolideA/PR/8/34 (H1N1)EC50 = 7.2 µM[86]
81Resveratrol derivatives R42A/PR/8/34 (H1N1)IC50 = 3.56 µM[87]
82Pterodontic acid derivativesPterodontic acid derivatives compound 15H1N1IC50 = 9.92 µM[88]
83Gastrodin derivativesMethyl 4-fluoro-3-((2S,3R,4S,5R,6R)-3,4,5-triacetoxy-6-(acetoxymethyl)-tetrahydro-2H-pyran-2-yloxy) benzoateA/FM/1/47(H1N1)IC50 = 34.4 μM[89]
84Mannich bases of abietic acid derivativesMannich bases of abietic acid derivatives compound 12A/PR/8/34 (H1N1)IC50 = 5.0 μM[90]
85Terpenophenols and some of their N- and O-derivatives2-(1,7,7-Trimethylbicyclo[2.2.1]hept-exo-2-yl)cyclohexa-2,5-dien-1,4-dioneA/PR/8/34 (H1N1)IC50 = 0.5 µM[91]
86Ferulic acid derivatives(E)-3-(4-Hydroxy-3-methoxyphenyl)-1-(4-methylpiperazin-1-yl)-prop-2-en-1-oneH1N1IC50 = 12.77 ± 0.47 μg/mL[92]
Table 2. Pharmacological and molecular parameters of Anti-H1N1 phytochemicals from natural resource.
Table 2. Pharmacological and molecular parameters of Anti-H1N1 phytochemicals from natural resource.
NoNameMWAlog PHdonHaccOB (%)Cao-2BBBDLRBN
24Kaempferol-3-sophoroside610.57−2.0710165.30−2.42−3.270.717
27Kaempferol-3-glucoside448.41−0.327112.77−1.36−1.990.744
28Quercetin-3-sophoroside626.57−2.3311173.37−3.390.670.307
29Luteolin286.252.074636.160.19−0.840.251
32Kaempferol286.251.774641.880.26−0.550.241
33Quercetin302.251.505746.430.05−0.770.281
34Herbacetin302.251.505736.070.12−0.650.271
35Gossypetin318.251.246835.00−0.09−1.020.311
36Cosmosiin432.410.436109.68−1.08−2.260.744
37Astragalin448.41−0.3271114.03−1.24−1.970.744
42Rutin610.57−1.4510163.20−1.93−2.750.686
43Nicotiflorin594.57−1.189153.64−1.77−2.550.736
46Hispidulin300.282.323630.970.48−0.490.272
47Nepetin316.282.054726.750.37−0.780.312
50Perlolyrine264.303.202365.950.880.150.272
531H-indole-3-carboxaldehyde145.171.881119.821.251.170.041
55Papaverine339.423.500564.041.220.570.386
65(-)-epigallocatechin306.291.656724.18−0.22−0.820.271
66Procyanidin B-2578.563.3610123.01−1.14−2.020.663
67ProcyanidinB-23,3′-di-O-gallate902.88−0.1316223.01−3.60−4.930.179
71Isoimperatorin270.303.650445.460.970.660.233
72Oxypeucedanin286.302.000524.900.850.110.303
74Imperatorin270.303.650434.551.130.920.223
MW, molecular weight; Alog P, lipophilicity; Hdon, hydrogen bond donors; Hacc, hydrogen bond acceptor; OB (%), oral bioavailability; Cao-2, Caco-2 permeability; BBB, blood–brain barrier; DL, drug-likeness; RBN, the number of bonds that allow free rotation around themselves.
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

Li, R.; Han, Q.; Li, X.; Liu, X.; Jiao, W. Natural Product-Derived Phytochemicals for Influenza A Virus (H1N1) Prevention and Treatment. Molecules 2024, 29, 2371. https://doi.org/10.3390/molecules29102371

AMA Style

Li R, Han Q, Li X, Liu X, Jiao W. Natural Product-Derived Phytochemicals for Influenza A Virus (H1N1) Prevention and Treatment. Molecules. 2024; 29(10):2371. https://doi.org/10.3390/molecules29102371

Chicago/Turabian Style

Li, Ruichen, Qianru Han, Xiaokun Li, Xinguang Liu, and Weijie Jiao. 2024. "Natural Product-Derived Phytochemicals for Influenza A Virus (H1N1) Prevention and Treatment" Molecules 29, no. 10: 2371. https://doi.org/10.3390/molecules29102371

Article Metrics

Back to TopTop