Next Article in Journal
Functional Diversity and Engineering of the Adenylation Domains in Nonribosomal Peptide Synthetases
Previous Article in Journal
High Degree of Polymerization of Chitin Oligosaccharides Produced from Shrimp Shell Waste by Enrichment Microbiota Using Two-Stage Temperature-Controlled Technique of Inducing Enzyme Production and Metagenomic Analysis of Microbiota Succession
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Terpenoids from Marine Sources: A Promising Avenue for New Antimicrobial Drugs

1
School of Pharmacy, Yantai University, Yantai 264005, China
2
Yantai Key Laboratory of Characteristic Agricultural Bioresource Conservation & Germplasm Innovative Utilization, School of life sciences, Yantai University, Yantai 264005, China
3
College of Pharmacy, University of Utah, Salt Lake City, UT 84108, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Mar. Drugs 2024, 22(8), 347; https://doi.org/10.3390/md22080347
Submission received: 29 June 2024 / Revised: 24 July 2024 / Accepted: 25 July 2024 / Published: 28 July 2024
(This article belongs to the Special Issue Marine Natural Products as Potential Drug Leads)

Abstract

:
Currently, there is an urgent need for new antibacterial and antifungal agents to combat the growing challenge of antibiotic resistance. As the largest ecosystem on Earth, the marine ecosystem includes a vast array of microorganisms (primarily bacteria and fungi), plants, invertebrates, and vertebrates, making it a rich source of various antimicrobial compounds. Notably, terpenoids, known for their complex structures and diverse bioactivities, are a significant and promising group of compounds in the battle against bacterial and fungal infections. In the past five years, numerous antimicrobial terpenoids have been identified from marine organisms such as bacteria, fungi, algae, corals, sea cucumbers, and sponges. This review article provides a detailed overview of 141 terpenoids with antibacterial and/or antifungal properties derived from marine organisms between 2019 and 2024. Terpenoids, a diverse group of natural organic compounds derived from isoprene units, are systematically categorized based on their carbon skeleton structures. Comprehensive information is provided about their names, structures, biological sources, and the extent of their antibacterial and/or antifungal effectiveness. This review aims to facilitate the rapid identification and development of prospective antimicrobials in the pharmaceutical sector.

1. Introduction

Antibiotics represent one of the most effective drugs for developing infections in humans and animals. Their extensive use is due to their broad spectrum of activity, which includes inhibiting the biosynthesis of the bacterial cell wall, disrupting the integrity of the cell membrane, suppressing the synthesis of nucleic acids and proteins, and interfering with metabolic processes [1].
Unfortunately, the advent of antibiotics has been accompanied by the escalating problem of antimicrobial drug resistance. In addition to inherent resistance, bacteria can acquire resistance to specific antimicrobial agents by transferring genetic material that confers resistance. To date, some of the most commonly observed strategies of bacterial resistance include modification of antibiotic target sites, increased cell wall permeability to antibiotics, active expulsion of antibiotics from the cell (known as efflux systems), and enzymatic inactivation [2]. Antibiotic resistance is a significant global public health concern, with an estimated 1.27 million deaths worldwide attributed to it [3]. It is projected that by 2050, the global death toll due to antibiotic resistance could reach 10 million per year, up from the current estimate of 700,000 deaths per year [4]. The widespread use of antibiotics in clinical and community settings and livestock and crop production is considered one of the main drivers of antimicrobial resistance [5,6,7]. The widespread use of antibiotics in clinical and community settings and livestock and crop production is considered one of the main drivers of antimicrobial resistance. Thus, it is necessary to improve the appropriate use of antibiotics and reduce unnecessary use. The World Health Organization has identified the ESKAPE pathogens—vancomycin-resistant Enterococcus faecalis (VRE), methicillin-resistant Staphylococcus aureus (MRSA), Klebsiella pneumoniae (K. pneumoniae), Acinetobacter baumannii (A. baumannii), Pseudomonas aeruginosa (P. aeruginosa), and vancomycin-resistant Enterobacter—as those with increasing multidrug resistance. Additionally, in epidemiology, Escherichia coli (E. coli), penicillin-resistant Streptococcus pneumoniae (PRSP), and extensively drug-resistant (XDR) Mycobacterium tuberculosis (M. tuberculosis) are well-known and significant multidrug-resistant bacteria [8,9,10]. The urgent need for new types of antibiotics to combat these pathogens highlights the importance of discovering and developing new antibacterial products for human, animal, agricultural, food, and environmental health [11].
It is well recognized that the marine ecosystem, the largest and most significant ecosystem on Earth, boasts immense biodiversity, including organisms ranging from nanoscale microorganisms to whales [12]. The marine environment offers a higher likelihood of discovering new antibacterial drug leads than terrestrial environments, making it a promising source for developing new antibiotics. Various marine organisms, such as bacteria, fungi, algae, corals, sea cucumbers, and sponges, have been explored for isolating antibacterial and antifungal bioactive compounds [13].
Terpenoids, significant both as natural products from terrestrial microorganisms and as metabolites in the ocean, are key candidates in the fight against microbial infections [14]. This review article provides a comprehensive account of 143 terpenoids identified between 2019 and 2024, with antibacterial and/or antifungal activities, sourced from a diverse array of marine organisms, including bacteria, fungi, algae, corals, sea cucumbers, and sponges. It details the names, structures, biological origins, and the compounds’ effectiveness against drug-resistant pathogens (most entries include the minimum inhibitory concentration (MIC) values against test bacterial and/or fungal strains). Additionally, certain compounds’ structure–activity relationships (SARs) were analyzed based on the magnitude of antimicrobial activity. The structures of these compounds are depicted in Figures 2–5, while the remaining information is presented in Tables 1–8. Figure 1 illustrates the analysis of statistical data. This review aims to facilitate and accelerate the identification and development of potentially innovative antimicrobial compounds to advance new pharmaceutical options.

2. Chemical Constitution

Terpenes are a diverse group of natural products synthesized from repeating units of isoprene. This class includes various compounds, such as monoterpenes (C10), sesquiterpenes (C15), diterpenes (C20), and triterpenoids (C30). Frequent skeletal rearrangements within these structures often deviate from the typical head-to-tail arrangement of isoprene units, introducing a significant degree of diversity to the terpenoid framework [13,15]. This review encompasses 141 antimicrobial terpenoids, including 48 sesquiterpenoids, 39 diterpenoids, 20 triterpenoids, and 34 meroterpenoids.

2.1. Sesquiterpenoids (148)

Sesquiterpenes are characterized by their basic carbon skeleton, which includes 15 carbon atoms arranged in three isoprene units. This section introduces 48 sesquiterpenoid compounds, including one linear sesquiterpenoid, three nardosinane-type sesquiterpenes, three neolemnane sesquiterpenes, one aristolane-type sesquiterpenoid, two drimane sesquiterpenes, and four each of carotane-style, merosesquiterpenoids, and three illudalane-style sesquiterpenoid. Additionally, there are eleven bisabolane-type sesquiterpenoids and eight sesquiterpene-derived compounds, such as sesquiterpene hydroquinones and sesquiterpene glycosides, and seven unclassified sesquiterpenoids.
The chemical structures of sesquiterpenoids 148 are depicted in Figure 2, while the remaining information, including names and marine sources, is presented in Table 1.

2.1.1. Linear Sesquiterpenoid

Chermesiterpenoid D (1), a new linear sesquiterpenoid, was isolated and identified from the fungus Penicillium rubens AS-130, which originates from the Magellan Seamount. The elucidation of its structure was achieved through nuclear magnetic resonance (NMR) and mass spectroscopic (MS) data analysis. The determination of its absolute configuration was accomplished by employing a synergistic approach of quantum mechanics (QM)-NMR and time-dependent density functional theory (TDDFT) computational methods [16].

2.1.2. Nardosinane-Type Sesquiterpenes

Three undescribed nardosinane-type sesquiterpenes, including 12-O-acetyl-nardosinan-6-en-1-one (2), 6β-acetyl-1(10)-α-13-nornardosin-7-one (3), and 6α-acetyl-1(10)-α-13-nornardosin-7-one (4), were isolated from the alcyonacean soft coral Rhytisma fulvum fulvum. Their chemical structures were elucidated based on 1D, 2D NMR, and MS spectral data [17].

2.1.3. Neolemnane Sesquiterpenes and Aristolane-Type Sesquiterpenoids

Three novel neolemnane sesquiterpenes, designated as Lineolemnenes E, F, and G (57), along with a new aristolane-type sesquiterpenoid, 2-acetoxy-aristolane (8), have been characterized. Their structural elucidation was achieved through comprehensive spectroscopic analyses coupled with the comparison of experimental and calculated electronic circular dichroism (ECD) data [18].

2.1.4. Drimane Sesquiterpenes

A marine-derived Penicillium sp. ZZ1283 yielded a novel drimane sesquiterpene lactone, purpuride D (9). The structure of purpuride D was elucidated through a multifaceted approach that included high-resolution electrospray ionization mass spectrometry (HRESIMS), NMR spectroscopic analyses, single-crystal X-ray diffraction, and ECD calculations [19]. Additionally, another drimane sesquiterpenoid, astellolide Q (10), was isolated from the culture of the marine fungus Penicillium sp. N-5. Its structure was also determined by a combination of spectroscopic methods, including MS, NMR, ECD, and X-ray diffraction [20].

2.1.5. Carotane-Style Sesquiterpenoids

Byssocarotins A−D (1114), four new carotane-style sesquiterpenoids, were obtained from a macroalga-associated strain (RR-dl-2-13) of the fungus Byssochlamys spectabilis. These isolates were identified through an integrated application of various spectroscopic techniques, encompassing MS, NMR, ECD, and X-ray diffraction [21].

2.1.6. Illudalane Sesquiterpenoids

From an Antarctic deep-water octocoral, four bioactive compounds were successfully isolated, comprising three illudalane sesquiterpenoids: Alcyopterosin T (15), Alcyopterosin U (16), and Alcyopterosin V (17). The structural characterization of these novel entities was accomplished by employing a thorough suite of 1D and 2D NMR analytical techniques [22].

2.1.7. Merosesquiterpenoids

Four merosesquiterpenoids, including a new sesquiterpenoid aminoquinone known as nakijiquinone V (18), were identified from the Indonesian marine sponge Dactylospongia elegans. Additionally, illimaquinone (19), smenospongine (20), and dyctioceratine C (21) were also found. The structure of compound 18 was elucidated by 1D and 2D NMR as well as by liquid chromatography HRESIMS data analysis [23].

2.1.8. Bisabolane-Type Phenolic Sesquiterpenoids

From the South China Sea marine sponge Plakortis simplex, a collection of six novel bisabolane-type phenolic sesquiterpenoids was successfully isolated, including plakordiols A to D (2225), (7R, 10R)-hydroxycurcudiol (26), and (7R, 10S)-hydroxycurcudiol (27). These compounds were extracted from the methanolic extract of Plakortis simplex through a series of reversed-phase chromatography and RP-HPLC separation techniques. The elucidation of their structures was facilitated by MS and NMR spectroscopy. Furthermore, compounds 2227’s stereochemical configurations were determined by combining coupling constant analysis, NOESY correlations, and applying the modified Mosher’s method [24].
Bisabolene derivatives, designated as compounds 2831, and a bisabolene dimer (32), were successfully isolated and characterized from Aspergillus versicolor AS-212, an Endozoic Fungus associated with Deep-Sea Coral of Magellan Seamounts. The chemical structures were ascertained through a comprehensive analysis of spectroscopic data, X-ray crystallography, specific rotation, ECD calculations, and spectral comparison [25].

2.1.9. Sesquiterpene-Derived Compounds

Four new sesquiterpene hydroquinones, named xishaeleganins A–D (compounds 3336), have been successfully isolated from the Xisha marine sponge Dactylospongia elegans (family Thorectida). Their structures were determined through a thorough examination using spectroscopic methods, ECD computations, and corroboration with spectral data documented in existing literature [26].
Three novel 2-guanidinoethanesulfonyl sesquiterpene analogs of (−)-agelasidine A, designated agelasidines G–I numbered compounds 3739, were isolated from a marine sponge Agelas nakamurai, which was collected in Orchid Island. The absolute configurations for compounds 3739 were ascertained by applying computational NMR data, the statistical DP4+ protocol, and correlating their experimental optical rotations with values predicted by B3LYP calculations [27].
Malfilanol C (40), a novel sesquiterpenoid, was first identified from the Aspergillus genus through its successful isolation from the rice solid-state fermentation products of the deep-sea-derived fungus Aspergillus puniceus A2. Its structure was elucidated based on comprehensive spectroscopic analysis, including HRESIMS and NMR, and comparing experimental and calculated ECD spectra to determine the absolute configuration [28].
A sesquiterpene glycoside, trichoacorside A (41), was isolated and identified from the culture extract of Trichoderma longibrachiatum EN-586, an endophytic fungus, obtained from the marine red alga Laurencia obtuse. The structures were deciphered from NMR and MS data, with absolute configurations confirmed through X-ray crystallography, derivatization, and DP4+ analysis [29].

2.1.10. Others Sesquiterpenoids

From the cultured mangrove-derived fungus Penicillium sp. HDN13-494, five novel sesquiterpenoids have been isolated: citreobenzofurans D to F (4244) and phomenones A to B (4546). Their structures were identified through comprehensive spectroscopic analysis, HRESIMS, and ECD calculations. Additionally, the absolute structure of compound 43 was confirmed by single-crystal X-ray diffraction [30].
Two novel sesquiterpenoids, identified as O8-ophiocomane (47) and O7-ophiocomane (48), have been successfully isolated from the brittle star Ophiocoma dentata, locally sourced from the Red Sea coast of Egypt. The structures were determined using 1D and 2D NMR, FT-IR, and MS [31].

2.2. Diterpenoids (4987)

Diterpenoids are a class of naturally occurring secondary metabolites with complex and diverse structures and biological activities, composed of 20 carbon atoms with extensive skeletal rearrangements [32]. This section presents 39 diterpenoid compounds, encompassing three 5,5,6,6,5-pentacyclic spongian diterpenes, nine indole diterpenes, one indole diterpene amino acid conjugate, three membrane diterpenes, two cyclopiane diterpenes, three diterpene alkaloids, six bicyclic diterpene glycosides, seven biflorane-type diterpenoids, and four decalin-type bicyclic diterpenes. The chemical structures of diterpenoids 4987 are depicted in Figure 3, while the remaining information, including names and marine sources, is presented in Table 2.

2.2.1. 5,5,6,6,5-Pentacyclic Spongian Diterpenes

From a Red Sea sponge specimen identified as Spongia sp., three novel 5,5,6,6,5-pentacyclic spongian diterpenes, designated Spongenolactones A–C and numbered 4951, have been isolated. Their structures were determined through comprehensive spectroscopic analysis, and the absolute configurations were ascertained by comparing experimental circular dichroism (CD) spectra with calculated ECD spectra [33].

2.2.2. Indole Diterpenes

Nine novel indole diterpenes, designated as Janthinellumine A through I and numbered 5260, were isolated from the co-culture of two marine-derived fungi, Penicillium janthinellium and Paecilomyces formosus. The chemical structures and absolute configurations were determined using extensive spectroscopic data and computational ECD and vibrational circular dichroism (VCD) methods [34].
A rare indole diterpene amino acid conjugate, noonidole A (61), was derived from the marine fungus Aspergillus sp. CMB-M0339 (identified as Aspergillus noonimiae). Its structure was determined through MS spectroscopy and X-ray crystallography [35].

2.2.3. Cembrane Diterpenes

Three novel cembrane diterpenes, Nephthecrassocolides A and B (6263), along with 6-acetoxynephthenol acetate, have been isolated from a population of the marine organism Nephthea sp. Their structures were determined using spectroscopic methods, including MS, NMR, and nuclear Overhauser effect spectroscopy (NOESY) [36].

2.2.4. Cyclopiane Diterpenes

Two novel cyclopiane diterpenes, 4-Hydroxyleptosphin C (65) and 13-epi-conidiogenone F (66), have been isolated from the marine sediment-derived fungus Penicillium antarcticum KMM 4670. Their absolute configurations were confirmed by the modified Mosher method and ECD spectrum calculations [37].

2.2.5. Diterpene Alkaloids

Three new diterpene alkaloids, (+)-8-epiagelasine T (67), (+)-10-epiagelasine B (68), and (+)-12-hydroxyagelasidine C (70), have been isolated from the sponge Agelas citrina, which was collected along the coasts of the Yucatan Peninsula. Their structures were identified through NMR spectroscopy, HRESIMS, and literature comparison [38].

2.2.6. Bicyclic Diterpene Glycosides

From the soft coral Lemnalia bournei, six new bicyclic diterpene glycosides—lemnaboursides E to G (7072) and lemnadiolboursides A to C (7375)—were meticulously isolated and characterized. Their structures were determined using spectroscopy (MS, NMR, heteronuclear single quantum coherence (HSQC), heteronuclear multiple bond correlation (HMBC), HRESIMS), ECD analysis, optical rotation, and literature data comparison [39].

2.2.7. Biflorane-Type Diterpenoids

A series of novel secondary metabolites, comprising five new biflorane-type diterpenoids designated as biofloranates E through I (7680) and two new bicyclic diterpene glycosides named lemnaboursides H through I (8182), have been successfully isolated from the soft coral Lemnalia bournei, collected from the South China Sea. Their chemical structures and stereochemistry were identified using various spectroscopic techniques and TDDFT ECD calculations and were confirmed by comparing them with reported values [40].

2.2.8. Decalin-Type Bicyclic Diterpenes

Four novel decalin-type bicyclic diterpenes, designated Biofloranates A to D (8386), along with a new aromadendrane-type diterpenoid (87), have been successfully isolated from the soft coral Lemnalia sp., collected from the South China Sea. The new compounds’ structures were determined using NMR, Mosher’s method, and ECD analysis [18].

2.3. Triterpenoids (88107)

Triterpenoids are a class of ubiquitous natural organic compounds in nature, consisting of six isoprene units, totaling 30 carbon atoms. This section introduces 20 triterpenoid compounds, including details on the names, sources, and structures of the nine isomalabaricane terpenoids and eleven fusicane-type nortriterpenoids. The chemical structures of triterpenoids 88107 are depicted in Figure 4, while the remaining information, including names and marine sources, is presented in Table 3.

2.3.1. Isomalabaricane Terpenoids

Nine novel isomalabaricane terpenoids, numbered 8896, have been successfully isolated from the sponge Rhabdastrella globostellata collected from Ximao Island. The structures were determined using spectroscopic methods, including MS, NMR, HMBC, and ECD, and by comparing them with known compounds’ data [41].

2.3.2. Fusicane-Type Nortriterpenoids

Eleven novel fusicane-type nortriterpenoids, named Implifusidic acids A–K and numbered 97107, have been successfully isolated from the marine-derived fungus Simplicillium sp. SCSIO 41513. Their structures were identified through spectroscopy, and the absolute configurations were confirmed by ECD calculations, spectral comparison, and X-ray diffraction [42].

2.4. Meroterpenoids (108141)

Meroterpenoids are hybrid secondary metabolites from mixed biosynthetic pathways, partially derived from terpenoid substrates. These compounds, produced widely by bacteria, algae, plants, and animals, exhibit remarkable chemical diversity by combining terpenoid frameworks with polyketides, alkaloids, phenols, and amino acids [43,44,45]. The chemical structures of meroterpenoids 108141 are depicted in Figure 5, while the remaining information, including names, marine sources, and results of antibacterial activity assays, is presented in Table 4.
Four novel meroterpenoids, designated as chermesins E–H (108111), were successfully isolated from Penicillium chermesinum EN-480, an endophyte derived from the marine red alga. The structures were ascertained using HRESIMS and NMR, with absolute configurations verified through NOESY, X-ray diffraction, and ECD cotton effect analysis [46].

2.4.1. Drimane-Type Meroterpenoid

The first drimane-type meroterpenoid featuring a C10 polyketide unit with an 8R-configuration, named taladrimanin A (112), has been isolated alongside three biogenetically related compounds (113115) from the marine-derived fungus Talaromyces sp. HM6-1–1 [47]. After two years, three new drimane-type meroterpenoids, designated as Taladrimanin B to D and numbered 116118, have been successfully isolated from the marine-derived fungus Talaromyces sp. M27416 [48]. The planar structure of 112 was identified using HRESIMS and NMR. Its relative configuration was deduced through quantum chemical NMR calculations of potential isomers and the DP4+ method. X-ray diffraction confirmed the relative and absolute configurations. Compound 116’s structure was elucidated by HRESIMS and NMR, with configuration confirmed by quantum chemical analysis and the DP4+ method and verified by X-ray crystallography. ECD calculations established the absolute configuration of 116, while comparative NMR and ECD analyses determined those of 117 and 118.

2.4.2. α-Pyrone Meroterpenoids

Six new α-pyrone meroterpenoids, designated as chevalones H–M and numbered 119124, have been successfully isolated from Aspergillus hiratsukae SCSIO 7S2001, a fungus derived from the gorgonian coral collected at Mischief Reef in the South China Sea. The structures and absolute configurations were determined by spectroscopy and X-ray diffraction [49].

2.4.3. Spiromeroterpenoids

The chemical exploration of the ethyl acetate (EtOAc) extract derived from the fermentation broth of the marine fungus Trametes sp. ZYX-Z-16 yielded eight meroterpenoids, numbered 125132, among which two novel spiromeroterpenoids, named asnovolin H (131) and asnovolin I (132), were identified. The structures of 131 and 132 were identified through 1D and 2D NMR, HRESIMS, and ECD spectral analysis [50].

2.4.4. Andrastin-Type Meroterpenoids

In addition to compound 10, three andrastin-type meroterpenoids, identified as Hemiacetalmeroterpenoids A to C (133135), were isolated from the marine-derived fungus Penicillium sp. N-5. Its structure was also determined by a combination of spectroscopic methods, including MS, NMR, ECD, and X-ray diffraction [20].

2.4.5. Chlorinated Meroterpenoids

From the cultivation of the marine sediment-derived bacterium strain Streptomyces sp. CNH-189, four novel chlorinated meroterpenoids, identified as merochlorins G through J (136139), have been successfully isolated. The planar structures of compounds 137140 were deduced from MS, ultraviolet-visible spectroscopy, and NMR data. Their relative configurations were inferred from nuclear Overhauser effect (NOE) data, and absolute configurations were confirmed by comparing ECD spectra with known models and DP4 calculations [51].

2.4.6. 3,5-Dimethylorsellinic Acid-Based Meroterpenoid

A chemical investigation of the extracts from the fungus Aspergillus sp. CSYZ-1 has led to the identification of aspergillactone (140), a novel 3,5-dimethylorsellinic acid-based meroterpenoid. NMR and mass spectrometry confirmed the structure and relative configuration of 140. Its absolute configuration was ascertained through TDDFT calculations and comparison with experimental ECD spectra [52].

2.4.7. Meroterpenoid-Type Alkaloid

Oxalicine C (141), a novel meroterpenoid-type alkaloid, has been isolated from the endophytic fungus Penicillium chrysogenum XNM-12, derived from marine algae. The planar structure of compound 141 was elucidated through spectroscopic analyses comprising ultraviolet–visible spectroscopy, 1D and 2D NMR, and HRESIMS. Its stereochemical configuration was determined by comparing experimental and calculated ECD spectra [53].

3. Antibacterial and/or Antifungal Activity

3.1. Sesquiterpenoids

This section provides a detailed account of the antimicrobial activities of 48 sesquiterpenoid compounds, with further details presented in Table 5.
Chermesiterpenoid D (1) demonstrated weak antibacterial activity, with MIC values for MRSA being 64 µg/mL [16]. Compounds 24 have exhibited against several Gram-positive (Bacillus cereus (B. cereus), Staphylococcus aureus (S. aureus)) and Gram-negative (E. coli, K. pneumoniae, and Pseudomonas sp.) bacteria. However, they have not shown inhibition against fungi Aspergillus niger (A. niger), Fusarium oxysporum (F. oxysporum), and C. albicans (C. albicans) [17].
Unfortunately, three new precious neolemnane sesquiterpene lineolemnenes, E, F, G (57), and a new aristolane-type sesquiterpenoid, 2-acetoxy-aristolane (8), have tested antibacterial activities against S. aureus and B. cereus with relatively high MIC values [18].
Compound 9, a drimane sesquiterpenoid, exhibited significant antibacterial activity with the MRSA, E. coli, and C. albicans with MIC values of 4, 3, and 8 µg/mL, respectively [19]. Another drimane sesquiterpenoid astellolide Q (10) showed remarkable antifungal activities against Penicillium italicum (P. italicum) and C. gloeosporioides with MIC values both being of 25 µg/mL [20].
Byssocarotins A–D (1114) displayed antagonism against the marine-derived bacteria Vibrio parahaemolyticus (V. parahaemolyticus) and V. harveyi with MIC values ranging from 13 to 50 µg/mL [21]. Alcyopterosin T (15), Alcyopterosin U (16), and Alcyopterosin V (17) were inactive against the ESKAPE panel of bacterial pathogens. However, compound 17 demonstrated significant efficacy against Clostridium difficile (C. difficile), an intestinal bacterium that is notoriously challenging to treat [22].
Compounds 1821 exhibited moderate to low antibacterial activity against Bacillus megaterium (B. megaterium) DSM32, with MIC values of 32 µg/mL for each. Furthermore, compounds 19 and 20 inhibited Micrococcus luteus (M. luteus) ATCC 4698 with MIC values of 32 µg/mL [23].
Plakordiols A to D (2225), (7R, 10R)-hydroxycurcudiol (26), and (7R, 10S)-hydroxycurcudiol (27). None of these compounds, ranging from 22 to 27, exhibited inhibitory effects on a panel of five bacterial strains, including S. aureus ATCC 25923, MRSA ATCC 43300, A. baumannii ATCC 19606, P. aeruginosa (clinical), and VRE CD27. However, compounds 27 and 28 demonstrated weak antibacterial activity against A. baumannii ATCC 19606 in disc diffusion tests, producing inhibition zone diameters of 5 mm each. Despite this, these compounds failed to exhibit significant activity against A. baumannii ATCC 19606 in MIC bioassays, with MIC values exceeding 64 µg/mL [24].
Compounds 2832 primarily displayed antimicrobial properties against V. harveyi and V. parahaemolyticus, with MIC values varying between 15.0 and 121.2 µg/mL. Notably, antimicrobial assays revealed that compound 29 exhibited superior efficacy to compounds 30 and 31 against the Vibrio species and C. gloeosporioides. These findings suggest hydroxylation at the C-10 or C-11 position may attenuate the antimicrobial activity against these microbial strains [25].
Compound 34 exhibited notable antibacterial potency, demonstrating significant inhibitory effects against S. aureus USA300 LAC, Streptococcus pyogenes (S. pyogenes) ATCC 12344, and Enterococcus faecium (E. faecium) Efm-HS0649. MICs for compound 34 were determined to be 1.5, 1.5, and 3.0 µg/mL for each bacterium, respectively. These MIC values are in the same range as those observed for the positive control, vancomycin, which tested a MIC of 1.0 µg/mL [26].
Remarkably, compound 37 displayed measurable antimicrobial activities when administered 25 mg per disk. This compound exhibited inhibitory effects against a panel of bacterial strains, including Bacillus subtilis (B. subtilis), E. coli, K. pneumoniae, Salmonella typhimurium (S. typhimurium), and S. aureus, with each strain demonstrating an inhibition zone of 3.0 mm in diameter [27]. Malfilanol C (40) exhibited weak antibacterial activity against S. aureus ATCC 29213 [28]. Sesquiterpene glycoside, trichoacorside A (41), demonstrated moderate activity against MRSA and Vibrio harveyi (V. harveyi), an aquatic pathogenic bacterium, with MIC values being 4 µg/mL, the tested plant–pathogenic fungi, including Alternaria brassicae, Calonectria cornigerum, Colletotrichum gloeosporioides with MIC values ranging from 8 to 64 µg/mL [29].
The majority of compounds 4246, specifically 4245, demonstrated high MIC values, indicating weak antibacterial potential. In contrast, compound 46 displayed a more pronounced effect, exhibiting moderate antibacterial activity against B. subtilis with an MIC value of 6.25 µg/mL [30]. Both compounds 47 and 48 have shown antibacterial efficacy against P. aeruginosa and Enterococcus faecalis (E. faecalis), with their antibacterial activities quantified in absolute activity units (AUs) [31].

3.2. Diterpenoids

This part elaborates on the antimicrobial effects of 39 diterpenoid substances (compounds 4987), with additional information found in Table 6.
Three novel 5,5,6,6,5-pentacyclic spongian diterpenes, numbered 4951, have been isolated. Subsequent in vitro assays were conducted to evaluate their growth inhibitory effects on S. aureus. Notably, spongenolactone A (49) demonstrated significant inhibitory activity, achieving 46%, 47%, and 93% growth inhibition at concentrations of 50, 100, and 200 µM, respectively. In contrast, spongenolactone B (50) showed comparatively lower inhibitory effects, with 24%, 42%, and 40% growth inhibition observed at the same concentration gradients [33].
Compounds 5260 have exhibited various biological activities, including anti-influenza A virus, protein tyrosine phosphatase inhibitory effects, and anti-Vibrio properties. In particular, their potential to resist Vibrio species has attracted significant interest. Notably, compounds 52 and 59 have demonstrated weak anti-Vibrio activity against Vibrio anguillarum (V. anguillarum), with MICs of 12.5 and 25.0 µg/mL, respectively [34].
Noonidole A (61) exhibited moderate antifungal activity. Regrettably, it did not demonstrate antibacterial activity against bacteria, including E. coli ATCC 11775, S. aureus ATCC 25923, and B. subtilis ATCC 6633 [35].
Compound 62 displayed significant antifungal activity, with a MIC value of 12.5 µg/mL against the hyphal growth inhibition of Lentinula thermophilum. SAR analysis revealed that the differential antifungal potency between compounds 62 and 63 may be ascribed to the presence of an epoxide ring in compound 63. Furthermore, the trisubstitution of methyl groups in the β-configuration within compound 63 could introduce steric hindrance compared to compound 62, potentially impacting its antifungal efficacy [36].
Compound 65 demonstrated a concentration-dependent inhibitory effect on the growth of S. aureus, with inhibition rates of 15.3% and 29.3% at concentrations of 12.5 µM and 100 µM, respectively. Additionally, compound 65 effectively reduced biofilm formation by S. aureus, with prevention rates of 15.9% and 34.5% at the same concentrations. Conversely, compound 66 showed a weaker effect on S. aureus growth, with inhibition of 19.1% at 100 µM and no significant impact at 12.5 µM. However, it notably inhibited biofilm formation, with prevention rates ranging from 37.9% at 12.5 µM to 52.6% at 100 µM. The half-maximal inhibitory concentration (IC50) for the inhibition of S. aureus biofilm formation by compound 66 was determined to be 76.1 µM [37]
(+)-8-epiagelasine T (67), (+)-10-epiagelasine B (68), and (+)-12-hydroxyagelasidine C (70) did not exhibit activity against the Gram-negative pathogens A. baumannii ATCC 17978, K. pneumoniae ATCC 700603, and P. aeruginosa ATCC 27823. However, these compounds did demonstrate antibacterial activity against a range of Gram-positive pathogens. This group included S. aureus ATCC 29213, S. aureus USA300LAC, Streptococcus pneumoniae (S. pneumoniae) ATCC 49619, S. pneumoniae 549 CHUAC, E. faecalis ATCC 29212, E. faecalis 256 CHUAC and E. faecium 214 CHUAC [38].
Lemnaboursides E-G (7072) and lemnadiolboursides A–C (7375) demonstrated discernible antibacterial activity, targeting both S. aureus and B. subtilis with MIC values ranging from 4 to 16 µg/mL. A comprehensive assessment integrating antimicrobial assays with detailed structural analyses has indicated a potential correlation between the steric hindrance of the glycosides and their antimicrobial efficacy, particularly with respect to the lemnaboursides [39].
The antibacterial activities of compounds 7682 were systematically evaluated against a panel of five pathogenic bacteria, including S. aureus, B. subtilis, V. harveyi, S. pneumoniae, and E. coli. Notably, all compounds within the series (7682) demonstrated antibacterial efficacy against S. aureus and B. subtilis, with MICs varying from 4 to 64 µg/mL [40]. Compounds 8387 exhibited antimicrobial properties, demonstrating antibacterial activity against S. aureus and B. cereus with MICs ranging from 4 to 16 µg/mL [18].

3.3. Triterpenoids

This section offers a comprehensive examination of the antimicrobial properties of 20 triterpenoid compounds, numbered 88 to 107, with further details available in Table 7.
Nine novel isomalabaricane terpenoids, numbered 8896, were tested against S. aureus USA300LAC and S. pyogenes ATCC12344. In this series, compounds 89 and 90 showed a substantial antibacterial effect against S. pyogenes, with MIC values recorded at 1.8 and 1.0 µg/mL, respectively [41].
Compound 105 exhibited potent antibacterial activity against S. aureus, with a remarkably low MIC value of 0.078 µg/mL. A subset of the compounds, specifically 98, 100102, 104, and 105, were selected for their targeted evaluation of antibacterial activity against S. aureus. However, compounds 97, 99, 103, 106, and 107 were excluded from the antibacterial panel due to their limited availability and the susceptibility of compound 97 to hydrolysis. Drawing on previous SAR studies, the chemical structure of fusidic acid has been established as optimal for antibacterial potency, with the C-21 carboxylic acid group being an essential moiety for activity [54]. The antibacterial findings for compounds 98, 100, 101, 102, 104, and 105 corroborated this established conclusion. Furthermore, a comparative analysis of the structures and antibacterial profiles of compounds 104, 105, and fusidic acid indicated that the oxidation of the C-11 position to a carbonyl group did not significantly impact antibacterial efficacy. In contrast, the oxidation of the hydrophobic side chain at the C-20 position was associated with decreased antibacterial activity [42].

3.4. Meroterpenoids

This section details the antimicrobial properties of meroterpenoid compounds 108141, with details in Table 8.
Chermesins E–H (108111) were subjected to a battery of assays to evaluate their antibacterial and antifungal activities against a spectrum of human and aquatic bacteria, including Aeromonas hydrophilia (A. hydrophilia), E. coli, Edwardsiella tarda (E. tarda), V. anguillarum, and V. harveyi, as well as against plant–pathogenic fungi, such as Coniothyrium diplodiella (C. diplodiella), and Fusarium graminearum (F. graminearum). Compound 108 displayed potent antibacterial activity against E. tarda and V. anguillarum, with MICs of 0.5 µg/mL, a value comparable to or exceeding the efficacy of the positive control, chloramphenicol, which showed MICs of 0.5 and 1 µg/mL, respectively. Additionally, compound 110 exhibited robust activity against the human pathogenic bacterium E. coli, with an MIC of 1 µg/mL, surpassing the activity of the positive control, chloromycetin, which had an MIC of 2 µg/mL [46].
Compound 112 exhibited selective antibacterial activity targeting S. aureus ATCC 6538P, demonstrating a noteworthy potency with a MIC of 15.2 µg/mL. This potency, however, was lower than that of the positive control, chloramphenicol, which had an MIC of 5.0 µg/mL. Compound 112’s antibacterial activity was less pronounced against V. parahaemolyticus and E. coli strains [47]. Moving on to compound 116, it also showed selective antibacterial activity against S. aureus CICC 10384, with an MIC of 12.5 µg/mL. This activity was on par with chloramphenicol, again serving as a positive control with an MIC of 5.0 µg/mL [48].
Compounds 119124 were subjected to a broth dilution assay to evaluate their antibacterial activity against a panel of bacterial strains, including M. luteus, K. pneumoniae, MRSA, and Streptococcus faecalis (S. faecalis). The compounds exhibited a range of antibacterial potencies, with MICs spanning from 6.25 to 100 µg/mL [49].
Compound 129 exhibited modest inhibitory effects against S. aureus with a MIC of 128 µg/mL. Compounds (125128 and 130132) demonstrated lackluster antibacterial properties, with MICs exceeding 128 µg/mL for both S. aureus and B. subtilis. Subsequently, the inhibitory activities of the same set of compounds were evaluated against five phytopathogenic fungi, including F. oxysporum f. sp. cubense, Fusarium spp, Peronophythora litchii (P. litchii), C. gloeosporioides, Hylocereus undatus (H. undatus) utilizing the broth microdilution technique. Regrettably, under the conditions tested, none of the compounds manifested definitive inhibitory effects against the tested fungi [50].
Hemiacetalmeroterpenoids A to C (133135) exhibit more potent antimicrobial activities against phytopathogenic fungi than their effects on bacteria. Specifically, compound 135 displayed significant antimicrobial effects against B. subtilis, P. italicum, and C. gloeosporioides with MIC of 6.25 µg/mL for all three organisms. Furthermore, compound 135 also demonstrated antibacterial activity against MRSA, with an MIC of 25 µg/mL [20].
Compound 138 exhibited robust antibacterial activity against Gram-positive strains: B. subtilis KCTC 1021, K. rhizophila KCTC 1915, and S. aureus KCTC 1927. The MICs for these strains were 1, 2, and 2 µg/mL, respectively. In contrast, compound 139 exhibited no significant antibacterial activity against the same strains, with all MICs greater than 128 µg/mL [51].
Compound 140 has demonstrated potent antimicrobial activity against various strains of Helicobacter pylori (H. pylori), including ATCC43504, G27, Hp159, and BY583, with MICs ranging from 1 to 4 µg/mL. Additionally, aspergillactone (140) exhibited significant inhibitory effects against multiple strains of S. aureus, such as ATCC25923, USA300, BKS231, and BKS233, with MICs in the range of 2 to 16 µg/mL [52].
Compound 141 demonstrated inhibitory effects against E. coli, M. luteus, and Ralstonia solanacearum (R. solanacearum) with MICs consistent at 8 µg/mL. However, against P. aeruginosa, the MIC was found to be higher at 16 µg/mL. In parallel with the antibacterial assessments, the antifungal activity of 141 was also evaluated. The compound exhibited notable antifungal activity against a panel of five plant pathogenic fungi, including Alternaria alternata (A. alternata), Botrytis cinerea (B. cinerea), F. oxysporum, Penicillium digitatum (P. digitatum), and Valsa mali (V. mali). Notably, the most potent activity was observed against V. mali, with an MIC value of 16 µg/mL [53].

4. Conclusions

This review summarizes 141 terpenoid compounds with antibacterial and/or antifungal activities discovered from marine biological resources between 2019 and 2024. These compounds are primarily derived from sponges, red algae, soft corals, fungi, bacteria, and marine sediments. They include 48 sesquiterpenes, 39 diterpenes, 20 triterpenes, and 34 meroterpenoids.
The antibacterial activity of these compounds is relatively evenly distributed and does not show a particular preference for any specific skeletal type. Among sesquiterpenes, compounds such as 9 and 34 exhibit significant antibacterial activity against S. aureus, with MICs ranging from 4 to 1.5 µg/mL. Diterpenes, including compounds 68, 71, 8387, also show promising antibacterial activity with MICs between 1 and 8 µg/mL. Notably, the fusidane-type nortriterpenoid compound 105 has the lowest MIC at 0.078 µg/mL, demonstrating strong antibacterial activity. Among meroterpenoids, compound 108 has significant antifungal activity against animal pathogens V. anguillarum and V. harveyi, with an MIC as low as 0.5 µg/mL. Compound 123 demonstrated potent antibacterial activity against M. lutea and S. faecalis, with MIC values of 12.5 µg/mL. Compound 140 showed strong antibacterial activity against H. pylori strains ATCC43504, G27, and Hp159, with MIC values as low as 1 µg/mL.
Among the 141 terpenoid compounds surveyed, those with antibacterial activity against S. aureus were most abundant, comprising 40 unique compounds. The count was followed by 21 compounds active against V. harveyi, 18 against B. subtilis, 14 against E. coli, and 10 each with efficacy against MRSA, V. anguillarum, and V. Parahaemolyticus. There were nine compounds with activity against S. pyogenes, eight compounds against B. Cereus, and terpenoid compounds with activity against C. Gloeosporioides, E. Faecium, K. Pneumoniae, A. hydrophilia, and E. Faecalis were five each. There were four compounds each against P. Italicum, P. Aeruginosa, F. Graminearum, and E. Tarda, and three against B. megaterium, M. Luteus, Exophiala sp., Haliphthoros sabahensis, H. Milfordensis, Lagenidium thermophilum, S. pneumoniae, and C. Diplodiella. The number of terpenoid compounds resistant to F.Oxysporum, P.Digitatum, K.Rhizophila, and C. albicans strains was 2, respectively. Notably, compound 17 exhibited antibacterial properties against C. difficile. Among the surveyed terpenoid compounds, compound 41 is the sole terpenoid with demonstrated antibacterial activity against the strains A. Brassicae, C. Cornigerum, Curvularia spicifera, Fusarium proliferatum, and P. Piricola. The sole compound exhibiting antibacterial activity against the A. baumannii strain is compound 27, producing inhibition zone diameters of 5.0 ± 0.6 mm. However, the compound’s MIC exceeds 64 µg/mL. Compound 62 is the only compound that exhibits resistance against the Fusarium solani strain. Compound 140 is the unique compound with antibacterial resistance against H. pylori and has demonstrated strong antibacterial activity against H. pylori strains, with an MIC of 2 µg/mL. Compound 141 stands out as the sole terpenoid to demonstrate resistance against R. solanacearum, with an MIC value of 8 µg/mL.
Furthermore, SAR analysis indicates that hydroxylation at C-10 and C-11 in compounds 29, 30, and 31 may reduce antibacterial activity; the β-configuration of the trimethyl substitution in compound 63 may decrease antifungal efficacy through steric hindrance. The structural analysis of compounds 7375 suggests that the spatial hindrance of glycosides is related to antimicrobial efficacy.
In conclusion, marine terpenoids have demonstrated considerable potential for development in the antimicrobial domain. This discovery is not only of profound significance to the pharmaceutical industry in the quest for novel antimicrobial agents but also holds promise for its potential applications in various fields, including animal nutrition and food preservation.

Author Contributions

Conceptualization, S.L.; methodology, X.L. and J.X.; software, X.L. and J.X.; validation, X.L. and J.X.; formal analysis, Y.S.; investigation, X.L.; resources, F.Z. and S.L.; data curation, Y.S.; writing—original draft preparation, X.L. and J.X.; writing—review and editing, F.Z., C.N. and S.L.; visualization, J.X. and S.L.; supervision, S.L.; project administration, S.L.; funding acquisition, S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation Project of Shandong Province, grant number ZR2023QH088, the Research Start-up Fund for Doctor in Yantai University, grant number YX20B03, and Yantai University 2024 Graduate Student Research and Innovation Fund Project, grant number GGIFYTU2438.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

SARstructure–activity relationships
MICminimum inhibitory concentration
VREvancomycin-resistant Enterococcus faecalis
MRSAmethicillin-resistant Staphylococcus aureus
K. pneumoniaeKlebsiella pneumoniae
A. baumanniiAcinetobacter baumannii
P. aeruginosaPseudomonas aeruginosa
E. coliEscherichia coli
PRSPpenicillin-resistant Streptococcus pneumoniae
XDRextensively drug-resistant
NMR nuclear magnetic resonance
MSmass spectroscopic
QMquantum mechanics
TDDFTtime-dependent density functional theory
ECDelectronic circular dichroism
HRESIMShigh-resolution electrospray ionization mass spectrometry
CDcircular dichroism
VCDvibrational circular dichroism
NOESYnuclear Overhauser effect spectroscopy
HSQCheteronuclear single quantum coherence
HMBCheteronuclear multiple bond correlation
NOEnuclear Overhauser effect
M. tuberculosisMycobacterium tuberculosis
V. harveyiVibrio harveyi
A. brassicaeAlternaria brassicae
C. cornigerumCalonectria cornigerum
C. gloeosporioidesColletotrichum gloeosporioides
B. cereusBacillus cereus
S. aureusStaphylococcus aureus
A. nigerAspergillus niger
F. oxysporumFusarium oxysporum
C. albicansCandida albicans
C. difficileClostridium difficile
P. italicumPenicillium italicum
V. parahaemolyticusVibrio parahaemolyticus
B. megateriumBacillus megaterium
M. luteusMicrococcus luteus
S. pyogenesStreptococcus pyogenes
E. faeciumEnterococcus faecium
B. subtilisBacillus subtilis
S. typhimuriumSalmonella typhimurium
E. faecalisEnterococcus faecalis
V. anguillarumVibrioanguillarum
IC50half-maximal inhibitory concentration
S. pneumoniaeStreptococcus pneumoniae
A. hydrophiliaAeromonas hydrophilia
E. tardaEdwardsiella tarda
C. diplodiellaConiothyrium diplodiella
F. graminearumFusarium graminearum
EtOAcethyl acetate
L. monocytogenesListeria monocytogenes
P. litchiiPeronophythora litchii
S. faecalisStreptococcus faecalis
H. undatusHylocereus undatus
H. pyloriHelicobacter pylori
R. solanacearumRalstonia solanacearum
A. alternataAlternaria alternata
B. cinereaBotrytis cinerea
P. digitatumPenicillium digitatum
V. maliValsa mali

References

  1. Uluseker, C.; Kaster, K.M.; Thorsen, K.; Basiry, D.; Shobana, S.; Jain, M.; Kumar, G.; Kommedal, R.; Pala-Ozkok, I. A Review on Occurrence and Spread of Antibiotic Resistance in Wastewaters and in Wastewater Treatment Plants: Mechanisms and Perspectives. Front. Microbiol. 2021, 12, 717809. [Google Scholar] [CrossRef]
  2. Li, T.; Wang, Z.; Guo, J.; de la Fuente-Nunez, C.; Wang, J.; Han, B.; Tao, H.; Liu, J.; Wang, X. Bacterial resistance to antibacterial agents: Mechanisms, control strategies, and implications for global health. Sci. Total Environ. 2023, 860, 160461. [Google Scholar] [CrossRef] [PubMed]
  3. Centers for Disease Control and Prevention. 2019 Antibiotic Resistance Threats Report. Available online: https://www.cdc.gov/antimicrobial-resistance/data-research/threats/ (accessed on 22 April 2024).
  4. Inoue, H. Strategic approach for combating antimicrobial resistance (AMR). Glob. Health Med. 2019, 1, 61–64. [Google Scholar] [CrossRef] [PubMed]
  5. Aslam, B.; Khurshid, M.; Arshad, M.I.; Muzammil, S.; Rasool, M.; Yasmeen, N.; Shah, T.; Chaudhry, T.H.; Rasool, M.H.; Shahid, A.; et al. Antibiotic Resistance: One Health One World Outlook. Front. Cell. Infect. Microbiol. 2021, 11, 771510. [Google Scholar] [CrossRef]
  6. Endale, H.; Mathewos, M.; Abdeta, D. Potential Causes of Spread of Antimicrobial Resistance and Preventive Measures in One Health Perspective—A Review. Infect Drug Resist. 2023, 16, 7515–7545. [Google Scholar] [CrossRef] [PubMed]
  7. Zhu, S.; Yang, B.; Wang, Z.; Liu, Y. Augmented dissemination of antibiotic resistance elicited by non-antibiotic factors. Ecotoxicol. Environ. Saf. 2023, 262, 115124. [Google Scholar] [CrossRef]
  8. Oliveira, D.M.P.D.; Forde, B.M.; Kidd, T.J.; Harris, P.N.A.; Schembri, M.A.; Beatson, S.A.; Paterson, D.L.; Walker, M.J. Antimicrobial Resistance in ESKAPE Pathogens. Clin. Microbiol. Rev. 2020, 33, e00181-19. [Google Scholar] [CrossRef]
  9. Venkateswaran, P.; Vasudevan, S.; David, H.; Shaktivel, A.; Shanmugam, K.; Neelakantan, P.; Solomon, A.P. Revisiting ESKAPE Pathogens: Virulence, resistance, and combating strategies focusing on quorum sensing. Front. Cell. Infect. Microbiol. 2023, 13, 1159798. [Google Scholar] [CrossRef]
  10. Alekshun, M.N.; Levy, S.B. Molecular mechanisms of antibacterial multidrug resistance. Cell 2007, 128, 1037–1050. [Google Scholar] [CrossRef]
  11. World Health Organization. Global Action Plan on Antimicrobial Resistance. Available online: https://www.who.int/publications/i/item/9789241509763 (accessed on 22 April 2024).
  12. Velmurugan, P.; Venil, C.K.; Veera Ravi, A.; Dufossé, L. Marine Bacteria Is the Cell Factory to Produce Bioactive Pigments: A Prospective Pigment Source in the Ocean. Front. Sustain. Food Syst. 2020, 4, 589655. [Google Scholar] [CrossRef]
  13. Thawabteh, A.M.; Swaileh, Z.; Ammar, M.; Jaghama, W.; Yousef, M.; Karaman, R.; Bufo, S.A.; Scrano, L. Antifungal and Antibacterial Activities of Isolated Marine Compounds. Toxins 2023, 15, 93. [Google Scholar] [CrossRef]
  14. Izzati, F.; Warsito, M.F.; Bayu, A.; Prasetyoputri, A.; Atikana, A.; Sukmarini, L.; Rahmawati, S.I.; Putra, M.Y. Chemical Diversity and Biological Activity of Secondary Metabolites Isolated from Indonesian Marine Invertebrates. Molecules 2021, 26, 1898. [Google Scholar] [CrossRef] [PubMed]
  15. Brahmkshatriya, P.P.; Brahmkshatriya, P.S. Terpenes: Chemistry, Biological Role, and Therapeutic Applications. In Natural Products: Phytochemistry, Botany and Metabolism of Alkaloids, Phenolics and Terpenes; Ramawat, K.G., Mérillon, J.-M., Eds.; Springer: Berlin/Heidelberg, Germany, 2013; pp. 2665–2691. [Google Scholar]
  16. Ying, Z.; Li, X.M.; Yang, S.Q.; Wang, B.G.; Li, H.L.; Meng, L.H. New Polyketide and Sesquiterpenoid Derivatives from the Magellan Seamount-Derived Fungus Penicillium rubens AS-130. Chem. Biodivers. 2023, 20, e202300229. [Google Scholar] [CrossRef] [PubMed]
  17. Ayyad, S.N.; Deyab, M.A.; Kosbar, T.; Alarif, W.M.; Eissa, A.H. Bio-active sesquiterpenoids and norsesquiternoids from the Red Sea octocoral Rhytisma fulvum fulvum. Nat. Prod. Res. 2021, 35, 4303–4310. [Google Scholar] [CrossRef] [PubMed]
  18. 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]
  19. Kaleem, S.; Ge, H.; Yi, W.; Zhang, Z.; Wu, B. Isolation, structural elucidation, and antimicrobial evaluation of the metabolites from a marine-derived fungus Penicillium sp. ZZ1283. Nat. Prod. Res. 2021, 35, 2498–2506. [Google Scholar] [CrossRef]
  20. Chen, T.; Yang, W.; Li, T.; Yin, Y.; Liu, Y.; Wang, B.; She, Z. Hemiacetalmeroterpenoids A-C and Astellolide Q with Antimicrobial Activity from the Marine-Derived Fungus Penicillium sp. N-5. Mar. Drugs 2022, 20, 514. [Google Scholar] [CrossRef]
  21. Liu, X.-H.; Song, Y.-P.; Yin, X.-L.; Ji, N.-Y. Antimicrobial Terpenoids and Polyketides from the Algicolous Fungus Byssochlamys spectabilis RR-dl-2-13. J. Agric. Food Chem. 2022, 70, 4658–4666. [Google Scholar] [CrossRef]
  22. Limon, A.D.; Patabendige, H.; Azhari, A.; Sun, X.; Kyle, D.E.; Wilson, N.G.; Baker, B.J. Chemistry and Bioactivity of the Deep-Water Antarctic Octocoral Alcyonium sp. Mar. Drugs 2022, 20, 576. [Google Scholar] [CrossRef]
  23. Balansa, W.; Mettal, U.; Wuisan, Z.G.; Plubrukarn, A.; Ijong, F.G.; Liu, Y.; Schaberle, T.F. A New Sesquiterpenoid Aminoquinone from an Indonesian Marine Sponge. Mar. Drugs 2019, 17, 158. [Google Scholar] [CrossRef]
  24. Wang, J.; Liu, L.; Hong, L.L.; Zhan, K.X.; Lin, Z.J.; Jiao, W.H.; Lin, H.W. New bisabolane-type phenolic sesquiterpenoids from the marine sponge Plakortis simplex. Chin. J. Nat. Med. 2021, 19, 626–631. [Google Scholar] [CrossRef] [PubMed]
  25. Dong, Y.L.; Li, X.M.; Shi, X.S.; Wang, Y.R.; Wang, B.G.; Meng, L.H. Diketopiperazine Alkaloids and Bisabolene Sesquiterpenoids from Aspergillus versicolor AS-212, an Endozoic Fungus Associated with Deep-Sea Coral of Magellan Seamounts. Mar. Drugs 2023, 21, 293. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, B.; Zhao, Q.; Gu, Y.C.; Lan, L.; Wang, C.Y.; Guo, Y.W. Xishaeleganins A-D, Sesquiterpenoid Hydroquinones from Xisha Marine Sponge Dactylospongia elegans. Mar. Drugs 2022, 20, 118. [Google Scholar] [CrossRef] [PubMed]
  27. Lin, Y.-C.; Chao, C.-H.; Fu, C.-W.; Chiou, S.-F.; Huang, T.-Y.; Yang, Y.-J.; Wu, S.-H.; Chen, S.-L.; Wang, H.-C.; Yu, M.-C.; et al. Computationally assisted structure elucidation of new 2-guanidinoethanesulfonyl sesquiterpenoid alkaloids: Agelasidines G–I from the marine sponge Agelas nakamurai. Tetrahedron 2022, 126, 133077. [Google Scholar] [CrossRef]
  28. Niu, S.; Huang, S.; Wang, J.; He, J.; Chen, M.; Zhang, G.; Hong, B. Malfilanol C, a new sesquiterpenoid isolated from the deep-sea-derived Aspergillus puniceus A2 fungus. Nat. Prod. Res. 2024, 38, 1362–1368. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, Y.; Li, X.M.; Yang, S.Q.; Zhang, F.Z.; Wang, B.G.; Li, H.L.; Meng, L.H. Sesquiterpene and Sorbicillinoid Glycosides from the Endophytic Fungus Trichoderma longibrachiatum EN-586 Derived from the Marine Red Alga Laurencia obtusa. Mar. Drugs 2022, 20, 177. [Google Scholar] [CrossRef] [PubMed]
  30. Wu, Q.; Chang, Y.; Che, Q.; Li, D.; Zhang, G.; Zhu, T. Citreobenzofuran D-F and Phomenone A-B: Five Novel Sesquiterpenoids from the Mangrove-Derived Fungus Penicillium sp. HDN13-494. Mar. Drugs 2022, 20, 137. [Google Scholar] [CrossRef] [PubMed]
  31. El Feky, S.E.; Abd El Hafez, M.S.M.; Abd El Moneim, N.A.; Ibrahim, H.A.H.; Okbah, M.A.; Ata, A.; El Sedfy, A.S.; Hussein, A. Cytotoxic and antimicrobial activities of two new sesquiterpenoids from red sea brittle star Ophiocoma dentata. Sci. Rep. 2022, 12, 8209. [Google Scholar] [CrossRef]
  32. Smanski, M.J.; Peterson, R.M.; Shen, B. Platensimycin and platencin biosynthesis in Streptomyces platensis, showcasing discovery and characterization of novel bacterial diterpene synthases. Methods Enzymol. 2012, 515, 163–186. [Google Scholar]
  33. Tai, C.J.; Ahmed, A.F.; Chao, C.H.; Yen, C.H.; Hwang, T.L.; Chang, F.R.; Huang, Y.M.; Sheu, J.H. Spongenolactones A-C, Bioactive 5,5,6,6,5-Pentacyclic Spongian Diterpenes from the Red Sea Sponge Spongia sp. Mar. Drugs 2022, 20, 498. [Google Scholar] [CrossRef]
  34. Cao, F.; Liu, X.M.; Wang, X.; Zhang, Y.H.; Yang, J.; Li, W.; Luo, D.Q.; Liu, Y.F. Structural diversity and biological activities of indole-diterpenoids from Penicillium janthinellum by co-culture with Paecilomyces formosus. Bioorg. Chem. 2023, 141, 106863. [Google Scholar] [CrossRef] [PubMed]
  35. Kankanamge, S.; Khalil, Z.G.; Bernhardt, P.V.; Capon, R.J. Noonindoles A-F: Rare Indole Diterpene Amino Acid Conjugates from a Marine-Derived Fungus, Aspergillus noonimiae CMB-M0339. Mar. Drugs 2022, 20, 698. [Google Scholar] [CrossRef] [PubMed]
  36. Tani, K.; Kamada, T.; Phan, C.S.; Vairappan, C.S. New cembrane-type diterpenoids from Bornean soft coral Nephthea sp. with antifungal activity against Lagenidium thermophilum. Nat. Prod. Res. 2019, 33, 3343–3349. [Google Scholar] [CrossRef] [PubMed]
  37. Yurchenko, A.N.; Zhuravleva, O.I.; Khmel, O.O.; Oleynikova, G.K.; Antonov, A.S.; Kirichuk, N.N.; Chausova, V.E.; Kalinovsky, A.I.; Berdyshev, D.V.; Kim, N.Y.; et al. New Cyclopiane Diterpenes and Polyketide Derivatives from Marine Sediment-Derived Fungus Penicillium antarcticum KMM 4670 and Their Biological Activities. Mar. Drugs 2023, 21, 584. [Google Scholar] [CrossRef] [PubMed]
  38. Pech-Puch, D.; Forero, A.M.; Fuentes-Monteverde, J.C.; Lasarte-Monterrubio, C.; Martinez-Guitian, M.; Gonzalez-Salas, C.; Guillen-Hernandez, S.; Villegas-Hernandez, H.; Beceiro, A.; Griesinger, C.; et al. Antimicrobial Diterpene Alkaloids from an Agelas citrina Sponge Collected in the Yucatan Peninsula. Mar. Drugs 2022, 20, 298. [Google Scholar] [CrossRef]
  39. Yan, X.; Ouyang, H.; Li, T.; Shi, Y.; Wu, B.; Yan, X.; He, S. Six New Diterpene Glycosides from the Soft Coral Lemnaliabournei. Mar. Drugs 2021, 19, 339. [Google Scholar] [CrossRef] [PubMed]
  40. Han, X.; Wang, H.; Li, B.; Chen, X.; Li, T.; Yan, X.; Ouyang, H.; Lin, W.; He, S. New Diterpenes and Diterpene Glycosides with Antibacterial Activity from Soft Coral Lemnalia bournei. Mar. Drugs 2024, 22, 157. [Google Scholar] [CrossRef] [PubMed]
  41. Chen, B.; Qiu, P.; Xu, B.; Zhao, Q.; Gu, Y.C.; Fu, L.; Bi, S.; Lan, L.; Wang, C.Y.; Guo, Y.W. Cytotoxic and Antibacterial Isomalabaricane Terpenoids from the Sponge Rhabdastrella globostellata. J. Nat. Prod. 2022, 85, 1799–1807. [Google Scholar] [CrossRef]
  42. Cheng, X.; Liang, X.; Yao, F.H.; Liu, X.B.; Qi, S.H. Fusidane-Type Antibiotics from the Marine-Derived Fungus Simplicillium sp. SCSIO 41513. J. Nat. Prod. 2021, 84, 2945–2952. [Google Scholar] [CrossRef]
  43. Gomm, A.; Nelson, A. A radical approach to diverse meroterpenoids. Nat. Chem. 2020, 12, 109–111. [Google Scholar] [CrossRef]
  44. Russo, D.; Milella, L. Chapter 14—Analysis of Meroterpenoids. In Recent Advances in Natural Products Analysis; Sanches Silva, A., Nabavi, S.F., Saeedi, M., Nabavi, S.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 477–501. [Google Scholar]
  45. Nazir, M.; Saleem, M.; Tousif, M.I.; Anwar, M.A.; Surup, F.; Ali, I.; Wang, D.; Mamadalieva, N.Z.; Alshammari, E.; Ashour, M.L.; et al. Meroterpenoids: A Comprehensive Update Insight on Structural Diversity and Biology. Biomolecules 2021, 11, 957. [Google Scholar] [CrossRef] [PubMed]
  46. Hu, X.Y.; Li, X.M.; Liu, H.; Wang, B.G.; Meng, L.H. Mining new meroterpenoids from the marine red alga-derived endophytic Penicillium chermesinum EN-480 by comparative transcriptome analysis. Bioorg. Chem. 2022, 128, 106021. [Google Scholar] [CrossRef]
  47. Hong, X.; Guan, X.; Lai, Q.; Yu, D.; Chen, Z.; Fu, X.; Zhang, B.; Chen, C.; Shao, Z.; Xia, J.; et al. Characterization of a bioactive meroterpenoid isolated from the marine-derived fungus Talaromyces sp. Appl. Microbiol. Biotechnol. 2022, 106, 2927–2935. [Google Scholar] [CrossRef] [PubMed]
  48. Tang, L.; Xia, J.; Chen, Z.; Lin, F.; Shao, Z.; Wang, W.; Hong, X. Cytotoxic and Antibacterial Meroterpenoids Isolated from the Marine-Derived Fungus Talaromyces sp. M27416. Mar. Drugs 2024, 22, 186. [Google Scholar] [CrossRef] [PubMed]
  49. Chen, X.Y.; Zeng, Q.; Chen, Y.C.; Zhong, W.M.; Xiang, Y.; Wang, J.F.; Shi, X.F.; Zhang, W.M.; Zhang, S.; Wang, F.Z. Chevalones H-M: Six New alpha-Pyrone Meroterpenoids from the Gorgonian Coral-Derived Fungus Aspergillus hiratsukae SCSIO 7S2001. Mar. Drugs 2022, 20, 71. [Google Scholar] [CrossRef] [PubMed]
  50. Ren, Z.; Yang, L.; Ma, Q.; Xie, Q.; Dai, H.; Sun, K.; Zhao, Y. Meroterpenoids and Steroids from the Marine-Derived Fungus Trametes sp. ZYX-Z-16. Molecules 2022, 27, 8782. [Google Scholar] [CrossRef]
  51. Ryu, M.J.; Hillman, P.F.; Lee, J.; Hwang, S.; Lee, E.Y.; Cha, S.S.; Yang, I.; Oh, D.C.; Nam, S.J.; Fenical, W. Antibacterial Meroterpenoids, Merochlorins G-J from the Marine Bacterium Streptomyces sp. Mar. Drugs 2021, 19, 618. [Google Scholar] [CrossRef] [PubMed]
  52. Cen, S.; Jia, J.; Ge, Y.; Ma, Y.; Li, X.; Wei, J.; Bai, Y.; Wu, X.; Song, J.; Bi, H.; et al. A new antibacterial 3,5-dimethylorsellinic acid-based meroterpene from the marine fungus Aspergillus sp. CSYZ-1. Fitoterapia 2021, 152, 104908. [Google Scholar] [CrossRef] [PubMed]
  53. Xu, K.; Wei, X.L.; Xue, L.; Zhang, Z.F.; Zhang, P. Antimicrobial Meroterpenoids and Erythritol Derivatives Isolated from the Marine-Algal-Derived Endophytic Fungus Penicillium chrysogenum XNM-12. Mar. Drugs 2020, 18, 578. [Google Scholar] [CrossRef]
  54. Singh, K.; Kaur, G.; Shanika, P.S.; Dziwornu, G.A.; Okombo, J.; Chibale, K. Structure-activity relationship analyses of fusidic acid derivatives highlight crucial role of the C-21 carboxylic acid moiety to its anti-mycobacterial activity. Bioorg. Med. Chem. 2020, 28, 115530. [Google Scholar] [CrossRef]
Figure 1. Illustration of statistical data analysis: (a) representative terpenoid compounds with antimicrobial activity; (b) count of antimicrobial strains and their corresponding compounds. (c) Terpenoid compounds with selective antimicrobial activity against targeted bacterial strains.
Figure 1. Illustration of statistical data analysis: (a) representative terpenoid compounds with antimicrobial activity; (b) count of antimicrobial strains and their corresponding compounds. (c) Terpenoid compounds with selective antimicrobial activity against targeted bacterial strains.
Marinedrugs 22 00347 g001
Figure 2. Chemical structures of sesquiterpenoids (148).
Figure 2. Chemical structures of sesquiterpenoids (148).
Marinedrugs 22 00347 g002
Figure 3. Chemical structures of diterpenoids (4987).
Figure 3. Chemical structures of diterpenoids (4987).
Marinedrugs 22 00347 g003
Figure 4. Chemical structures of triterpenoids (88107).
Figure 4. Chemical structures of triterpenoids (88107).
Marinedrugs 22 00347 g004
Figure 5. Chemical structures of meroterpenoids (108141).
Figure 5. Chemical structures of meroterpenoids (108141).
Marinedrugs 22 00347 g005
Table 1. Names, classes, skeletons, and marine sources of sesquiterpenoids (148).
Table 1. Names, classes, skeletons, and marine sources of sesquiterpenoids (148).
No.NamesClassesMarine SourcesRef.
1Chermesiterpenoid DLinear SesquiterpenoidMagellan Seamount-Derived fungus Penicillium rubens AS-130[16]
212-O-acetyl-nardosinan-6-en-1-oneNardosinane SesquiterpeneOctocoral Rhytisma fulvum fulvum[17]
36β-acetyl-1(10)-α-13-nornardosin-7-one[17]
46α-acetyl-1(10)-α-13-nornardosin-7-one[17]
5Lineolemnene ENeolemnane SesquiterpeneSoft coral Lemnalia sp.[18]
6Lineolemnene F
7Lineolemnene G
82-acetoxy-aristolaneAristolane SesquiterpenoidSoft coral Lemnalia sp.[18]
9Lactone purpuride DDrimane SesquiterpeneThe marine-derived Penicillium sp. ZZ1283[19]
10Astellolide QDrimane SesquiterpeneThe marine-derived fungus Penicillium sp. N-5[20]
11Byssocarotin ACarotane SesquiterpenoidMacroalga-Derived Algicolous Fungus Penicillium rubens RR-dl-2-13[21]
12Byssocarotin B[21]
13Byssocarotin C[21]
14Byssocarotin D[21]
15Alcyopterosin TIlludalane SesquiterpenoidOctocoral Alcyonium sp.[22]
[22]
16Alcyopterosin U[22]
17Alcyopterosin V[22]
18Nakijiquinone VSesquiterpenoid AminoquinoneIndonesian marine Dactylospongia elegans sponge[23]
19IllimaquinoneMerosesquiterpenoidIndonesian marine Dactylospongia elegans sponge[23]
20Smenospongine[23]
21Dyctioceratine C[23]
22Plakordiol ABisabolane Phenolic SesquiterpenoidThe marine sponge Plakortis simplex[24]
23Plakordiol B[24]
24Plakordiol C[24]
25Plakordiol D[24]
26(7R, 10R)-hydroxycurcudiol[24]
27(7R, 10S)-hydroxycurcudiol[24]
28Sydonic acidBisabolene SesquiterpenoidAspergillus versicolor AS-212[25]
29(S)-(+)-11-dehydrosydonic acid[25]
30(−)-10-hydroxysydonic acid[25]
31hydroxysydonic acid[25]
32Peniciaculin B[25]
33Xishaeleganins ASesquiterpenoid HydroquinoneXisha Marine Sponge Dactylospongia elegans[26]
34Xishaeleganins B[26]
35Xishaeleganins C[26]
36Xishaeleganins D[26]
37Agelasidine GSesquiterpenoid AlkaloidSponge Agelas nakamurai[27]
38Agelasidine H[27]
39Agelasidine I[27]
40Malfilanol CSesquiterpenoidThe deep-sea-derived fungus Aspergillus puniceus A2[28]
41Trichoacorside ASesquiterpene GlycosideRed Alga Laurencia obtuse -Derived Endophytic Fungus Trichoderma longibrachiatum EN-586 [29]
42Citreobenzofuran DSesquiterpenoidMangrove-Derived Fungus Penicillium sp. HDN13-494[30]
43Citreobenzofuran E[30]
44Citreobenzofuran F[30]
45Phomenone A[30]
46Phomenone B[30]
47O8-ophiocomaneSesquiterpenoidBrittle star; Ophiocoma dentata[31]
48O7-ophiocomane[31]
Table 2. Names, classes, skeletons, and marine sources of diterpenoids (5088).
Table 2. Names, classes, skeletons, and marine sources of diterpenoids (5088).
No.NamesClassesMarine SourcesRef.
49Spongenolactone A5,5,6,6,5-Pentacyclic Spongian DiterpenesRed Sea sponge Spongia sp.[33]
50Spongenolactone B[33]
51Spongenolactone C[33]
52Janthinellumine AIndole DiterpeneCo-culturing the marine-derived fungi Penicillium janthinellium with Paecilomyces formosus[34]
53Janthinellumine B[34]
54Janthinellumine C[34]
55Janthinellumine D[34]
56Janthinellumine E[34]
57Janthinellumine F[34]
58Janthinellumine G[34]
59Janthinellumine H[34]
60Janthinellumine I[34]
61Noonindole AIndole Diterpene Amino AcidFungus Aspergillus noonimiae CMB-M0339[35]
62Nephthecrassocolide ACembrane DiterpeneBornean soft coral Nephthea sp.[36]
63Nephthecrassocolide B[36]
646-Acetoxy Nephthenol Acetate[36]
654-Hydroxyleptosphin CCyclopiane DiterpeneThe marine sediment-derived fungus Penicillium antarcticum KMM 4670[37]
6613-Epi-Conidiogenone F[37]
67(+)-8-Epiagelasine TDiterpene AlkaloidAgelas citrina Sponge[38]
68(+)-10-Epiagelasine B[38]
69(+)-12-Hydroxyagelasidine C[38]
70Lemnabourside EBicyclic Diterpene GlycosideSoft coral Lemnalia bournei[39]
71Lemnabourside F[39]
72Lemnabourside G[39]
73Lemnadiolbourside A[39]
74Lemnadiolbourside B[39]
75Lemnadiolbourside C[39]
76Biofloranate EBiflorane-Type DiterpenoidSoft coral Lemnalia bournei[40]
77Biofloranate F[40]
78Biofloranate G[40]
79Biofloranate H[40]
80Biofloranate I[40]
81Lemnabourside HBicyclic Diterpene GlycosideSoft coral Lemnalia bournei[40]
82Lemnabourside I[40]
83Biofloranate ADecalin-Type Bicyclic DiterpeneSoft coral Lemnalia sp.[18]
84Biofloranate B[18]
85Biofloranate C[18]
86Biofloranate D[18]
87Cneorubin KAromadendrane-Type DiterpenoidSoft coral Lemnalia sp.[18]
Table 3. Names, classes, skeletons, and marine sources of triterpenoids (88107).
Table 3. Names, classes, skeletons, and marine sources of triterpenoids (88107).
No.NamesClassesMarine SourcesRef.
8813-(E)-geoditin AIsomalabaricane TerpenoidSponge Rhabdastrella globostellata[41]
8913-(E)-isogeoditin B[41]
903-Acetylstelliferin D[41]
9129-Acetylstelliferin D[41]
92Hainanstelletin A[41]
93Hainanstelletin B[41]
9423,24-Ene-25-hydroxystelliferin D[41]
9525,26-Ene-24-hydroxystelliferin D[41]
96Hainanstelletin C[41]
97Implifusidic acid AFusidane-Type NortriterpenoidThe marine-derived fungus Simplicillium sp. SCSIO 41513.[42]
98Implifusidic acid B[42]
99Implifusidic acid C[42]
100Implifusidic acid D[42]
101Implifusidic acid E[42]
102Implifusidic acid F[42]
103Implifusidic acid G[42]
104Implifusidic acid H[42]
105Implifusidic acid I[42]
106Implifusidic acid J[42]
107Implifusidic acid K[42]
Table 4. Names, classes, skeletons, and marine sources of meroterpenoids (108141).
Table 4. Names, classes, skeletons, and marine sources of meroterpenoids (108141).
No.NamesClassesMarine SourcesRef.
108Chermesin EMeroterpenoidRed alga-derived endophytic Penicillium chermesinum EN-480[46]
109Chermesin F[46]
110Chermesin G[46]
111Chermesin H[46]
112Taladrimanin ADrimane-Type MeroterpenoidFungus Talaromyces sp. HM6-1–1[47]
116Taladrimanin BMeroterpenoidThe marine-derived fungus Talaromyces sp. M27416[48]
119Chevalone Hα-Pyrone MeroterpenoidGorgonian coral-derived fungus Aspergillus hiratsukae SCSIO 7S2001[49]
120Chevalone I[49]
121Chevalone J[49]
122Chevalone K[49]
123Chevalone L[49]
124Chevalone M[49]
125Asnovolin C 5′6′-dehydrohydrogenSpiromeroterpenoidConch snail-derived fungus Trametes sp. ZYX-Z-16[50]
126Asnovolin C[50]
127Chermesin A[50]
128Chrodrimanin E[50]
129Chrodrimanin H[50]
130Thailandolide B[50]
131Asnovolin H[50]
132Asnovolin I[50]
133Hemiacetalmeroterpenoid AAndrastin-Type MeroterpenoidThe marine-derived fungus Penicillium sp. N-5[20]
134Hemiacetalmeroterpenoid B[20]
135Hemiacetalmeroterpenoid C[20]
136Merochlorin GChlorinated MeroterpenoidMarine sediment-derived bacterium strain Streptomyces sp. CNH-189[51]
137Merochlorin H[51]
138Merochlorin I[51]
139Merochlorin J[51]
140AspergillactoneMeroterpenoidThe marine fungus Aspergillus sp. CSYZ-1[52]
141Oxalicine CMeroterpenoid-Type AlkaloidThe marine-algal-derived endophytic fungus Penicillium chrysogenum XNM-12[53]
Table 5. Antibacterial and/or antifungal activities of sesquiterpenoids (148).
Table 5. Antibacterial and/or antifungal activities of sesquiterpenoids (148).
No.Test StrainsActivityBioassaysRef.
1MRSAAntibacterialMIC = 64 µg/mL[16]
2B. cereusAntibacterialdiameters of inhibition zone 6 ± 0.03 mm (50 µg/mL)[17]
S. aureusdiameters of inhibition zone 5 ± 0.00 mm (50 µg/mL)
E. colinegative
Pseudomonas sp.diameters of inhibition zone 4 ± 0.00 mm (50 µg/mL)
3B. cereusAntibacterialdiameters of inhibition zone 6 ± 0.00 mm (100 µg/mL)[17]
S. aureusdiameters of inhibition zone 5 ± 0.00 mm (100 µg/mL)
E. colidiameters of inhibition zone 4 ± 0.00 mm (100 µg/mL)
Pseudomonas sp.negative
4B. cereusAntibacterialdiameters of inhibition zone 6 ± 0.00 mm (100 µg/mL)[17]
S. aureusdiameters of inhibition zone 5 ± 0.00 mm (100 µg/mL)
E. colidiameters of inhibition zone 4 ± 0.00 mm (100 µg/mL)
Pseudomonas sp.negative
5S. aureusAntibacterialMIC > 128 µg/mL[18]
B. cereus
6S. aureusAntibacterialMIC > 128 µg/mL[18]
B. cereus
7S. aureusAntibacterialMIC > 128 µg/mL[18]
B. cereus
8S. aureusAntibacterialMIC > 128 µg/mL[18]
B. cereus
9MRSAAntibacterialMIC = 4 µg/mL[19]
E. coliMIC = 3 µg/mL
C. albicansAntifungalMIC = 8 µg/mL
10MRSAAntibacterialMIC >50 µg/mL[20]
B. cereus
P. italicumAntifungalMIC = 25 µg/mL
C. gloeosporioides
11V. anguillarumAntibacterialdiameters of inhibition zone 6.3 ± 0.6 mm (50 µg/disk)[21]
V. harveyinegative
V. parahaemolyticusdiameters of inhibition zone 6.7 ± 0.6 mm (50 µg/disk)
12V. anguillarumAntibacterialdiameters of inhibition zone 6.7 ± 0.6 mm (50 µg/disk)[21]
V. harveyinegative
V. parahaemolyticusdiameters of inhibition zone 7.3 ± 0.6 mm (50 µg/disk)
13V. anguillarumAntibacterialnegative[21]
V. harveyinegative
V. parahaemolyticusdiameters of inhibition zone 7.3 ± 0.6 mm (50 µg/disk)
14V. anguillarumAntibacterialnegative[21]
V. harveyinegative
V. parahaemolyticusdiameters of inhibition zone 6.7 ± 0.6 mm (50 µg/disk)
15ESKAPEInactiveinactive against the ESKAPE[22]
[22]
16ESKAPEInactiveinactive against the ESKAPE[22]
17C. difficileAntibacterialMIC 8.1 µg/mL[22]
ESKAPEinactive against the ESKAPE
18B. megaterium DSM32Inactiveinactive against B. Megaterium DSM32[23]
M. luteus ATCC 4698inactive against M. Luteus ATCC 4698
19B. megaterium DSM32AntibacterialMIC = 32 µg/mL[23]
M. luteus ATCC 4698
20B. megaterium DSM32AntibacterialMIC = 32 µg/mL[23]
M. luteus ATCC 4698
21B. megaterium DSM32AntibacterialMIC = 32 µg/mL[23]
M. luteus ATCC 4698MIC = 64 µg/mL
22S. aureus ATCC 25923AntibacterialMIC > 64 µg/mL[24]
MRSA ATCC 43300
A. baumannii ATCC19606
P. aeruginosa (clinical)
VRE CD27
23S. aureus ATCC 25923AntibacterialMIC > 64 µg/mL[24]
MRSA ATCC 43300
A. baumannii ATCC19606
P. aeruginosa (clinical)
VRE CD27
24S. aureus ATCC 25923AntibacterialMIC > 64 µg/mL[24]
MRSA ATCC 43300
A. baumannii ATCC19606
P. aeruginosa (clinical)
VRE CD27
25S. aureus ATCC 25923AntibacterialMIC > 64 µg/mL[24]
MRSA ATCC 43300
A. baumannii ATCC19606
P. aeruginosa (clinical)
VRE CD27
26S. aureus ATCC 25923AntibacterialMIC > 64 µg/mL[24]
MRSA ATCC 43300
A. baumannii ATCC19606
P. aeruginosa (clinical)
VRE CD27
27A. baumannii ATCC19606Antibacterialdiameters of inhibition zone 5.0 ± 0.6 mm, but MIC > 64 µg/mL[24]
28V. harveyiAntibacterialMIC = 15.0 µg/mL[25]
V. Parahaemolyticus
C. GloeosporioidesAntifungalMIC = 120.3 µg/mL
29V. harveyiAntibacterialMIC = 15.2 µg/mL[25]
V. ParahaemolyticusMIC = 121.2 µg/mL
C. GloeosporioidesAntifungalMIC = 121.2 µg/mL
30V. harveyiAntibacterialMIC = 28.4 µg/mL[25]
V. ParahaemolyticusMIC = 113.5 µg/mL
C. GloeosporioidesAntifungalMIC > 200 µg/mL
31V. harveyiAntibacterialMIC > 200 µg/mL[25]
V. Parahaemolyticus
C. GloeosporioidesAntifungalMIC > 200 µg/mL
32V. harveyiAntibacterialMIC > 200 µg/mL[25]
V. ParahaemolyticusMIC = 64 µg/mL
C. GloeosporioidesAntifungalMIC > 200 µg/mL
33S. aureus USA300 LACInactiveinactive against S. aureus USA300 LAC[26]
S. pyogenes ATCC 12344inactive against S. pyogenes ATCC 12344
E. Faecium Efm-HS0649inactive against E. Faecium Efm-HS0649
34S. aureus USA300 LACAntibacterialMIC = 1.5 µg/mL[26]
S. pyogenes ATCC 12344MIC = 1.5 µg/mL
E. Faecium Efm-HS0649MIC = 3.0 µg/mL
35S. aureus USA300 LACAntibacterialMIC = 11.1 µg/mL[26]
S. pyogenes ATCC 12344MIC = 2.8 µg/mL
E. Faecium Efm-HS0649MIC = 5.6 µg/mL
36S. aureus USA300 LACAntibacterialMIC > 186.0 µg/mL[26]
S. pyogenes ATCC 12344MIC = 11.6 µg/mL
E. Faecium Efm-HS0649MIC > 186.0 µg/mL
37B. subtilisAntibacterialdiameters of inhibition zone 3.0 mm (25 mg/disk)[27]
E. coli
K. Pneumoniae
S. aureus
38B. subtilisInactiveinactive against B. subtilis[27]
E. coliinactive against E. coli
K. Pneumoniaeinactive against K. Pneumoniae
S. aureusinactive against S. aureus
40S. aureus ATCC 29213Antibacterialdiameters of inhibition zone 8 mm (200 mg/disk)[28]
41E. coliAntibacterialMIC > 64 µg/mL[29]
MRSAMIC = 64 µg/mL
P. AeruginosaMIC > 64 µg/mL
V. harveyiMIC = 4 µg/mL
V. ParahaemolyticusMIC > 64 µg/mL
A. BrassicaeAntifungalMIC = 32 µg/mL
C. CornigerumMIC = 64 µg/mL
C. GloeosporioidesMIC = 16 µg/mL
C. Gloeosporioides PenzMIC = 16 µg/mL
Curvularia spiciferaMIC = 8 µg/mL
F. GraminearumMIC > 64 µg/mL
F. OxysporumMIC = 32 µg/mL
F. Oxysporum f. Sp. Radicis lycopersiciMIC = 32 µg/mL
Fusarium proliferatumMIC = 32 µg/mL
P. DigitatumMIC = 64 µg/mL
P. Piricola NoseMIC = 32 µg/mL
A. hydrophiliaMIC = 64 µg/mL
42B. subtilisAntibacterialMIC > 50 µg/mL[30]
A. Baumannii
E. coil
MRSA
C. albicansAntifungalMIC > 50 µg/mL
43B. subtilisAntibacterialMIC > 50 µg/mL[30]
A. Baumannii
E. coil
MRSA
C. albicansAntifungalMIC > 50 µg/mL
44B. subtilisAntibacterialMIC > 50 µg/mL[30]
A. Baumannii
E. coil
MRSA
C. albicansAntifungalMIC > 50 µg/mL
45B. subtilisAntibacterialMIC 6.25 µg/mL[30]
A. BaumanniiMIC > 50 µg/mL
E. coilMIC > 50 µg/mL
MRSAMIC > 50 µg/mL
C. albicansAntifungalMIC > 50 µg/mL
46B. subtilisAntibacterialMIC > 50 µg/mL[30]
A. Baumannii
E. coil
MRSA
C. albicansAntifungalMIC > 50 µg/mL
47P. AeruginosaAntibacterial2.25 ± 0.04 mm AU[31]
E. Faecalis1.36 ± 0.04 mm AU
48P. AeruginosaAntibacterial2.8 ± 0.05 mm AU[31]
E. Faecalis1.8 ± 0.02 mm AU
Table 6. Antibacterial and/or Antifungal Activities Diterpenoids (4987).
Table 6. Antibacterial and/or Antifungal Activities Diterpenoids (4987).
No.Test StrainsActivityBioassaysRef.
49S. aureusAntibacterial24% (50 µM), 42% (100 µM), 40% (200 µM) inhibition[33]
50S. aureusAntibacterial46% (50 µM), 47% (100 µM), 93% (200 µM) inhibition[33]
51S. aureusInactiveInactive against S. aureus[33]
52V. anguillarumAntibacterialMIC = 12.5 µg/mL[34]
53V. anguillarumInactiveInactive against V. anguillarum[34]
54V. anguillarumInactiveInactive against V. anguillarum[34]
55V. anguillarumInactiveInactive against V. anguillarum[34]
56V. anguillarumInactiveInactive against V. anguillarum[34]
57V. anguillarumInactiveInactive against V. anguillarum[34]
58V. anguillarumInactiveInactive against V. anguillarum[34]
59V. anguillarumAntibacterialMIC = 12.5 µg/mL[34]
60V. anguillarumInactiveInactive against V. anguillarum[34]
61C. albicansAntifungal-[35]
E. coli ATCC 11775Inactive-
S. aureus ATCC 25923
B. subtilis ATCC 6633
62Exophiala sp. NJM 1551AntifungalMIC = 25 µg/mL[36]
Fusarium moniliforme NJM 8995MIC > 100 µg/mL
F. Oxysporum NJM 0179MIC = 50 µg/mL
Fusarium solani NJM 8996MIC = 50 µg/mL
Haliphthoros sabahensis IPMB 1402MIC = 25 µg/mL
H. Milfordensis IPMB 1603MIC = 25 µg/mL
Lagenidium thermophilum IPMB 1401MIC = 12.5 µg/mL
63Exophiala sp. NJM 1551AntifungalMIC = 50 µg/mL[36]
Fusarium moniliforme NJM 8995MIC > 100 µg/mL
F. Oxysporum NJM 0179MIC > 100 µg/mL
Fusarium solani NJM 8996MIC > 100 µg/mL
Haliphthoros sabahensis IPMB 1402MIC = 25 µg/mL
H. Milfordensis IPMB 1603MIC = 50 µg/mL
Lagenidium thermophilum IPMB 1401MIC = 25 µg/mL
64Exophiala sp. NJM 1551AntifungalMIC = 50 µg/mL[36]
Fusarium moniliforme NJM 8995MIC > 100 µg/mL
F. Oxysporum NJM 0179MIC > 100 µg/mL
Fusarium solani NJM 8996MIC > 100 µg/mL
Haliphthoros sabahensis IPMB 1402MIC = 50 µg/mL
H. Milfordensis IPMB 1603MIC = 50 µg/mL
Lagenidium thermophilum IPMB 1401MIC = 25 µg/mL
65S. aureusAntibacterial0 (12.5 µM), 19.1% (100 µM) inhibition[37]
66S. aureusAntibacterial15.3% (12.5 µM), 29.3% (100 µM) inhibition[37]
67S. aureus ATCC 29213AntibacterialMIC = 16 µg/mL[38]
S. aureus USA300LACMIC = 16 µg/mL
S. Pneumoniae ATCC 49619MIC = 16 µg/mL
S. Pneumoniae 549 CHUACMIC = 32 µg/mL
E. Faecalis ATCC 29212MIC = 32 µg/mL
E. Faecalis 256 CHUACMIC > 64 µg/mL
E. Faecium 214 CHUACMIC = 32 µg/mL
68S. aureus ATCC 29213AntibacterialMIC = 1 µg/mL[38]
S. aureus USA300LACMIC = 2 µg/mL
S. Pneumoniae ATCC 49619MIC = 4 µg/mL
S. Pneumoniae 549 CHUACMIC = 8 µg/mL
E. Faecalis ATCC 29212MIC = 4 µg/mL
E. Faecalis 256 CHUACMIC = 4 µg/mL
E. Faecium 214 CHUACMIC = 4 µg/mL
69S. aureus ATCC 29213AntibacterialMIC = 8 µg/mL[38]
S. aureus USA300LACMIC = 8 µg/mL
S. Pneumoniae ATCC 49619MIC = 16 µg/mL
S. Pneumoniae 549 CHUACMIC > 64 µg/mL
E. Faecalis ATCC 29212MIC = 16 µg/mL
E. Faecalis 256 CHUACMIC = 32 µg/mL
E. Faecium 214 CHUACMIC = 8 µg/mL
70S. aureusAntibacterialMIC = 8 µg/mL[39]
B. subtilisMIC = 4 µg/mL
71S. aureusAntibacterialMIC = 8 µg/mL[39]
B. subtilisMIC = 4 µg/mL
72S. aureusInactiveMIC > 128 µg/mL[39]
B. subtilisMIC = 64 µg/mL
73S. aureusInactiveMIC > 128 µg/mL[39]
B. subtilisMIC = 64 µg/mL
74S. aureusInactiveMIC > 128 µg/mL[39]
B. subtilis
75S. aureusInactiveMIC > 128 µg/mL[39]
B. subtilisMIC = 64 µg/mL
76S. aureusAntibacterialMIC = 32 µg/mL[40]
B. subtilisMIC = 32 µg/mL
V. harveyiMIC = 64 µg/mL
S. PneumoniaeMIC > 128 µg/mL
E. coliMIC > 128 µg/mL
77S. aureusAntibacterialMIC = 32 µg/mL[40]
B. subtilisMIC = 32 µg/mL
V. harveyiMIC = 64 µg/mL
S. PneumoniaeMIC > 128 µg/mL
E. coliMIC > 128 µg/mL
78S. aureusAntibacterialMIC = 64 µg/mL[40]
B. subtilisMIC = 32 µg/mL
V. harveyiMIC > 128 µg/mL
S. PneumoniaeMIC > 128 µg/mL
E. coliMIC > 128 µg/mL
79S. aureusAntibacterialMIC = 64 µg/mL[40]
B. subtilisMIC = 64 µg/mL
V. harveyiMIC > 128 µg/mL
S. PneumoniaeMIC > 128 µg/mL
E. coliMIC > 128 µg/mL
80S. aureusAntibacterialMIC = 64 µg/mL[40]
B. subtilisMIC = 32 µg/mL
V. harveyiMIC > 128 µg/mL
S. PneumoniaeMIC > 128 µg/mL
E. coliMIC > 128 µg/mL
81S. aureusAntibacterialMIC = 16 µg/mL[40]
B. subtilisMIC = 16 µg/mL
V. harveyiMIC > 128 µg/mL
S. PneumoniaeMIC > 128 µg/mL
E. coliMIC = 8 µg/mL
82S. aureusAntibacterialMIC = 32 µg/mL[40]
B. subtilisMIC = 16 µg/mL
V. harveyiMIC = 64 µg/mL
S. PneumoniaeMIC > 128 µg/mL
E. coliMIC = 32 µg/mL
83S. aureusAntibacterialMIC = 8 µg/mL[18]
B. Cereus
84S. aureusAntibacterialMIC = 4 µg/mL[18]
B. CereusMIC = 16 µg/mL
85S. aureusAntibacterialMIC = 4 µg/mL[18]
B. CereusMIC = 16 µg/mL
86S. aureusAntibacterialMIC = 16 µg/mL[18]
B. CereusMIC = 8 µg/mL
87S. aureusAntibacterialMIC = 16 µg/mL[18]
B. CereusMIC = 8 µg/mL
“-” indicates that the MIC value is beyond the detectable range.
Table 7. Antibacterial and/or Antifungal Activities Triterpenoids (88105).
Table 7. Antibacterial and/or Antifungal Activities Triterpenoids (88105).
No.Test StrainsActivityBioassaysRef.
88S. aureus USA300LACAntibacterialMIC = 28.1 µg/mL[41]
S. pyogenes ATCC12344MIC = 1.8 µg/mL
89S. aureus USA300LACAntibacterialMIC = 30.9 µg/mL[41]
S. pyogenes ATCC12344MIC = 1.0 µg/mL
90S. aureus USA300LACAntibacterialMIC > 250.0 µg/mL[41]
S. pyogenes ATCC12344MIC = 120.0 µg/mL
91S. aureus USA300LACAntibacterialMIC > 250.0 µg/mL[41]
S. pyogenes ATCC12344
92S. aureus USA300LACAntibacterialMIC > 250.0 µg/mL[41]
S. pyogenes ATCC12344MIC = 65.9 µg/mL
93S. aureus USA300LACAntibacterialMIC > 250.0 µg/mL[41]
S. pyogenes ATCC12344MIC = 124.5 µg/mL
94S. aureus USA300LACAntibacterialMIC > 250.0 µg/mL[41]
S. pyogenes ATCC12344
95S. aureus USA300LACAntibacterialMIC > 250.0 µg/mL[41]
S. pyogenes ATCC12344
96S. aureus USA300LACAntibacterialMIC > 250.0 µg/mL[41]
S. pyogenes ATCC12344MIC = 240.0 µg/mL
98S. aureusAntibacterialMIC > 100 µg/mL[42]
100S. aureusAntibacterialMIC > 100 µg/mL[42]
101S. aureusAntibacterialMIC > 100 µg/mL[42]
102S. aureusAntibacterialMIC > 100 µg/mL[42]
104S. aureusAntibacterialMIC = 2.5 µg/mL[42]
105S. aureusAntibacterialMIC = 0.078 µg/mL[42]
Table 8. Antibacterial and/or antifungal activities meroterpenoids (108141).
Table 8. Antibacterial and/or antifungal activities meroterpenoids (108141).
No.Test StrainsActivityBioassaysRef.
108A. hydrophiliaAntibacterialMIC = 32 µg/mL[46]
E. coliMIC = 16 µg/mL
E. TardaMIC = 0.5 µg/mL
V. anguillarumMIC = 0.5 µg/mL
V. harveyiMIC = 32 µg/mL
C. Diplodiella, AntifungalMIC = 8 µg/mL
F. GraminearumMIC = 32 µg/mL
109A. hydrophiliaAntibacterialMIC = 16 µg/mL
E. coliMIC = 1 µg/mL[46]
E. TardaMIC = 32 µg/mL
V. anguillarumMIC = 4 µg/mL
V. harveyiMIC = 32 µg/mL
C. DiplodiellaAntifungalMIC = 64 µg/mL
F. Graminearum
110A. hydrophiliaAntibacterialMIC = 32 µg/mL[46]
E. coliMIC = 32 µg/mL
E. TardaMIC = 16 µg/mL
V. anguillarumMIC = 32 µg/mL
V. harveyiMIC = 16 µg/mL
C. DiplodiellaAntifungalMIC > 64 µg/mL
F. GraminearumMIC > 32 µg/mL
111A. hydrophiliaAntibacterialMIC = 32 µg/mL[46]
E. coliMIC = 16 µg/mL
E. TardaMIC = 0.5 µg/mL
V. anguillarumMIC = 0.5 µg/mL
V. harveyiMIC = 32 µg/mL
C. DiplodiellaAntifungalMIC = 8 µg/mL
F. GraminearumMIC = 32 µg/mL
112S. aureus ATCC6538PAntibacterialMIC = 15.2 µg/mL[47]
E. coli-
V. ParahaemolyticusAntifungal-
116S. aureus CICC 10384AntibacterialMIC = 12.5 µg/mL[48]
119M. luteaAntibacterialMIC = 6.25 µg/mL[49]
K. PneumoniaeMIC = 50 µg/mL
MRSAMIC = 6.25 µg/mL
S. faecalisMIC = 6.25 µg/mL
120M. luteaAntibacterialMIC = 25 µg/mL[49]
K. PneumoniaeMIC > 100 µg/mL
MRSAMIC = 6.25 µg/mL
S. faecalisMIC = 25 µg/mL
121M. luteaAntibacterialMIC = 25 µg/mL[49]
K. PneumoniaeMIC = 25 µg/mL
MRSAMIC = 12.5 µg/mL
S. faecalisMIC > 100 µg/mL
122M. luteaAntibacterialMIC > 100 µg/mL[49]
K. PneumoniaeMIC = 6.25 µg/mL
MRSAMIC = 25 µg/mL
S. faecalisMIC = 50 µg/mL
123M. luteaAntibacterialMIC = 12.5 µg/mL[49]
K. PneumoniaeMIC > 100 µg/mL
MRSAMIC = 12.5 µg/mL
S. faecalisMIC = 12.5 µg/mL
124M. luteaInactiveMIC > 100 µg/mL[49]
K. Pneumoniae
MRSA
S. faecalis
125S. aureus ATCC6538AntibacterialMIC > 128 µg/mL[50]
B. subtilis ATCC 6633
E. coli ATCC 25922
L. Monocytogenes ATCC 1911
F. Oxysporum f. Sp. CubenseInactiveMIC > 128 µg/mL
Fusarium spp.
P. Litchii
C. Gloeosporioides
H. Undatus
126S. aureus ATCC6538AntibacterialMIC > 128 µg/mL[50]
B. subtilis ATCC 6633
E. coli ATCC 25922
L. Monocytogenes ATCC 1911
F. Oxysporum f. Sp. CubenseInactiveMIC > 128 µg/mL
Fusarium spp.
P. Litchii
C. Gloeosporioides
H. Undatus
127S. aureus ATCC6538AntibacterialMIC > 128 µg/mL[50]
B. subtilis ATCC 6633
E. coli ATCC 25922
L. Monocytogenes ATCC 1911
F. Oxysporum f. Sp. CubenseInactiveMIC > 128 µg/mL
Fusarium spp.
P. Litchii
C. Gloeosporioides
H. Undatus
128S. aureus ATCC6538AntibacterialMIC > 128 µg/mL[50]
B. subtilis ATCC 6633
E. coli ATCC 25922
L. Monocytogenes ATCC 1911
F. Oxysporum f. Sp. CubenseInactiveMIC > 128 µg/mL
Fusarium spp.
P. Litchii
C. Gloeosporioides
H. Undatus
129S. aureus ATCC6538AntibacterialMIC = 128 µg/mL[50]
B. subtilis ATCC 6633MIC > 128 µg/mL
E. coli ATCC 25922MIC > 128 µg/mL
L. Monocytogenes ATCC 1911MIC > 128 µg/mL
F. Oxysporum f. Sp. CubenseInactiveInactive against F. Oxysporum f. Sp. Cubense
Fusarium spp.Inactive against Fusarium spp.
P. LitchiiInactive against P. Litchii
C. GloeosporioidesInactive against C. Gloeosporioides
H. UndatusInactive against H. Undatus
130S. aureus ATCC6538AntibacterialMIC > 128 µg/mL[50]
B. subtilis ATCC 6633
E. coli ATCC 25922
L. Monocytogenes ATCC 1911
F. Oxysporum f. Sp. CubenseInactiveMIC > 128 µg/mL
Fusarium spp.
P. Litchii
C. Gloeosporioides
H. Undatus
131S. aureus ATCC6538AntibacterialMIC > 128 µg/mL[50]
B. subtilis ATCC 6633
E. coli ATCC 25922
L. Monocytogenes ATCC 1911
F. Oxysporum f. Sp. CubenseInactiveMIC > 128 µg/mL
Fusarium spp.
P. Litchii
C. Gloeosporioides
H. Undatus
132S. aureus ATCC6538AntibacterialMIC > 128 µg/mL[50]
B. subtilis ATCC 6633
E. coli ATCC 25922
L. Monocytogenes ATCC 1911
F. Oxysporum f. Sp. CubenseInactiveMIC > 128 µg/mL
Fusarium spp.
P. Litchii
C. Gloeosporioides
H. Undatus
133MRSAAntibacterialMIC = 25 µg/mL[20]
B. subtilisMIC = 6.25 µg/mL
P. AeruginosaMIC > 50 µg/mL
S. TyphimuriumMIC > 50 µg/mL
P. ItalicumAntifungalMIC = 6.25 µg/mL
C. Gloeosporioides
134MRSAAntibacterialMIC = 25 µg/mL[20]
B. subtilisMIC = 25 µg/mL
P. AeruginosaMIC = 25 µg/mL
S. TyphimuriumMIC > 50 µg/mL
P. ItalicumAntifungalMIC = 50 µg/mL
C. GloeosporioidesMIC > 50 µg/mL
135MRSAAntibacterialMIC > 50 µg/mL[20]
B. subtilis
P. Aeruginosa
S. Typhimurium
P. ItalicumAntifungalMIC = 50 µg/mL
C. GloeosporioidesMIC > 50 µg/mL
136B. subtilis KCTC 1021AntibacterialMIC = 16 µg/mL[51]
K. Rhizophila KCTC 1915MIC = 32 µg/mL
S. aureus KCTC 1927MIC = 16 µg/mL
E. coli KCTC 2441MIC > 128 µg/mL
S. Typhimurium KCTC 2515MIC > 128 µg/mL
K. Pneumonia KCTC 2690MIC > 128 µg/mL
137B. subtilis KCTC 1021AntibacterialMIC = 64 µg/mL[51]
K. Rhizophila KCTC 1915MIC > 128 µg/mL
S. aureus KCTC 1927MIC > 128 µg/mL
E. coli KCTC 2441MIC > 128 µg/mL
S. Typhimurium KCTC 2515MIC > 128 µg/mL
K. Pneumonia KCTC 2690MIC > 128 µg/mL
138B. subtilis KCTC 1021AntibacterialMIC = 1 µg/mL[51]
K. Rhizophila KCTC 1915MIC = 2 µg/mL
S. aureus KCTC 1927MIC = 2 µg/mL
E. coli KCTC 2441MIC > 128 µg/mL
S. Typhimurium KCTC 2515MIC > 128 µg/mL
K. Pneumonia KCTC 2690MIC > 128 µg/mL
139B. subtilis KCTC 1021AntibacterialMIC > 128 µg/mL[51]
K. Rhizophila KCTC 1915
S. aureus KCTC 1927
E. coli KCTC 2441
S. Typhimurium KCTC 2515
K. Pneumonia KCTC 2690
140H. pylori ATCC43504AntibacterialMIC = 2 µg/mL[52]
H. pylori G27MIC = 1 µg/mL
H. pylori Hp159MIC = 1 µg/mL
H. pylori BY583MIC = 4 µg/mL
S. aureus ATCC25923MIC = 16 µg/mL
S. aureus USA300MIC = 2 µg/mL
S. aureus BKS231MIC = 4 µg/mL
S. aureus BKS233MIC = 8 µg/mL
141E. coliAntibacterialMIC = 8 µg/mL[53]
M. LuteusMIC = 8 µg/mL
P. AeruginosaMIC = 16 µg/mL
R. SolanacearumMIC = 8 µg/mL
A. AlternataAntifungalMIC > 64 µg/mL
B. CinereaMIC = 32 µg/mL
F. OxysporumMIC > 64 µg/mL
P. DigitatumMIC = 32 µg/mL
V. MaliMIC = 16 µg/mL
“-” indicates that the MIC value is beyond the detectable range.
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

Liu, X.; Xin, J.; Sun, Y.; Zhao, F.; Niu, C.; Liu, S. Terpenoids from Marine Sources: A Promising Avenue for New Antimicrobial Drugs. Mar. Drugs 2024, 22, 347. https://doi.org/10.3390/md22080347

AMA Style

Liu X, Xin J, Sun Y, Zhao F, Niu C, Liu S. Terpenoids from Marine Sources: A Promising Avenue for New Antimicrobial Drugs. Marine Drugs. 2024; 22(8):347. https://doi.org/10.3390/md22080347

Chicago/Turabian Style

Liu, Xiao, Jianzeng Xin, Yupei Sun, Feng Zhao, Changshan Niu, and Sheng Liu. 2024. "Terpenoids from Marine Sources: A Promising Avenue for New Antimicrobial Drugs" Marine Drugs 22, no. 8: 347. https://doi.org/10.3390/md22080347

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

Article Metrics

Back to TopTop