Next Article in Journal
Physico-Chemical Conversion of Lignocellulose: Inhibitor Effects and Detoxification Strategies: A Mini Review
Next Article in Special Issue
Enantiomeric Mixtures in Natural Product Chemistry: Separation and Absolute Configuration Assignment
Previous Article in Journal
Synthesis and Biological Evaluations of NO-Donating Oxa- and Aza-Pentacycloundecane Derivatives as Potential Neuroprotective Candidates
Previous Article in Special Issue
Chiral Drug Analysis in Forensic Chemistry: An Overview
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Marine Natural Peptides: Determination of Absolute Configuration Using Liquid Chromatography Methods and Evaluation of Bioactivities

by
Ye’ Zaw Phyo
1,2,†,
João Ribeiro
3,†,
Carla Fernandes
2,3,*,
Anake Kijjoa
1,2,* and
Madalena M. M. Pinto
2,3
1
ICBAS-Instituto de Ciências Biomédicas Abel Salazar, Universidade do Porto, Rua de Jorge Viterbo Ferreira, 228, 4050-313 Porto, Portugal
2
Interdisciplinary Centre of Marine and Environmental Research (CIIMAR), Edifício do Terminal de Cruzeiros do Porto de Leixões, Av. General Norton de Matos s/n, 4050-208 Matosinhos, Portugal
3
Laboratório de Química Orgânica e Farmacêutica, Departamento de Ciências Químicas, Faculdade de Farmácia da Universidade do Porto, Rua de Jorge Viterbo Ferreira, 228, 4050-313 Porto, Portugal
*
Authors to whom correspondence should be addressed.
These authors equally contributed to this work.
Molecules 2018, 23(2), 306; https://doi.org/10.3390/molecules23020306
Submission received: 30 December 2017 / Revised: 22 January 2018 / Accepted: 27 January 2018 / Published: 31 January 2018
(This article belongs to the Special Issue Chirality in Health and Environment: Recent developments)

Abstract

:
Over the last decades, many naturally occurring peptides have attracted the attention of medicinal chemists due to their promising applicability as pharmaceuticals or as models for drugs used in therapeutics. Marine peptides are chiral molecules comprising different amino acid residues. Therefore, it is essential to establish the configuration of the stereogenic carbon of their amino acid constituents for a total characterization and further synthesis to obtain higher amount of the bioactive marine peptides or as a basis for structural modifications for more potent derivatives. Moreover, it is also a crucial issue taking into account the mechanisms of molecular recognition and the influence of molecular three-dimensionality in this process. In this review, a literature survey covering the report on the determination of absolute configuration of the amino acid residues of diverse marine peptides by chromatographic methodologies is presented. A brief summary of their biological activities was also included emphasizing to the most promising marine peptides. A case study describing an experience of our group was also included.

1. Introduction

In recent years, it has become well known that the oceans represent a rich source of structurally unique bioactive compounds from the perspective of potential therapeutic agents [1,2]. Bioactive compounds can be isolated from a myriad of marine invertebrates such as mollusks, sponges, tunicates and bryozoans, in addition to algae and marine microorganisms, especially cyanobacteria, bacteria and fungi [3,4,5].
Over the last decades, novel bioactive compounds from marine organisms with important bioactivities, such as antifungal, antibacterial, cytotoxic and anti-inflammatory properties, have been widely explored, and many of them are considered as lead compounds for drug discovery as well as biologically useful agents in pharmaceutical research [6,7,8,9,10]. In fact, owing to their pharmacological potential, either directly as drugs or as models for molecular modifications and/or total synthesis, marine natural products are certainly an interesting source, exploited by many researchers [11].
Ziconotide (Prialt®), a peptide first isolated from the venom of the cone snail (Conus magus), and trabectedin (Yondelis®), an alkaloid originally isolated from a marine tunicate Ectenascidia turbinata and now obtained by semisynthesis, are examples of marine natural products that have already been approved as human therapeutics [3,12,13,14]. Ziconotide is an analgesic used for treatment of patients suffering from chronic pain, and trabectedin for the treatment of soft tissue sarcomas and ovarian cancer.
In terms of the overall number of marine natural products, peptides are one of the most described due to their novel chemistry and diverse biological properties [15]. Actually, marine peptides are known to exhibit various biological activities such as antiviral, antiproliferative, antioxidant, anticancer, antidiabetic, anti-obesity, anticoagulant, antihypertensive, and calcium-binding activities [6,15,16,17].
Marine peptides are chiral molecules comprising different amino acid residue subunits. For their total characterization, and taking into account the mechanisms of molecular recognition and the influence of molecular three-dimensionality in this process, it is essential to define the configuration of the amino acids components of the peptide fractions, isolated from marine sources. Besides, it is also crucial to obtain the bioactive marine peptides by synthesis in order to achieve higher amount of compound for future assays or as a basis for structural modifications to obtain more potent derivatives.
Nowadays, there are different methodologies for the determination of the absolute configuration of amino acids, such as X-ray crystallography, NMR techniques, vibrational circular dichroism (VCD), enantioselective chromatography, optical rotatory dispersion (ORD), among others [18,19,20,21,22,23,24,25,26].
For the determination of the absolute configuration of amino acid residues of marine peptides, separation methodologies by using Marfey’s method, chiral high performance liquid chromatography (HPLC) analysis or both have proved to be suitable and the most described, as will be shown in this review. Regardless of the method used, the evaluation of peptides stereochemistry is based on the determination of the amino acid composition in peptide hydrolysates. Two main steps are involved, specifically the total or partial hydrolysis of peptides to obtain amino acid residues followed by their analysis by comparison with appropriated standards [27] (Figure 1).
Marfey’s method was first reported by Marfey in 1984 [28]. After the acid hydrolysis of peptides, the amino acid residues are derivatized with chiral Marfey’s reagents such as 1-fluoro-2-4-dinitrophenyl-5-d,l-alanine amide (FDAA) or 1-fluoro-2-4-dinitrophenyl-5-d,l-leucine amide (FDLA). Subsequent analysis via reverse phase liquid chromatography (LC), using generally C18 columns, and by comparison the retention times of the derivatized amino acids with suitable standards, afforded the stereochemistry of the peptides [29,30,31]. This method is often used for determination of the absolute configuration of amino acids, mainly because it is a simple method, offering a better resolution when compared to chiral HPLC methodologies; furthermore, several derivatization agents, such as FDAA and FDLA, are commercially available. However, this methodology has some disadvantages, including low availability of some standards, and the possibility of occurring racemization of the analyte during the derivatization reaction, prior to the chromatographic analysis [30,31].
The chiral analysis by HPLC is based on a formation of transient diastereomeric complexes between the amino acids present in the hydrolysates and the chiral stationary phase (CSP) employed, being the less stable complex the first to elute [32]. There are several types of CSPs, such as polysaccharide-based, Pirkle-type, protein-based, macrocyclic antibiotic-based, crown ether-based, ligand exchange type, among others [33,34,35]; however, the last three types are the most used for the separation of primary amine-containing compounds and amino acids [36,37]. Chiral HPLC offers several advantages, when comparing to Marfey’s method, including the direct analysis of the amino acid hydrolysates without further derivatization; moreover, the analysis often provides quicker results. However, poor chemical sensitivity, low sample capacity, and low availability and expensiveness of commercial chiral columns are some of the disadvantages of chiral HPLC method [38].
A number of reviews on marine peptides have appeared in recent years, focusing mainly on their biological activities, applications and biosynthesis as well as isolation procedures [16,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57]. In this review, several works related to the methods used for determination of the absolute configuration of marine peptides by chromatographic methods are presented in different sections according to the source of the marine peptides. Diverse types of peptides such as cyclic peptides, cyclic depsipeptides and lipopeptides are reported. A literature survey covering all the reports on liquid chromatographic methods (Marfey’s method and chiral HPLC) is presented (from 1996 to 2017). Furthermore, a case study describing an experience of our group is included.

2. Peptides from Marine Cyanobacteria and Other Bacteria

Cyanobacteria (blue-green algae), the most ancient known microorganisms on Earth, are a rich source of novel secondary metabolites possessing a broad spectrum of biological activities including antitumor, antibacterial, anticoagulant, antifungal, antiviral, antimalarial, antiprotozoal, and anti-inflammatory activities [58]. Currently, cyanobacteria are one of the most interesting sources of novel marine compounds [59]. Actually, the number of biologically active cyclic peptides, depsipeptides, lipopeptides, and other acyclic or small peptides, many of which containing unusual amino acid residues or modified amino acid units, is impressive. In addition to cyanobacteria, this type of compounds has also been isolated from other marine-derived bacteria.

2.1. Cyclic Peptides

Scattered publications concerning the stereochemistry determination of the amino acid residues of several cyclic peptides, isolated from marine cyanobacteria and other bacteria, were reported (Table 1). Marfey’s method, using FDAA as derivatization reagent, allowed the successful determination of the absolute configuration of the amino acid residues of cyclic peptides 14 (Figure 2).
For the new cyclic tetrapeptide 1 isolated from the bacterium Nocardiopsis sp. [60], the absolute configuration of all the amino acid residues was found to be L. Similarly, the absolute configuration of the amino acid residues of three novel anabaenopeptins labeled NZ825 (2), NZ841 (3), and NZ857 (4) [61], were successfully determined by Marfey’s method combined with HPLC.
However, as Marfey’s method was not accurate enough to determine the absolute configuration of all the amino acid residues of some cyclic peptides 516 (Figure 2), it was necessary to associate this method with chiral HPLC.
This strategy, i.e., using a ligand exchange-type CSP in chiral HPLC associated with Marfey’s method, was used for the determination of amino acids stereochemistry of several cyclic peptides, including aurilide B (5) and C (6), isolated from the cyanobacterium Lyngbya majuscula [62], urukthapelstatin A (7), isolated from a culture broth of thermoactinomycetaceae bacterium Mechercharimyces asporophorigenens YM11-542 [63], pompanoeptpins A (8) and B (9), isolated from the cyanobacterium Lyngbya confervoides [64], marthiapeptide A (10) isolated from the deep South China Sea-derived Marinactinospora thermotolerance SCSIO 00652 [65], norcardiamides A (11) and B (12), isolated from the marine-derived actinomycete Nocardiopsis sp. CNX037 [66], destomides B–D (1315), isolated from the deep South China Sea-derived Streptomyces scopuliridis SCSIO ZJ46 [67], and jandolide (16) isolated from the marine cyanobacterium Okeania sp. [68].
The cyclic peptides aurilides B (5) and C (6) were reported to have the in vitro cytotoxicity toward NCl-H460, human lung tumor, and neuro-2a mouse neuroblastoma cell lines, with lethal concentration 50 (LC50) values between 0.01 and 0.13 µM [62]. Aurilide B (5) was evaluated in the NCl 60 cell line panel and was found to exhibit a high level of cytotoxicity, particularly against leukemia, renal, and prostate cancer cell lines [62]. The cyclic peptide pompanopeptpin A (8) was shown to exhibit trypsin inhibitory activity with an IC50 value of 2.4 ± 0.4 µg/mL [64]. A polythiazole cyclopeptide, marthiapeptide A (10) showed antibacterial activity against a panel of Gram-positive bacteria with minimum inhibitory concentration (MIC) values ranging from 2.0 to 8.0 μg/mL, and strong cytotoxicity against a panel of human cancer cell lines with IC50 values ranging from 0.38 to 0.52 μM [65]. The cyclohexapeptide destomide B (13) also showed antimicrobial activity against Staphylococcus aureus ATCC 29213, Streptococcus pneumoniae NCTC 7466 and MRSE shhs-E1 with MIC values of 16.0, 12.5, 32.0 µg/mL, respectively [67]. A cyclic polyketide-peptide hybrid, janadolide (16) exhibited potent antitrypanosomal activity with an IC50 value of 47 nM [68].
Recently, the configuration of the amino acids of a cytotoxic cyanobactin, wewakazole B (17), isolated from the cyanobacterium Moorea producens (Figure 3), was determined using only chiral HPLC [69]. Two different types of CSPs, under reverse phase mode, were used to perform the analysis. A macrocyclic antibiotic-based CSP afforded the assignment of the l-configuration for its Ala, Phe, and Pro residues, while a ligand exchange type CSP clearly identified the presence of l-Ile, which could not be distinguished by the first CSP [69].

2.2. Cyclic Depsipeptides

As mentioned above, there are many publications describing the isolation and characterization, including the determination of the stereochemistry of their amino acids, of new cyclic depsipeptides from marine cyanobacteria and other bacteria (Table 2). However, contrary to cyclic peptides, several works reported the use of chiral HPLC as the only method for determination of the configuration of amino acids. Figure 4 shows the structure of cyclic depsipeptides 1846, isolated from marine cyanobacteria and other bacteria, whose stereochemistry of the amino acids was determined only by this method.
The ligand exchange-type CSPs were the most widely used by different research groups. Cai et al. employed a penicillamine ligand exchange-type CSP to determine the absolute configuration of the amino acids constituent of malevamide B (18) and C (19) isolated from the cyanobacterium Symploca laete-viridis [71]. Three different mobile phases in reverse phase elution mode were used. Nevertheless, the stereochemistry of Amha and Amoa residues present in both compounds were not determined [71]. The same CSP was employed to establish that all the amino acids of the cytotoxic depsipeptide lyngbyapeptin B (20) [72], tasipeptins A (21) and B (22) [73], wewakamide A (23) [74], cocosamide A (24) and B (25) [75], and the antiparasitic depsipeptides dudawelamides A–D (2629) [76], isolated from cyanobacteria Lyngbya majuscula, Symploca sp., Lyngbya semiplena, Lyngbya majuscula, and Moorea producens, respectively, has l-configuration. The only exception was for allo-Hiva amino acid of dudawelamide C (29), which has d-configuration [76]. The configuration of the amino acids of the cyclic depsipeptides pitipeptolides A (30) and B (31), isolated from cyanobacterium Lyngbya majuscula, was assigned to be L by a ligand exchange-type CSP comprising N,N-dioctyl-l-alanine as chiral selector (Chiralpack MA (+) from Daicel) and different proportion of CuSO4:ACN as mobile phase [77]. By using the same CSP, the absolute configuration of three new cyclic depsipeptides, kohamamides A–C (3234) were also successfully established [78].
Zhou et al. [79] described the determination of the absolute configuration of new anti-infective cycloheptadepsipeptides marformycins A–F (3540), produced by the deep sea-derived Streptomyces drozdowiczii SCSIO 1014, using a ligand exchange type CSP containing the same chiral selector as the previous ones (N,N-dioctyl-l(or d)-alanine) but purchased from Mitsubishi Chemical Corporation (MCI GEL CRS10W). Another type of CSP, specifically the macrocyclic antibiotic-based Chirobiotic TAG, confirmed the presence of l-Pro and l-Val in an unusual cyclic depsipeptide, pitiprolamide (41), isolated from Lyngbya majuscula [80]. Interestingly, in some works, more than one CSP were employed to elucidate the configuration of all the amino acids contained in the hydrolysates of cyclic depsipeptides. For example, two different types of ligand exchange type CSPs were used to elucidate the stereochemistry of the amino acid residues of palau’amide (42), depsipeptide with strong cytotoxicity against KB cell line (IC50 value of 13 nM) [81].
In the case of pitipeptolides C–F (4346), which were isolated from the cyanobacterium Lyngbya majuscula, the configuration of most of the amino acid residues was determined using the macrocyclic antibiotic-based Chirobiotic TAG under reverse phase elution conditions [82]. Then, the N,N-dioctyl-l-alanine ligand exchange CSP Chiralpack MA (+), under the same elution mode, was used for the assignment of S configuration for Hiva residue [82].
The concurrent applicability of chiral HPLC and Marfey’s methods for determination of the absolute configuration of all the amino acid residues of cyclic depsipeptides 4778 (Figure 5) was also described in several reports, among which ten described the use of ligand exchange-type CSPs to perform the analysis in association with Marfey’s method [71,72,74,75,76,78,79,81]. Furthermore, the use of macrocyclic antibiotic-based CSPs was reported by Montaser et al. [82].
Considering the biological activities of cyclic depsipeptides, whose stereochemistry of the amino acids was determined by a combination of Marfey’s method and chiral HPLC, it is worth mentioning the following compounds. Ulongapeptin (47), isolated from a Palauan marine cyanobacterium Lyngbya sp. displayed significant cytotoxic activity against KB cells with IC50 value of 0.63 µM [83]. Largamides A–H (4855), isolated from the marine cyanobacterium Oscillatoria sp., inhibited chymotrypsin with IC50 values ranging from 4 to 25 µM [84]. Symplocamide A (60), isolated from the marine cyanobacterium Symploca sp., showed cytotoxicity against NCI-460, non-small cell lung cancer cells (IC50 = 40 nM), and neuro-2a mouse neuroblastoma cells (IC50 = 29 nM). It was also reported that 60 was active against three tropical parasites: malaria (Plasmodium falciparum, IC50 = 0.95 µM), chagas disease, (Trypanasoma cruzi, IC50 > 9.5 µM), and leishmaniasis (Leishmania donovani, IC50 > 9.5 µM) [87]. It was found that, kempopeptins A (61) and B (62), isolated from the marine cyanobacterium Lyngbya sp., exhibited inhibitory activity against elastase and chymotrypsin with IC50 values of 0.32 µM and 2.6 µM, respectively [88]. Palmyramide A (67), isolated from the marine cyanobacterium Lyngbya majuscula, showed sodium channel blocking activity in the neuro-2a cells as well as cytotoxic activity in H-460 human lung carcinoma cell line [91]. Companeramides A (77) and B (78), isolated from a marine cyanobacterial assemblage comprising a small filament Leptolyngbya species, showed high nanomolar in vitro antiplasmodial activity against Plasmodium falciparum strains D6, Dd2, and 7G8 [94].
Moreover, HPLC analysis after derivatization with a Marfey’s reagent has been reported as the only method to determine the stereochemistry of the amino acid residues of cyclic depsipeptides 7994 (Figure 6). FDAA was used as derivatization reagent for piperazimycins A–C (7981), cyclic hexadepsipeptides isolated from the fermentation broth of a marine-derived bacterium Streptomyces sp. Strain, collected from a sediment [95], grassypeptolides D (82) and E (83), cyclic depsipeptides isolated from the marine cyanobacterium Leptolyngbya sp. [96], fijimycins A–C (8486), cyclic depsipeptides isolated from a marine bacteria Streptomyces sp. [97]. The Marfey’s reagent FDLA was employed for the assignment of the absolute configuration of the amino acid residues of several cyclic depsipeptides such as itralamide A (87) and B (88) and carriebowmide sulfone (89), isolated from the marine cyanobacterium Lyngbya majuscula [98], viequeamide A (90), isolated from the marine button cyanobacterium (Rivularia sp.) [99], ngercheumicins F–I (9194) [100].
Many cyclic depsipeptides whose stereochemistry of their amino acids was determined only by Marfey’s method, exhibited various interesting biological activities. Thus, piperazimycin A (79) was found to exhibit potent cytotoxicity against a panel of sixty cancer cell lines (mean values of growth inhibition (GI50) = 100 nM, and LC50 = 2 µM) [95]. While, grassypeptolides D (82) and E (83) exhibited significant cytotoxicity to HeLa (IC50 = 335 and 192 nM, respectively) and mouse neuro-2a blastoma (IC50 = 559 and 407 nM, respectively) cell lines [96], itralamide B (88) was active against HEK293 cells (IC50 value of 6 ± 1 µM) [98]. Fijimycins A–C (8486) exhibited strong growth inhibitory activity against three MRSA strains in a concentration range of 4–32 µg/mL−1 [97].

2.3. Lipopeptides

To the best of our knowledge, there are only two reports describing simultaneously the isolation and characterization of lipopeptides from marine cyanobacteria (Figure 7) as well as the stereochemistry determination of the amino acids present in their hydrolysates (Table 3).
The configuration of N-Me-Hph of the lipopeptide antillatoxin B (95), isolated from the cyanobacterium Lyngbya majuscula, was assigned as L using FDAA as Marfey’s derivatization reagent [101]. Compound 95 exhibited significant sodium channel activation (EC50 = 1.77 μM) and ichthyotoxicity (LC50 = 1 μM) [101]. The hydrolysates of lipopeptides lobocyclamides A–C (9698), isolated from the cyanobacterium Lyngbya confervoides, were analyzed by either direct chiral HPLC, using the d-penicillamine ligand exchange type CSP or by prior derivatization by Marfey’s method and reverse phase HPLC [102]. Both compounds displayed modest in vitro antifungal activity against a panel of Candida sp., including two fluconazole-resistant strains. Interestingly, synergistic antifungal activity was also observed [102].

3. Peptides from Marine-Derived Fungi

Marine fungi have been isolated from various marine sources like algae, marine invertebrates, sediment or water, mangroves and sponges. Most of the fungal species isolated from marine sponges are related to the genera Aspergillus and Penicillium [103]. Marine fungi are a rich source of structurally unique and biologically active compounds with a wide range of biological activities, such as antimalarial, anticancer, antifungal, antibacterial, cytotoxicity and among others [104]. More than 1000 compounds have been already isolated from marine derived fungi and among them around 150–200 new compounds were bioactive [103,104].

3.1. Cyclic Peptides

A large number of cyclic peptides have been isolated from marine-derived fungi (Figure 8) and Table 4 shows the marine fungal cyclic peptides whose stereochemistry of their amino acid residues were determined. To the best of our knowledge, only three reports described the use of FDAA and FDLA as Marfey’s derivatization reagents, specifically for analysis of the peptides 99112.
The cyclic peptide cyclo-(l-leucyl)-trans-4-hydroxyl-l-prolyl-d-leucyl-trans-4-hydroxy-l-proline) (99), isolated from the marine mangrove-derived fungi Phomopsis sp. K38, and Alternaria sp. E33, was found to exhibit antifungal activity, particularly the fungus Helminthosporium sativum. By using a combination of Marfey’s method and a reverse phase HPLC, the presence of 4-OH-l-Pro and both l- and d-Leu residues in its structure was confirmed [105]. Scytalidamides A (100) and B (101), and clonostachysins A (102) and B (103), isolated from marine sponge derived fungus Clonostachys rogersoniana strain HJK9, were found to comprise l-configuration for all their amino acids [106,107]. Scytalidamides A (100) and B (101) showed cytotoxicity against human colon carcinoma tumor cell line (HCT-116) with IC50 values of 2.7 and 11.0 µM, respectively, and the NCI 60 cell-line, with 7.9 and 4.1 µM GI-50, respectively [106], while clonostachysins A (102) and B (103) exhibited inhibitory effect on the Prorocentrum micans alga at concentration higher than 30 µM [107].
Both Marfey’s method and chiral HPLC analysis were also used for the analysis of the absolute configuration of the amino acids of asperterrestide A (104), a cyclic peptide isolated from the marine-derived fungus Aspergillus terreus SCSGAF0162 which revealed the presence of d-Ala in its structure [108]. Nevertheless, it was not possible to distinguish between d-Ile and d-allo-Ile. Compound 104 showed promising inhibitory effects to the influenza virus strains A/WSN/33, and A/Hong Kong/8/68 (IC50 values of 15 and 8.1 µM, respectively) as well as cytotoxicity against U937 and MOLT4 cell lines (IC50 values of 6.5 and 6.2 µM, respectively) [108].
There are some reports describing the application of different types of CSPs, including crown ethers and macrocyclic antibiotics, for a chiral HPLC as the only method for analysis of the absolute configuration of the amino acids of peptides. Thus, the determination of the stereochemistry of the amino acids in the cyclic peptides sclerotides A (105) and B (106), isolated from the marine-derived fungus Aspergillus sclerotiorum PT06-1 [109], and cordyheptapeptides C–E (107109), isolated from the marine-derived fungus Acremonium persicinum SCSIO 115 [110], was achieved via chiral HPLC analysis of the hydrolysates using the crown ether-based CSP Crownpak CR (+). Sclerotides A (105) and B (106) were found to comprise l-Thr, l-Ala, d-Phe, and d-Ser [109]. Moreover, the presence of N-Me-d-Gly, and l-Val in cordyheptapeptides C (107) and D (108) and N-Me-l-Gly, N-Me-d-Tyr, and l-allo-Ile in cordyheptapeptide E (109) was confirmed, in addition to the present of other amino acids common to the three cyclic peptides [110]. Sclerotides A (105) and B (106) displayed antifungal activity against Candida albicans, with MIC values of 7.0 and 3.5 µM, respectively. Furthermore, sclerotide B (106) also exhibited cytotoxicity against HL-60 cell line as well as antibacterial activity against Pseudomonas aeruginosa [109] whereas cordyheptapeptides C (107) and E (109) exhibited cytotoxic activity against SF-268 (IC50 values of 3.7 and 3.2 µM, respectively), MCF-7 (IC50 values of 3.0 and 2.7 µM, respectively), and NCI-H460 (IC50 values of 11.6 and 4.5 µM, respectively) tumor cell lines [110]. Recently, the macrocyclic antibiotic-based CSP Chirobiotic T was employed in our group to determine the stereochemistry of amino acid residues of a new cyclic hexapeptide, similanamide (110), isolated from a marine sponge-associated fungus Aspergillus similanensis KUFA 0013 [111] which confirmed the presence of l-Ala, d-Leu, l-Val and d-pipecolic acid as its amino acids constituent. By using a similar approach, the absolute configuration of all the amino acids of two new cyclotetrapeptides, sartoryglabramides A (111) and B (112), isolated from the marine sponge-associated fungus Neosartorya glabra KUFA 0702, were assigned to be L-configuration in both cyclic peptides [112]. Further details are described in the case-study presented below.

3.2. Cyclic Depsipeptides

Most of the works describing the stereochemistry determination of amino acid residues of cyclic depsipeptides, isolated from marine fungus (Figure 9), employed Marfey’s method coupled with HPLC, using FDAA or FDLA as derivatization reagents (Table 5).
The structures of exumolides A (113) and B (114), cyclic depsipeptides isolated from the marine fungus of the genus Scytalidium, were confirmed to have l-Pro, l-Phe and N-Me-l-Leu in their composition [113]. Moreover, guangomide A (115), isolated from an unidentified sponge-derived fungus, was found to comprise N-Me-d-Phe [114]. The absolute configuration of common amino acid residues in destruxin E chlorohydrin (116) and pseudodestruxin C (117), isolated from the marine-derived fungus Beauveria felina, indicated the presence of N-Me-l-Ala and l-Ile in 116, l-Phe in 117, and N-Me-l-Val in both cyclic depsipeptides [115]. Furthermore, the absolute configuration of amino acid residues in zygosporamide (118), isolated from the marine-derived fungus Zygosporium masonii [116], petriellin A (119), isolated from the coprophilous fungus Petriella sordida [117], alternaramdie (120), isolated from the marine derived fungus Alternaria sp. SF-5016 [118], petrosifungins A (121) and B (122), isolated from a Penicillum brevicompac-tum strain of the Mediterranean sponge Petrosia ficiformis Poiret [119], were also successfully determined by Marfey’s method coupled with HPLC. Zygosporamide (118) displayed cytotoxic activity against RXF 393 and SF-268 cancer cell lines, with mean values of GI-50 of 6.0 and <5.6 nM, respectively [116] whereas guangomide A (115) [114] and alternaramdie (120) [118] showed antibacterial activity against Staphylococcus epidermidis and Staphylococcus aureus, respectively.
In the last few years, ultra-high-pressure liquid chromatography (UHPLC) is becoming an essential technique for ultra-fast separations, since it offers many benefits, including high efficiency in short analysis time and low solvent consumption [120,121]. Thus, the absolute configuration of the amino acid residues of oryzamides A–E (123127), isolated from the sponge-derived fungus Nigrospora oryzae PF18, was achieved by Marfey’s analysis with FDLA, combined with UHPLC [122].
Spicellamides A (128) and B (129), which were isolated from the marine-derived fungus Spicellum roseum, exhibited cytotoxicity against rat neuroblastoma B104 cell line, with an IC50 value of 6.2 µg/mL for spicellamide B (129) [123]. It is interesting to note that Marfey’s method was not suitable for the determination of the configuration of all amino acid residues of these two peptides. Therefore, a chiral HPLC approach was also employed, using a ligand exchange type CSP [123]. Furthermore, the chiral HPLC, using the crown ether-based CSP Crownpak CR (+), was used as the only method for determination of the configuration of the amino acids residues to confirm the presence of l-Tyr, l-Val, d-Leu, and (S)-O-Leu in the cyclic depsipeptides 1962 A (130) and B (131), isolated from the endophytic fungus Kandelia candel [124]. The cyclic depsipeptide 1962 A (130) exhibited growth inhibitory activity against the human breast cancer cell line, MCF-7, with IC50 of 100 µg/mL [124].

4. Peptides from Marine Sponges

Marine sponges are an important source of new metabolites from the marine environment [125]. They are considered one of the most prolific sources of novel bioactive compounds, such as terpenoids, alkaloids, macrolides, nucleoside derivatives, polyethers, fatty acids, sterols, peroxides and other numerous organic compounds [17,126]. In addition, cyclic peptides and depsipeptides have also been isolated from marine sponges. Most bioactive compounds from sponges displayed myriad of biological activities including anti-inflammatory, antibiotic, antitumor, antimalarial, antiviral, antifouling, and immuno- or neurosuppressive [127]. However, a significant number of marine natural products isolated from sponges were tested for the anticancer activity, and many of them were successfully undergoing to preclinical and clinical trials [126,128]. More recently, among bioactive compounds discovered from marine sponges, bioactive peptides have aroused attention of many researchers [8,17].

4.1. Cyclic Peptides

Several works reported the determination of the stereochemistry of the amino acid residues of diverse peptides isolated from marine sponges (Figure 10 and Figure 11), most of which described the application of Marfey’s method, using FDAA as the derivatization reagent (Table 6). By using Marfey’s method, Randazzo et al. [129] have showed that a 16-membered cyclic peptide, haliclamide (132), isolated from the Vanuatu marine sponge Haliclona sp., comprised the amino acid N-Me-l-Phe. The absolute configuration analysis of the amino acid residues of microsclerodermins J (133) and K (134), isolated from the sponge Microscleroderma herdmani, indicated, besides the amino acids common to both microsclerodermins, the presence of l-Phe, and l-Gly in 133, and l-Val, and l-Ala in 134 [130]. Moreover, in the case of euryjanicins E–G (135137), isolated from the Caribbean sponge Prosuberites laughlini [131], chujamide A (138), isolated from Suberites waedoensis [132], and kapakahines A–D (139142), isolated from Cribrochalina olemda [133], all the amino acid residues were proved to have the L configuration. However, except for d-Phe, all the amino acid residues of koshikamide B (143), isolated from the marine sponge Theonella sp., were shown to possess L-configuration [134]. Furthermore, perthamides C–F (144147), isolated from the sponge marine Theonella swinhoei, were found to comprise l-ThrOMe, and l-Phe; while perthamides C (144) and D (145) also comprise in their structures l-Asp, and (2R,3S)-βOHAsp [135,136]. Marfey’s method was also successfully used for evaluation of the stereochemistry of the amino acids of the cyclic peptides stylisins 1 (148) and 2 (149), stylissatins B–D (152154), and carteritins A (150) and B (151), isolated from marine sponge Stylissa sp. [137,138,139], as well as of callyaerin G (155), isolated from the marine sponge Callyspongia aerizusa [140].
The marine sponge cyclic peptides whose configuration of their amino acids constituent was determined by Marfey’s method, were found to display interesting biological activities. For examples, haliclamide (132) exhibited cytotoxicity against NSCLC-N6 cell line, with an IC50 value of 4.0 µg/mL [129], while koshikamide B (143) showed growth inhibitory activity against P388 and HCT-116 cell lines, with IC50 values of 0.45 and 7.5 µg/mL, respectively [134]. Callyaerin G (155) also exhibited cytotoxicity against mouse lymphoma cell line (L5178Y), and HeLa cell line, with ED50 values of 0.53 and 5.4 µg/mL, respectively [140]. Moreover, perthamides C (144), D (145) and F (147) showed anti-inflammatory activity, with perthamide F (147) having a promising antipsoriatic effect [135,136].
The simultaneous application of Marfey’s method, using FDAA as derivatization reagent, and chiral HPLC, using a ligand exchange type CSP, afforded the total assignment of the configuration of all the amino acid residues of reniochalistatins A–E (156160) [141]. Reniochalistatins A–E (156160), the cyclic peptides isolated from the marine sponge Reniochalina stalagmitis, were found to have all the amino acid residues with l configuration, including l-Asn and l-Trp in reniochalistatins A (156) and E (160) respectively [141]. The octapeptide reniochalistatin E (160) exhibited cytotoxicity towards myeloma RPMI-8226, and gastric MGC-803 cell lines (IC50 values of 4.9 and 9.7 µM, respectively) [141].
Phakellistatins 15–18 (161164) were analysed only by chiral HPLC, using the ligand exchange type Chirex 3126 d-penicillamine CSP, being able to identify that all the amino acids presented l-configuration. Furthermore, phakellistatins 15 (161) and 16 (162) exhibited cytotoxicity against P388 cancer cell line, with IC50 values of 8.5 and 5.4 µM, respectively, while phakellistatin 16 (162) was also active against BEL-7402 cancer cell line, with an IC50 value of 14.3 µM [142].

4.2. Cyclic Depsipeptides

A number of cyclic depsipeptides (Figure 12 and Figure 13), have been reported from marine sponges and Marfey’s method using FDAA as the derivatization reagent was the most used for the determination of absolute configuration of the amino acid residues. Table 7 gives some examples of the cyclic depsipeptides, isolated from marine sponges, whose stereochemistry of their amino acid residues was determined by Marfey’s method. By application of this method, callipeltins B (165) and C (166), isolated from the marine lithistida sponge Callipelta sp., were found to have in their structure l-Ala, N-Me-l-Ala, l-Leu, l-Thr and d-Arg [143]. For halipeptins A (167) and B (168), isolated from the marine sponge Haliclona sp., the referred method was only able to determine the configuration for l-Ala [144]. Marfey’s method was successfully used to determine the absolute configuration of the amino acid constituents of several marine sponge cyclic peptides including phoriospongin A (169) and B (170), isolated from the marine sponges Phoriospongia sp. and Callyspongia bilamellata [145], mirabamides A–D (171174), isolated from the marine sponge Siliquarias-pongia mirabilis [146], and neamphamides B–D (175177), isolated from the marine sponge Naemphius huxleyi [147]. Furthermore, the stereochemistry determination of amino acid residues in pipecolidepsins A (178) and B (179), isolated from the marine sponge Homophymia lamellose, confirmed the presence of several l and d amino acid residues, besides the (3S,4R) diMe-l-Glu and (2S,3S)-EtO-Asp present in both peptides [148]. Stellatolide A (180), a cyclic depsipeptide isolated from Ecionemia acervus, was found to have N-Me-d-Ser and d-allo-Thr, among other l-configured amino acids [149]. The amino acid constituents of the cyclic depsipeptides cyclolithistide A (181) and nagahamide A (182), both isolated from the sponge Theonella swinhoei, were all found to have the S or l-configuration, and the 3-amino-5-hydroxybenzoic acid (AHBA) residue in nagahamide A (182) was established to have 3S configuration [150,151].
Almost all the cyclic peptides isolated from marine sponges displayed a variety of biological activities. Thus, callipeltin C (166) [143], cyclolithistide A (181) [150], and mirabamides A–D (171174) [146] exhibited growth inhibitory activity against Candida albicans. Moreover, mirabamides A–D (171174) also exhibited potent anti-HIV activities towards several HIV strains [146] whereas neamphamides B–D (175177) displayed cytotoxic activity against several human cancer cell lines, including A549, HeLa, LNCaP, PC3, HEK, and NFF, with IC50 values ranging from 88 to 370 nM [147].
A simultaneous use of Marfey’s method and chiral HPLC analysis for stereochemical analysis of the amino acids of this type of peptides have been reported (Table 7). For examples, the absolute configuration of the amino acids of theopapuamides B (183) and C (184) and celebesides A–C (185187), isolated from an Indonesian sponge Siliquariaspongia mirabilis, was successful assigned by HPLC-MS analysis of FDAA derivatives as well as via chiral HPLC analysis using a ligand exchange type CSP [152]. In the case of theopapuamide (188), isolated from a papua new Guinea Lithistid Sponge Theonella swinhoei, Marfey’s method was used to confirm the presence of d-allo-Thr, whereas chiral HPLC using a ligand exchange type CSP, revealed the presence of N-Me-l-Leu, d-Asp, l-Leu and N-Me-l-Glu in its structure [153]. The absolute configuration of the amino acid residues of a new sulfated cyclic depsipeptide, mutremdamide A (189) and six new highly N-methylated peptides, koshikamides C–H (190195), isolated from different deep-water specimens of Theonella swinhoei and Theonella cupola, was also established by using both approaches. However, two different columns (C12 and C18) were used in Marfey’s method. By using chiral HPLC, it was possible to identify the amino acid residue N-Me-allo-l-Ile in koshikamide H (195) [154]. These cyclic peptides showed interesting biological activities. While theopapuamide (188) was cytotoxic against CEM-TART and HCT cell lines (IC50 values of 0.5 and 0.9 µM, respectively) [153], koshikamides F (193) and H (195) were active against a CCR5-using viral envelope, with IC50 values of 2.3 and 5.5 µM [154].

4.3. Lipopeptides

The absolute configuration of the amino acids of new N-sulfoureidylated lipopeptides sulfolipodiscamides A–C (196198), isolated from the n-butanol fraction of the marine sponge Discodermia kiiensis (Figure 14), was determined by Marfey’s method to be l-Uda and l-Gly (Table 8). Compound 196 was found to be cytotoxic against the murine leukemia cell line P388 with a IC50 value of 15 µM [155].

5. Peptides from Other Marine Invertebrates and Algae

A number of diverse bioactive peptides such as cyclic peptides, cyclic depsipeptides and linear peptides have been isolated from other marine invertebrates including ascidians, commonly called tunicates, mollusks, among others [17]. Moreover, the potential applications of many bioactive compounds from marine algae, mainly red and brown as well as some green algae, were reported [156].

5.1. Cyclic Peptides

To the best of our knowledge, only five works described the analysis of the stereochemistry of the cyclic peptides from marine invertebrates and algae (Figure 15). In all reported works, Marfey’s method was employed (Table 9). Among these, the determination of the absolute configuration of the cyclic hexapeptides didmolamides A (199) and B (200) and mollamides B (201) and C (202), isolated from the marine ascidian Didemnum molle from Madagascar and Indonesia, respectively, was performed by Marfey’s method using FDAA as derivatization reagent [157,158]. These compounds showed interesting biological activities, particularly, cytotoxicity against A549, HT29 MEL28 tumor cell lines, with IC50 values ranging from 10 to 20 µg/mL for didmolamides A (199) and B (200) [157] while 201 showed antimalarial activity against Plasmodium falciprum, clones D6 and W2, with IC50 values of 2.0 and 21 µg/mL, respectively [158].
Furthermore, the stereochemical determination of antatollamides A (203) and B (204), isolated from the marine ascidian Didemnum molle, sanguinamide A (205), isolated from the sea slug Hexabranchus sanguineus, and gamakamide E (206), isolated from the oysters Crassostrea giga, was carried out by Marfey’s method using FDLA as a derivatization reagent. The analysis demonstrated that most of their amino acids have the L-configuration, with the exception of d-Ala and d-Lys in antatollamides A (203) and B (204), and gamakamide E (206), respectively [159,160,161].

5.2. Cyclic Depsipeptides

To the best of our knowledge, only four works reported the determination of the stereochemistry of amino acid constituents of the cyclic depsipeptides from marine invertebrates and algae (Figure 16). Among these, three employed only Marfey’s method, specifically for peptides 207216. However, for peptide 217, Marfey’s method was not efficient and, as a consequence, a ligand exchange type CSP was also used for complete determination of the configuration of its amino acids (Table 10).
The determination of the absolute configuration of the amino acids in kahalalides A–F (207212), isolated from the marine mollusk Elysia rufescens, was performed by using FDLA as the derivatization reagent and the presence of diverse residues of l- and d-Val in these peptides was confirmed [162]. Using FDAA as the Marfey derivatization reagent, the absolute configuration of tamandarins A (213) and B (214), isolated from an unidentified Brazilian marine ascidian of the family Didemnidae [163], and kahalalides P (215) and Q (216), isolated from green algae Bryopsis species [164] were elucidated. In the case of kahalalide O (217), the absolute configuration of its amino acid constituents was determined by Marfey’s method and chiral HPLC analysis, using a ligand exchange type CSP [165]. Tamandarin A (213) was found to display cytotoxicity against BX-PC3, DU-145, and UMSCC10b human cancer cell lines, with IC50 values of 1.79, 1.36, and 0.99 µg/mL, respectively [163].

5.3. Lipopeptides

For lipopeptides isolated from other marine invertebrates and algae, there are only two works which reported the use of a chiral HPLC for the stereochemistry determination of the amino acid residues (Table 11) of the peptides 218221 (Figure 17).
Chiral HPLC analysis by using a ligand exchange type CSP (Phenomenex Chirex Phase 3126) was used to determine the configuration of the amino acid residues in eudistomides A (218) and B (219), isolated from an ascidian Eudistoma sp. It was possible to verify the presence of l-Pro, l-Ala and l-Leu in both compounds as well as the presence of l-Cyp in eudistomide A (218) [166]. Similarly, a chiral HPLC analysis using a ligand exchange type CSP (CHIRALPAK (MA+)) was able to confirm the presence of four l-amino acid residues and d-Ala, d-Phe, and d-Ser in mebamamides A (220) and B (221), isolated from the green alga Derbesia marina [167].

6. Case-Study: Chiral HPLC in the Analysis of the Stereochemistry of Cyclopeptides Isolated from Marine Sponge-Associated Fungi

Recently, the determination of the stereochemistry of the amino acid residues of three bioactive marine natural products, by chiral HPLC analysis of their acidic hydrolysates, using appropriate d- and l-amino acid standards was achieved in our group [111,112]. The marine sponge-associated fungus Aspergillus similanensis KUFA 0013 was the source of the cyclohexapeptide similanamide (110) (Figure 8), while cyclotetrapeptides sartoryglabramides A (111) and B (112) (Figure 8) were isolated from the marine sponge-associated fungus Neosartorya glabra KUFA 0702. The enantioseparations of the amino acids were successfully performed on Chirobiotic T column under reverse phase elution conditions. Actually, the teicoplanin selector of this column has several characteristic features that make it suitable for amino acid analysis [168,169]. Figure 18 shows selected chromatograms of the enantioseparation of standard amino acids.
The elution order of all the standard enantiomers of amino acids was confirmed by injecting solutions of the racemic or enantiomeric mixtures of amino acids and then each enantiomer separately. As an example, Figure 19 shows the chromatograms obtained during the method development for the determination of the elution order of Ala. As expected, the d-enantiomer was always more strongly retained than the corresponding l-enantiomer on Chirobiotic T column [168]. Mixed HPLC analyses of the acidic hydrolysates with appropriate standard amino acids (co-injection) (Table 12), confirmed the stereochemistry of the amino acids of the three cyclopeptides [111,112]. Chiral HPLC technique demonstrated to be decisive leading to the unambiguous elucidation of the amino acid constituents of the three marine natural products.
Additionally, the in vitro growth inhibitory activity against MCF-7, breast adenocarcinoma, NCI-H460, non-small cell lung cancer and A373, melanoma, cell lines, as well as antibacterial activity against reference strains and the environmental multidrug-resistant isolates (MRS and VRE) were evaluated for cyclopeptide 110. Only weak activity against the three cancer cell lines was observed [111]. Moreover, cyclopeptides 111 and 112 were tested for their antifungal activity against filamentous (Aspergillus fumigatus ATCC 46645), dermatophyte (Trichophyton rubrum ATCC FF5) and yeast (Candida albicans ATCC 10231), as well as for their antibacterial activity against Gram-positive (Escherichia coli ATCC 25922) and Gram-negative (Staphyllococus aureus ATCC 25923) bacteria. None of them exhibited antibacterial or antifungal activities [112].

7. Conclusions

In summary, concerning all the reported studies surveyed in this review, which are related to the determination of the absolute configuration of the marine peptides, their distribution according to the methods used, is shown in Figure 20. It is possible to conclude that Marfey’s method is the most employed accounting for 52% of the reported studies, while only 21% of the studies described the use of chiral HPLC analysis. Moreover, 27% of the studies included the application of both methods. In fact, in some cases, the complementarity of both methods demonstrated to be crucial for the stereochemical analysis of all the amino acid residues.
Figure 21 compares the reported studies before and after 2007. Interestingly, it is possible to observe that in the last ten years, Marfey’s method is still the most used for determination of the absolute configuration of amino acid residues in marine peptides. However, it is important to point out a notable increase of the number of studies related to a chiral HPLC analysis, either as the only method or in a combination with Marfey’s method.
In our opinion, the current trend is to use chiral HPLC for stereochemical analysis due to many advantages of this method. For examples, there is no need for prior derivatization, it requires much less sample manipulation and the results are more rapid to obtain. In contrast, Marfey’s method involves time-consuming and labor-intensive procedure.
We believe that the reasons that can justify the actual low number of studies using chiral HPLC is due to the price of the commercially available CSPs and the fact that there is no universal CSP, i.e., one CSP can only separate a limited number of chiral compounds and, in many cases, the choice of CSP may become a very difficult task.

Acknowledgments

This research was partially supported by the Strategic Funding UID/Multi/04423/2013 through national funds provided by FCT—Foundation for Science and Technology and European Regional Development Fund (ERDF), in the framework of the programme PT2020, the project PTDC/MAR-BIO/4694/2014 (reference POCI-01-0145-FEDER-016790; Project 3599—Promover a Produção Científica e Desenvolvimento Tecnológico e a Constituição de Redes Temáticas (3599-PPCDT)) as well as by the project INNOVMAR—Innovation and Sustainability in the Management and Exploitation of Marine Resources (reference NORTE-01-0145-FEDER-000035, within Research Line NOVELMAR), supported by North Portugal Regional Operational Programme (NORTE 2020), under the PORTUGAL 2020 Partnership Agreement, through the European Regional Development Fund (ERDF), and Chiral_Drugs_CESPU_2017. Phyo thanks the Erasmus Mundus Action 2 (Lotus Plus project) for a PhD’s scholarship.

Author Contributions

Ye’ Zaw Phyo collected the primary data, analyzed the references, and compiled the draft manuscript. João Ribeiro contributed in writing of the manuscript and data analysis. Carla Fernandes, Anake Kijjoa and Madalena M. M. Pinto supervised development of the manuscript, and assisted in data interpretation, manuscript evaluation, and editing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pinto, M.M.M.; Castanheiro, R.A.P.; Kijjoa, A. Xanthones from marine-derived microorganisms: Isolation, structure elucidation and biological activities. In Encyclopedia of Analytical Chemistry; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2006. [Google Scholar]
  2. Blunt, J.W.; Copp, B.R.; Keyzers, R.A.; Munro, M.H.G.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2017, 34, 235–294. [Google Scholar] [CrossRef] [PubMed]
  3. Kijjoa, A.; Sawangwong, P. Drugs and cosmetics from the sea. Mar. Drugs 2004, 2, 73–82. [Google Scholar] [CrossRef]
  4. Saleem, M.; Ali, M.S.; Hussain, S.; Jabbar, A.; Ashraf, M.; Lee, Y.S. Marine natural products of fungal origin. Nat. Prod. Rep. 2007, 24, 1142–1152. [Google Scholar] [CrossRef] [PubMed]
  5. Xu, J. Biomolecules produced by mangrove-associated microbes. Curr. Med. Chem. 2011, 18, 5224–5266. [Google Scholar] [CrossRef] [PubMed]
  6. Cheung, R.C.; Ng, T.B.; Wong, J.H. Marine peptides: Bioactivities and applications. Mar. Drugs 2015, 13, 4006–4043. [Google Scholar] [CrossRef] [PubMed]
  7. Mayer, A.M.; Rodriguez, A.D.; Berlinck, R.G.; Fusetani, N. Marine pharmacology in 2007–8: Marine compounds with antibacterial, anticoagulant, antifungal, anti-inflammatory, antimalarial, antiprotozoal, antituberculosis, and antiviral activities; affecting the immune and nervous system, and other miscellaneous mechanisms of action. Comp. Biochem. Physiol. C 2011, 153, 191–222. [Google Scholar]
  8. Aneiros, A.; Garateix, A. Bioactive peptides from marine sources: Pharmacological properties and isolation procedures. J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 2004, 803, 41–53. [Google Scholar] [CrossRef] [PubMed]
  9. Donia, M.; Hamann, M.T. Marine natural products and their potential applications as anti-infective agents. Lancet Infect. Dis. 2003, 3, 338–348. [Google Scholar] [CrossRef]
  10. Harada, K.-I.; Fujii, K.; Mayumi, T.; Hibino, Y.; Suzuki, M.; Ikai, Y.; Oka, H. A method usingL/CMS for determination of absolute configuration of constituent amino acids in peptide—Advanced Marfey’s method. Tetrahedron Lett. 1995, 36, 1515–1518. [Google Scholar] [CrossRef]
  11. Blunt, J.W.; Copp, B.R.; Hu, W.P.; Munro, M.H.; Northcote, P.T.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2009, 26, 170–244. [Google Scholar] [CrossRef] [PubMed]
  12. Jimeno, J.; Faircloth, G.; Sousa-Faro, J.M.F.; Scheuer, P.; Rinehart, K. New marine derived anticancer therapeutics—A journey from the sea to clinical trials. Mar. Drugs 2004, 2, 14–29. [Google Scholar] [CrossRef]
  13. Baker, D.D.; Chu, M.; Oza, U.; Rajgarhia, V. The value of natural products to future pharmaceutical discovery. Nat. Prod. Rep. 2007, 24, 1225–1244. [Google Scholar] [CrossRef] [PubMed]
  14. Glaser, K.B.; Mayer, A.M. A renaissance in marine pharmacology: From preclinical curiosity to clinical reality. Biochem. Pharmacol. 2009, 78, 440–448. [Google Scholar] [CrossRef] [PubMed]
  15. Gogineni, V.; Hamann, M.T. Marine natural product peptides with therapeutic potential: Chemistry, biosynthesis, and pharmacology. Biochim. Biophys. Acta 2018, 1862, 81–196. [Google Scholar] [CrossRef] [PubMed]
  16. Anjum, K.; Abbas, S.Q.; Akhter, N.; Shagufta, B.I.; Shah, S.A.A.; Hassan, S.S.U. Emerging biopharmaceuticals from bioactive peptides derived from marine organisms. Chem. Biol. Drug Des. 2017, 90, 12–30. [Google Scholar] [CrossRef] [PubMed]
  17. Jo, C.; Khan, F.F.; Khan, M.I.; Iqbal, J. Marine bioactive peptides: Types, structures, and physiological functions. Food Rev. Int. 2016, 33, 44–61. [Google Scholar] [CrossRef]
  18. He, Y.; Wang, B.; Dukor, R.K.; Nafie, L.A. Determination of absolute configuration of chiral molecules using vibrational optical activity: A review. Appl. Spectrosc. 2011, 65, 699–723. [Google Scholar] [CrossRef] [PubMed]
  19. Seco, J.M.; Quiñoá, E.; Riguera, R. The assignment of absolute configuration by NMR†. Chem. Rev. 2004, 104, 17–118. [Google Scholar] [CrossRef]
  20. Hoye, T.R.; Jeffrey, C.S.; Shao, F. Mosher ester analysis for the determination of absolute configuration of stereogenic (chiral) carbinol carbons. Nat. Protoc. 2007, 2, 2451–2458. [Google Scholar] [CrossRef] [PubMed]
  21. Li, X.; Hopmann, K.H.; Hudecova, J.; Isaksson, J.; Novotna, J.; Stensen, W.; Andrushchenko, V.; Urbanova, M.; Svendsen, J.S.; Bour, P.; et al. Determination of absolute configuration and conformation of a cyclic dipeptide by NMR and chiral spectroscopic methods. J. Phys. Chem. A 2013, 117, 1721–1736. [Google Scholar] [CrossRef] [PubMed]
  22. Berova, N.; Di Bari, L.; Pescitelli, G. Application of electronic circular dichroism in configurational and conformational analysis of organic compounds. Chem. Soc. Rev. 2007, 36, 914–931. [Google Scholar] [CrossRef] [PubMed]
  23. Toome, V.; De Bernardo, S.; Weigele, M. A simple method for determining the absolute configuration of α-amino acids. Tetrahedron 1975, 31, 2625–2627. [Google Scholar] [CrossRef]
  24. Urriolabeitia, E.P.; Díaz-de-Villegas, M.D. A simple method for determining the absolute configuration of alpha-amino acids. J. Chem. Ed. 1999, 76, 77. [Google Scholar] [CrossRef]
  25. Gomez, E.D.; Duddeck, H. A new method proposed for the determination of absolute configurations of alpha-amino acids. Magn. Reson. Chem. 2009, 47, 222–227. [Google Scholar] [CrossRef] [PubMed]
  26. Calcaterra, A.; D’Acquarica, I. The market of chiral drugs: Chiral switches versus de novo enantiomerically pure compounds. J. Pharm. Biomed. Anal. 2018, 147, 323–340. [Google Scholar] [CrossRef] [PubMed]
  27. Dolowy, M.; Pyka, A. Application of TLC, HPLC and GC methods to the study of amino acid and peptide enantiomers: A review. Biomed. Chromatogr. 2014, 28, 84–101. [Google Scholar] [CrossRef] [PubMed]
  28. Marfey, P. Determination ofd-amino acids. II. Use of a bifunctional reagent, 1,5-difluoro-2,4-dinitrobenzene. Carlsberg Res. Commun. 1984, 49, 591–596. [Google Scholar] [CrossRef]
  29. Vijayasarathy, S.; Prasad, P.; Fremlin, L.J.; Ratnayake, R.; Salim, A.A.; Khalil, Z.; Capon, R.J. C3 and 2D C3 Marfey’s methods for amino acid analysis in natural products. J. Nat. Prod. 2016, 79, 421–427. [Google Scholar] [CrossRef] [PubMed]
  30. Bhushan, R.; Bruckner, H. Use of Marfey’s reagent and analogs for chiral amino acid analysis: Assessment and applications to natural products and biological systems. J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 2011, 879, 3148–3161. [Google Scholar] [CrossRef] [PubMed]
  31. Bhushan, R.; Bruckner, H. Marfey’s reagent for chiral amino acid analysis: A review. Amino Acids 2004, 27, 231–247. [Google Scholar] [CrossRef] [PubMed]
  32. Tiritan, M.E.; Ribeiro, A.R.; Fernandes, C.; Pinto, M. Chiral pharmaceuticals. In Kirk-Othmer Encyclopedia of Chemicl Technology; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2016. [Google Scholar]
  33. Okamoto, Y.; Ikai, T. Chiral HPLC for efficient resolution of enantiomers. Chem. Soc. Rev. 2008, 37, 2593–2608. [Google Scholar] [CrossRef] [PubMed]
  34. Fernandes, C.; Phyo, Y.Z.; Silva, A.S.; Tiritan, M.E.; Kijjoa, A.; Pinto, M.M.M. Chiral stationary phases based on small molecules: An update of the last 17 years. Sep. Purif. Methods 2017, 33, 1–35. [Google Scholar] [CrossRef]
  35. Ribeiro, J.; Tiritan, M.; Pinto, M.; Fernandes, C. Chiral stationary phases for liquid chromatography based on chitin- and chitosan-derived marine polysaccharides. Symmetry 2017, 9, 190. [Google Scholar] [CrossRef]
  36. Fernandes, C.; Tiritan, M.E.; Pinto, M. Small molecules as chromatographic tools for hplc enantiomeric resolution: Pirkle-type chiral stationary phases evolution. Chromatographia 2013, 76, 871–897. [Google Scholar] [CrossRef]
  37. Ribeiro, A.R.; Castro, P.M.L.; Tiritan, M.E. Chiral pharmaceuticals in the environment. Environ. Chem. Lett. 2012, 10, 239–253. [Google Scholar] [CrossRef]
  38. Li, B.; T Haynie, D. Chiral Drug Separation. In Encyclopedia of Chemical Processing; Taylor & Francis: Boca Raton, FL, USA, 2006. [Google Scholar]
  39. Auckloo, B.N.; Wu, B. Structure, biological properties and applications of marine-derived polysaccharides. Curr. Org. Chem. 2016, 20, 2002–2012. [Google Scholar] [CrossRef]
  40. Bennur, T.; Ravi Kumar, A.; Zinjarde, S.S.; Javdekar, V. Nocardiopsis species: A potential source of bioactive compounds. J. Appl. Microbiol. 2016, 120, 1–16. [Google Scholar] [CrossRef] [PubMed]
  41. Ciavatta, M.L.; Lefranc, F.; Carbone, M.; Mollo, E.; Gavagnin, M.; Betancourt, T.; Dasari, R.; Kornienko, A.; Kiss, R. Marine mollusk-derived agents with antiproliferative activity as promising anticancer agents to overcome chemotherapy resistance. Med. Res. Rev. 2017, 37, 702–801. [Google Scholar] [CrossRef] [PubMed]
  42. Fang, W.Y.; Dahiya, R.; Qin, H.L.; Mourya, R.; Maharaj, S. Natural proline-rich cyclopolypeptides from marine organisms: Chemistry, synthetic methodologies and biological status. Mar. Drugs 2016, 14. [Google Scholar] [CrossRef] [PubMed]
  43. Fernando, I.P.S.; Nah, J.W.; Jeon, Y.J. Potential anti-inflammatory natural products from marine algae. Environ. Toxicol. Pharmacol. 2016, 48, 22–30. [Google Scholar] [CrossRef] [PubMed]
  44. França, P.H.B.; da Silva-Júnior, E.F.; Santos, B.V.O.; Alexandre-Moreira, M.S.; Quintans-Júnior, L.J.; de Aquino, T.M.; de Araújo-Júnior, J.X. Antileishmanial marine compounds: A review. Rec. Nat. Prod. 2017, 11, 92–113. [Google Scholar]
  45. Hou, X.M.; Xu, R.F.; Gu, Y.C.; Wang, C.Y.; Shao, C.L. Biological and chemical diversity of coral-derived microorganisms. Curr. Med. Chem. 2015, 22, 3707–3762. [Google Scholar] [CrossRef] [PubMed]
  46. Ji, N.-Y.; Wang, B.-G. Mycochemistry of marine algicolous fungi. Fungal Divers. 2016, 80, 301–342. [Google Scholar] [CrossRef]
  47. Jin, Z. Muscarine, imidazole, oxazole and thiazole alkaloids. Nat. Prod. Rep. 2016, 33, 1268–1317. [Google Scholar] [CrossRef] [PubMed]
  48. Lu, W.; Yu, W.; Sun, D. Advances in application of marine bioactive peptides in drug development. Chin. J. Org. Chem. 2017, 37, 1681. [Google Scholar] [CrossRef]
  49. Maciel, E.; Costa Leal, M.; Lillebo, A.I.; Domingues, P.; Domingues, M.R.; Calado, R. Bioprospecting of marine macrophytes using MS-based lipidomics as a new approach. Mar. Drugs 2016, 14. [Google Scholar] [CrossRef] [PubMed]
  50. Oh, G.-W.; Ko, S.-C.; Lee, D.H.; Heo, S.-J.; Jung, W.-K. Biological activities and biomedical potential of sea cucumber (Stichopus japonicus): A review. Fish. Aquat. Sci. 2017, 20. [Google Scholar] [CrossRef]
  51. Pangestuti, R.; Kim, S.K. Bioactive peptide of marine origin for the prevention and treatment of non-communicable diseases. Mar. Drugs 2017, 15. [Google Scholar] [CrossRef] [PubMed]
  52. Rangel, M.; Santana, C.; Pinheiro, A.; Anjos, L.; Barth, T.; Júnior, O.; Fontes, W.; Castro, M. Marine depsipeptides as promising pharmacotherapeutic agents. Curr. Protein Pept. Sci. 2016, 18, 72–91. [Google Scholar] [CrossRef]
  53. Shannon, E.; Abu-Ghannam, N. Antibacterial derivatives of marine algae: An overview of pharmacological mechanisms and applications. Mar. Drugs 2016, 14. [Google Scholar] [CrossRef] [PubMed]
  54. Shukla, S. Therapeutic importance of peptides from marine source: A mini review. Indian J. Geo Mar. Sci. 2016, 45, 1422–1431. [Google Scholar]
  55. Sila, A.; Bougatef, A. Antioxidant peptides from marine by-products: Isolation, identification and application in food systems. A review. J. Funct. Foods 2016, 21, 10–26. [Google Scholar] [CrossRef]
  56. Suleria, H.A.R.; Gobe, G.; Masci, P.; Osborne, S.A. Marine bioactive compounds and health promoting perspectives; innovation pathways for drug discovery. Trends Food Sci. Technol. 2016, 50, 44–55. [Google Scholar] [CrossRef]
  57. Wang, L.; Dong, C.; Li, X.; Han, W.; Su, X. Anticancer potential of bioactive peptides from animal sources (Review). Oncol. Rep. 2017, 38, 637–651. [Google Scholar] [CrossRef] [PubMed]
  58. Burja, A.M.; Banaigs, B.; Abou-Mansour, E.; Grant Burgess, J.; Wright, P.C. Marine cyanobacteria—A prolific source of natural products. Tetrahedron 2001, 57, 9347–9377. [Google Scholar] [CrossRef]
  59. Blunt, J.W.; Copp, B.R.; Munro, M.H.G.; Northcote, P.T.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2006, 23, 26–78. [Google Scholar] [CrossRef] [PubMed]
  60. Shin, J.; Seo, Y.; Lee, H.S.; Rho, J.R.; Mo, S.J. A new cyclic peptide from a marine-derived bacterium of the genus Nocardiopsis. J. Nat. Prod. 2003, 66, 883–884. [Google Scholar] [CrossRef] [PubMed]
  61. Grach-Pogrebinsky, O.; Carmeli, S. Three novel anabaenopeptins from the cyanobacterium Anabaena sp. Tetrahedron 2008, 64, 10233–10238. [Google Scholar] [CrossRef]
  62. Han, B.; Gross, H.; Goeger, D.E.; Mooberry, S.L.; Gerwick, W.H. Aurilides B and C, cancer cell toxins from a Papua New Guinea collection of the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2006, 69, 572–575. [Google Scholar] [CrossRef] [PubMed]
  63. Matsuo, Y.; Kanoh, K.; Imagawa, H.; Adachi, K.; Nishizawa, M.; Shizuri, Y. Urukthapelstatin A, a novel cytotoxic substance from marine-derived Mechercharimyces asporophorigenens YM11–542. II. Physico-chemical properties and structural elucidation. J. Antibiot. 2007, 60, 256–260. [Google Scholar] [CrossRef] [PubMed]
  64. Matthew, S.; Ross, C.; Paul, V.J.; Luesch, H. Pompanopeptins A and B, new cyclic peptides from the marine cyanobacterium Lyngbya confervoides. Tetrahedron 2008, 64, 4081–4089. [Google Scholar] [CrossRef]
  65. Zhou, X.; Huang, H.; Chen, Y.; Tan, J.; Song, Y.; Zou, J.; Tian, X.; Hua, Y.; Ju, J. Marthiapeptide A, an anti-infective and cytotoxic polythiazole cyclopeptide from a 60 L scale fermentation of the deep sea-derived Marinactinospora thermotolerans SCSIO 00652. J. Nat. Prod. 2012, 75, 2251–2255. [Google Scholar] [CrossRef] [PubMed]
  66. Wu, Z.C.; Li, S.; Nam, S.J.; Liu, Z.; Zhang, C. Nocardiamides A and B, two cyclohexapeptides from the marine-derived actinomycete Nocardiopsis sp. CNX037. J. Nat. Prod. 2013, 76, 694–701. [Google Scholar] [CrossRef] [PubMed]
  67. Song, Y.; Li, Q.; Liu, X.; Chen, Y.; Zhang, Y.; Sun, A.; Zhang, W.; Zhang, J.; Ju, J. Cyclic Hexapeptides from the deep south china sea-derived streptomyces scopuliridis scsio zj46 active against pathogenic gram-positive bacteria. J. Nat. Prod. 2014, 77, 1937–1941. [Google Scholar] [CrossRef] [PubMed]
  68. Ogawa, H.; Iwasaki, A.; Sumimoto, S.; Kanamori, Y.; Ohno, O.; Iwatsuki, M.; Ishiyama, A.; Hokari, R.; Otoguro, K.; Omura, S.; et al. Janadolide, a cyclic polyketide-peptide hybrid possessing a tert-butyl group from an okeania sp. marine cyanobacterium. J. Nat. Prod. 2016, 79, 1862–1866. [Google Scholar] [CrossRef] [PubMed]
  69. Lopez, J.A.; Al-Lihaibi, S.S.; Alarif, W.M.; Abdel-Lateff, A.; Nogata, Y.; Washio, K.; Morikawa, M.; Okino, T. Wewakazole B, a cytotoxic cyanobactin from the cyanobacterium moorea producens collected in the red sea. J. Nat. Prod. 2016, 79, 1213–1218. [Google Scholar] [CrossRef] [PubMed]
  70. Davyt, D.; Serra, G. Thiazole and oxazole alkaloids: Isolation and synthesis. Mar. Drugs 2010, 8, 2755–2780. [Google Scholar] [CrossRef] [PubMed]
  71. Horgen, F.D.; Yoshida, W.Y.; Scheuer, P.J. Malevamides A–C, new depsipeptides from the marine cyanobacterium Symploca laete-viridis. J. Nat. Prod. 2000, 63, 461–467. [Google Scholar] [CrossRef] [PubMed]
  72. Luesch, H.; Yoshida, W.Y.; Moore, R.E.; Paul, V.J. Isolation and structure of the cytotoxin lyngbyabellin b and absolute configuration of lyngbyapeptin a from the marine cyanobacteriumlyngbya majuscula. J. Nat. Prod. 2000, 63, 1437–1439. [Google Scholar] [CrossRef] [PubMed]
  73. Williams, P.G.; Yoshida, W.Y.; Moore, R.E.; Paul, V.J. Tasipeptins A and B: New cytotoxic depsipeptides from the marine cyanobacterium Symploca sp. J. Nat. Prod. 2003, 66, 620–624. [Google Scholar] [CrossRef] [PubMed]
  74. Han, B.; Gross, H.; McPhail, K.L.; Goeger, D.; Maier, C.S.; Gerwick, W.H. Wewakamide A and guineamide G, cyclic depsipeptides from the marine cyanobacteria Lyngbya semiplena and Lyngbya majuscula. J. Microbiol. Biotechnol. 2011, 21, 930–936. [Google Scholar] [CrossRef] [PubMed]
  75. Gunasekera, S.P.; Owle, C.S.; Montaser, R.; Luesch, H.; Paul, V.J. Malyngamide 3 and cocosamides A and B from the marine cyanobacterium Lyngbya majuscula from Cocos Lagoon, Guam. J. Nat. Prod. 2011, 74, 871–876. [Google Scholar] [CrossRef] [PubMed]
  76. Almaliti, J.; Malloy, K.L.; Glukhov, E.; Spadafora, C.; Gutierrez, M.; Gerwick, W.H. Dudawalamides A-D, antiparasitic cyclic depsipeptides from the marine cyanobacterium moorea producens. J. Nat. Prod. 2017, 80, 1827–1836. [Google Scholar] [CrossRef] [PubMed]
  77. Luesch, H.; Pangilinan, R.; Yoshida, W.Y.; Moore, R.E.; Paul, V.J. Pitipeptolides A and B, new cyclodepsipeptides from the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2001, 64, 304–307. [Google Scholar] [CrossRef] [PubMed]
  78. Iwasaki, A.; Shiota, I.; Sumimoto, S.; Matsubara, T.; Sato, T.; Suenaga, K. Kohamamides A, B, and C, cyclic depsipeptides from an okeania sp. marine cyanobacterium. J. Nat. Prod. 2017, 80, 1948–1952. [Google Scholar] [CrossRef] [PubMed]
  79. Zhou, X.; Huang, H.; Li, J.; Song, Y.; Jiang, R.; Liu, J.; Zhang, S.; Hua, Y.; Ju, J. New anti-infective cycloheptadepsipeptide congeners and absolute stereochemistry from the deep sea-derived streptomyces drozdowiczii SCSIO 10141. Tetrahedron 2014, 70, 7795–7801. [Google Scholar] [CrossRef]
  80. Montaser, R.; Abboud, K.A.; Paul, V.J.; Luesch, H. Pitiprolamide, a proline-rich dolastatin 16 analogue from the marine cyanobacterium Lyngbya majuscula from Guam. J. Nat. Prod. 2011, 74, 109–112. [Google Scholar] [CrossRef] [PubMed]
  81. Williams, P.G.; Yoshida, W.Y.; Quon, M.K.; Moore, R.E.; Paul, V.J. The structure of Palau’amide, a potent cytotoxin from a species of the marine cyanobacterium Lyngbya. J. Nat. Prod. 2003, 66, 1545–1549. [Google Scholar] [CrossRef] [PubMed]
  82. Montaser, R.; Paul, V.J.; Luesch, H. Pitipeptolides C-F, antimycobacterial cyclodepsipeptides from the marine cyanobacterium Lyngbya majuscula from Guam. Phytochemistry 2011, 72, 2068–2074. [Google Scholar] [CrossRef] [PubMed]
  83. Williams, P.G.; Yoshida, W.Y.; Quon, M.K.; Moore, R.E.; Paul, V.J. Ulongapeptin, a cytotoxic cyclic depsipeptide from a Palauan marine cyanobacterium Lyngbya sp. J. Nat. Prod. 2003, 66, 651–654. [Google Scholar] [CrossRef] [PubMed]
  84. Plaza, A.; Bewley, C.A. Largamides A–H, unusual cyclic peptides from the marine cyanobacterium Oscillatoria sp. J. Org. Chem. 2006, 71, 6898–6907. [Google Scholar] [CrossRef] [PubMed]
  85. Bunyajetpong, S.; Yoshida, W.Y.; Sitachitta, N.; Kaya, K. Trungapeptins A–C, cyclodepsipeptides from the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2006, 69, 1539–1542. [Google Scholar] [CrossRef] [PubMed]
  86. Gunasekera, S.P.; Ritson-Williams, R.; Paul, V.J. Carriebowmide, a new cyclodepsipeptide from the marine cyanobacterium Lyngbya polychroa. J. Nat. Prod. 2008, 71, 2060–2063. [Google Scholar] [CrossRef] [PubMed]
  87. Linington, R.G.; Edwards, D.J.; Shuman, C.F.; McPhail, K.L.; Matainaho, T.; Gerwick, W.H. Symplocamide A, a potent cytotoxin and chymotrypsin inhibitor from the marine Cyanobacterium Symploca sp. J. Nat. Prod. 2008, 71, 22–27. [Google Scholar] [CrossRef] [PubMed]
  88. Taori, K.; Paul, V.J.; Luesch, H. Kempopeptins A and B, serine protease inhibitors with different selectivity profiles from a marine cyanobacterium, Lyngbya sp. J. Nat. Prod. 2008, 71, 1625–1629. [Google Scholar] [CrossRef] [PubMed]
  89. Matthew, S.; Paul, V.J.; Luesch, H. Tiglicamides A–C, cyclodepsipeptides from the marine cyanobacterium Lyngbya confervoides. Phytochemistry 2009, 70, 2058–2063. [Google Scholar] [CrossRef] [PubMed]
  90. Tripathi, A.; Puddick, J.; Prinsep, M.R.; Lee, P.P.; Tan, L.T. Hantupeptins B and C, cytotoxic cyclodepsipeptides from the marine cyanobacterium Lyngbya majuscula. Phytochemistry 2010, 71, 307–311. [Google Scholar] [CrossRef] [PubMed]
  91. Taniguchi, M.; Nunnery, J.K.; Engene, N.; Esquenazi, E.; Byrum, T.; Dorrestein, P.C.; Gerwick, W.H. Palmyramide A, a cyclic depsipeptide from a Palmyra Atoll collection of the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2010, 73, 393–398. [Google Scholar] [CrossRef] [PubMed]
  92. Salvador, L.A.; Biggs, J.S.; Paul, V.J.; Luesch, H. Veraguamides A-G, cyclic hexadepsipeptides from a dolastatin 16-producing cyanobacterium Symploca cf. hydnoides from Guam. J. Nat. Prod. 2011, 74, 917–927. [Google Scholar] [CrossRef] [PubMed]
  93. Meickle, T.; Gunasekera, S.P.; Liu, Y.; Luesch, H.; Paul, V.J. Porpoisamides A and B, two novel epimeric cyclic depsipeptides from a Florida Keys collection of Lyngbya sp. Bioorg. Med. Chem. 2011, 19, 6576–6580. [Google Scholar] [CrossRef] [PubMed]
  94. Vining, O.B.; Medina, R.A.; Mitchell, E.A.; Videau, P.; Li, D.; Serrill, J.D.; Kelly, J.X.; Gerwick, W.H.; Proteau, P.J.; Ishmael, J.E.; et al. Depsipeptide companeramides from a Panamanian marine cyanobacterium associated with the coibamide producer. J. Nat. Prod. 2015, 78, 413–420. [Google Scholar] [CrossRef] [PubMed]
  95. Miller, E.D.; Kauffman, C.A.; Jensen, P.R.; Fenical, W. Piperazimycins: Cytotoxic hexadepsipeptides from a marine-derived bacterium of the genus Streptomyces. J. Org. Chem. 2007, 72, 323–330. [Google Scholar] [CrossRef] [PubMed]
  96. Thornburg, C.C.; Thimmaiah, M.; Shaala, L.A.; Hau, A.M.; Malmo, J.M.; Ishmael, J.E.; Youssef, D.T.; McPhail, K.L. Cyclic depsipeptides, grassypeptolides D and E and Ibu-epidemethoxylyngbyastatin 3, from a Red Sea Leptolyngbya cyanobacterium. J. Nat. Prod. 2011, 74, 1677–1685. [Google Scholar] [CrossRef] [PubMed]
  97. Sun, P.; Maloney, K.N.; Nam, S.J.; Haste, N.M.; Raju, R.; Aalbersberg, W.; Jensen, P.R.; Nizet, V.; Hensler, M.E.; Fenical, W. Fijimycins A–C, three antibacterial etamycin-class depsipeptides from a marine-derived Streptomyces sp. Bioorg. Med. Chem. 2011, 19, 6557–6562. [Google Scholar] [CrossRef] [PubMed]
  98. Jimenez, J.I.; Vansach, T.; Yoshida, W.Y.; Sakamoto, B.; Porzgen, P.; Horgen, F.D. Halogenated fatty acid amides and cyclic depsipeptides from an eastern Caribbean collection of the cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2009, 72, 1573–1578. [Google Scholar] [CrossRef] [PubMed]
  99. Boudreau, P.D.; Byrum, T.; Liu, W.T.; Dorrestein, P.C.; Gerwick, W.H. Viequeamide A, a cytotoxic member of the kulolide superfamily of cyclic depsipeptides from a marine button cyanobacterium. J. Nat. Prod. 2012, 75, 1560–1570. [Google Scholar] [CrossRef] [PubMed]
  100. Kjaerulff, L.; Nielsen, A.; Mansson, M.; Gram, L.; Larsen, T.O.; Ingmer, H.; Gotfredsen, C.H. Identification of four new agr quorum sensing-interfering cyclodepsipeptides from a marine Photobacterium. Mar. Drugs 2013, 11, 5051–5062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Nogle, L.M.; Okino, T.; Gerwick, W.H. Antillatoxin B, a neurotoxic lipopeptide from the marine cyanobacterium Lyngbya majuscula. J. Nat. Prod. 2001, 64, 983–985. [Google Scholar] [CrossRef] [PubMed]
  102. MacMillan, J.B.; Ernst-Russell, M.A.; de Ropp, J.S.; Molinski, T.F. Lobocyclamides A–C, lipopeptides from a cryptic cyanobacterial mat containing Lyngbya confervoides. J. Org. Chem. 2002, 67, 8210–8215. [Google Scholar] [CrossRef] [PubMed]
  103. Rateb, M.E.; Ebel, R. Secondary metabolites of fungi from marine habitats. Nat. Prod. Rep. 2011, 28, 290–344. [Google Scholar] [CrossRef] [PubMed]
  104. Blunt, J.W.; Copp, B.R.; Keyzers, R.A.; Munro, M.H.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2014, 31, 160–258. [Google Scholar] [CrossRef] [PubMed]
  105. Li, C.; Wang, J.; Luo, C.; Ding, W.; Cox, D.G. A new cyclopeptide with antifungal activity from the co-culture broth of two marine mangrove fungi. Nat. Prod. Res. 2014, 28, 616–621. [Google Scholar] [CrossRef] [PubMed]
  106. Tan, L.T.; Cheng, X.C.; Jensen, P.R.; Fenical, W. Scytalidamides A and B, new cytotoxic cyclic heptapeptides from a marine fungus of the genus Scytalidium. J. Org. Chem. 2003, 68, 8767–8773. [Google Scholar] [CrossRef] [PubMed]
  107. Adachi, K.; Kanoh, K.; Wisespongp, P.; Nishijima, M.; Shizuri, Y. Clonostachysins A and B, new anti-dinoflagellate cyclic peptides from a marine-derived fungus. J. Antibiot. 2005, 58, 145–150. [Google Scholar] [CrossRef] [PubMed]
  108. 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]
  109. Zheng, J.; Zhu, H.; Hong, K.; Wang, Y.; Liu, P.; Wang, X.; Peng, X.; Zhu, W. Novel cyclic hexapeptides from marine-derived fungus, Aspergillus sclerotiorum PT06–1. Org. Lett. 2009, 11, 5262–5265. [Google Scholar] [CrossRef] [PubMed]
  110. Chen, Z.; Song, Y.; Chen, Y.; Huang, H.; Zhang, W.; Ju, J. Cyclic heptapeptides, cordyheptapeptides C-E, from the marine-derived fungus Acremonium persicinum SCSIO 115 and their cytotoxic activities. J. Nat. Prod. 2012, 75, 1215–1219. [Google Scholar] [CrossRef] [PubMed]
  111. Prompanya, C.; Fernandes, C.; Cravo, S.; Pinto, M.M.; Dethoup, T.; Silva, A.M.; Kijjoa, A. A new cyclic hexapeptide and a new isocoumarin derivative from the marine sponge-associated fungus Aspergillus similanensis KUFA 0013. Mar. Drugs 2015, 13, 1432–1450. [Google Scholar] [CrossRef] [PubMed]
  112. May Zin, W.W.; Buttachon, S.; Dethoup, T.; Fernandes, C.; Cravo, S.; Pinto, M.M.; Gales, L.; Pereira, J.A.; Silva, A.M.; Sekeroglu, N.; et al. New cyclotetrapeptides and a new diketopiperzine derivative from the marine sponge-associated fungus neosartorya glabra KUFA 0702. Mar. Drugs 2016, 14. [Google Scholar] [CrossRef] [PubMed]
  113. Jenkins, K.M.; Renner, M.K.; Jensen, P.R.; Fenical, W. Exumolides A and B: Antimicroalgal cyclic depsipeptides produced by a marine fungus of the genus Scytalidium. Tetrahedron Lett. 1998, 39, 2463–2466. [Google Scholar] [CrossRef]
  114. Amagata, T.; Morinaka, B.I.; Amagata, A.; Tenney, K.; Valeriote, F.A.; Lobkovsky, E.; Clardy, J.; Crews, P. A chemical study of cyclic depsipeptides produced by a sponge-derived fungus. J. Nat. Prod. 2006, 69, 1560–1565. [Google Scholar] [CrossRef] [PubMed]
  115. Lira, S.P.; Vita-Marques, A.M.; Seleghim, M.H.; Bugni, T.S.; LaBarbera, D.V.; Sette, L.D.; Sponchiado, S.R.; Ireland, C.M.; Berlinck, R.G. New destruxins from the marine-derived fungus Beauveria felina. J. Antibiot. 2006, 59, 553–563. [Google Scholar] [CrossRef] [PubMed]
  116. Oh, D.-C.; Jensen, P.R.; Fenical, W. Zygosporamide, a cytotoxic cyclic depsipeptide from the marine-derived fungus Zygosporium masonii. Tetrahedron Lett. 2006, 47, 8625–8628. [Google Scholar] [CrossRef]
  117. Aurelio, L.; Brownlee, R.T.C.; Dang, J.; Hughes, A.B.; Polya, G.M. Determination of the complete absolute configuration of petriellin A. Aust. J. Chem. 2006, 59, 407. [Google Scholar] [CrossRef]
  118. Kim, M.Y.; Sohn, J.H.; Ahn, J.S.; Oh, H. Alternaramide, a cyclic depsipeptide from the marine-derived fungus Alternaria sp. SF-5016. J. Nat. Prod. 2009, 72, 2065–2068. [Google Scholar] [CrossRef] [PubMed]
  119. Bringmann, G.; Lang, G.; Steffens, S.; Schaumann, K. Petrosifungins A and B, novel cyclodepsipeptides from a sponge-derived strain of Penicillium brevicompactum. J. Nat. Prod. 2004, 67, 311–315. [Google Scholar] [CrossRef] [PubMed]
  120. Kotoni, D.; Ciogli, A.; D’Acquarica, I.; Kocergin, J.; Szczerba, T.; Ritchie, H.; Villani, C.; Gasparrini, F. Enantioselective ultra-high and high performance liquid chromatography: A comparative study of columns based on the Whelk-O1 selector. J. Chromatogr. A 2012, 1269, 226–241. [Google Scholar] [CrossRef] [PubMed]
  121. Kotoni, D.; Ciogli, A.; Molinaro, C.; D’Acquarica, I.; Kocergin, J.; Szczerba, T.; Ritchie, H.; Villani, C.; Gasparrini, F. Introducing enantioselective ultrahigh-pressure liquid chromatography (eUHPLC): Theoretical inspections and ultrafast separations on a new sub-2-mum Whelk-O1 stationary phase. Anal. Chem. 2012, 84, 6805–6813. [Google Scholar] [CrossRef] [PubMed]
  122. Ding, L.J.; Yuan, W.; Liao, X.J.; Han, B.N.; Wang, S.P.; Li, Z.Y.; Xu, S.H.; Zhang, W.; Lin, H.W. Oryzamides A-E, cyclodepsipeptides from the sponge-derived fungus nigrospora oryzae PF18. J. Nat. Prod. 2016, 79, 2045–2052. [Google Scholar] [CrossRef] [PubMed]
  123. Kralj, A.; Kehraus, S.; Krick, A.; van Echten-Deckert, G.; Konig, G.M. Two new depsipeptides from the marine fungus Spicellum roseum. Planta Med. 2007, 73, 366–371. [Google Scholar] [CrossRef] [PubMed]
  124. Huang, H.; She, Z.; Lin, Y.; Vrijmoed, L.L.; Lin, W. Cyclic peptides from an endophytic fungus obtained from a mangrove leaf (Kandelia candel). J. Nat. Prod. 2007, 70, 1696–1699. [Google Scholar] [CrossRef] [PubMed]
  125. Kang, H.K.; Seo, C.H.; Park, Y. Marine peptides and their anti-infective activities. Mar. Drugs 2015, 13, 618–654. [Google Scholar] [CrossRef] [PubMed]
  126. Kamalakkannan, P. Marine sponges a good source of bioactive compounds in anticancer agents. Int. J. Pharm. Sci. Rev. Res. 2015, 31, 132–135. [Google Scholar]
  127. Muller, W.E.; Schroder, H.C.; Wiens, M.; Perovic-Ottstadt, S.; Batel, R.; Muller, I.M. Traditional and modern biomedical prospecting: Part II-the Benefits: Approaches for a sustainable exploitation of biodiversity (secondary metabolites and biomaterials from sponges). Evid. Based Complement. Alternat. Med. 2004, 1, 133–144. [Google Scholar] [CrossRef] [PubMed]
  128. Mehbub, M.F.; Lei, J.; Franco, C.; Zhang, W. Marine sponge derived natural products between 2001 and 2010: Trends and opportunities for discovery of bioactives. Mar. Drugs 2014, 12, 4539–4577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Randazzo, A.; Debitus, C.; Gomez-Paloma, L. Haliclamide, a novel cyclic metabolite from the Vanuatu marine sponge Haliclona sp. Tetrahedron 2001, 57, 4443–4446. [Google Scholar] [CrossRef]
  130. Zhang, X.; Jacob, M.R.; Rao, R.R.; Wang, Y.H.; Agarwal, A.K.; Newman, D.J.; Khan, I.A.; Clark, A.M.; Li, X.C. Antifungal cyclic peptides from the marine sponge Microscleroderma herdmani. Res. Rep. Med. Chem. 2012, 2, 7–14. [Google Scholar] [PubMed]
  131. Aviles, E.; Rodriguez, A.D. Euryjanicins E-G, poly-phenylalanine and poly-proline cyclic heptapeptides from the Caribbean sponge Prosuberites laughlini. Tetrahedron 2013, 69, 10797–10804. [Google Scholar] [CrossRef] [PubMed]
  132. Song, J.; Jeon, J.E.; Won, T.H.; Sim, C.J.; Oh, D.C.; Oh, K.B.; Shin, J. New cyclic cystine bridged peptides from the sponge Suberites waedoensis. Mar. Drugs 2014, 12, 2760–2770. [Google Scholar] [CrossRef] [PubMed]
  133. Yeung, B.K.; Nakao, Y.; Kinnel, R.B.; Carney, J.R.; Yoshida, W.Y.; Scheuer, P.J.; Kelly-Borges, M. The kapakahines, cyclic peptides from the marine sponge cribrochalina olemda. J. Org. Chem. 1996, 61, 7168–7173. [Google Scholar] [CrossRef] [PubMed]
  134. Araki, T.; Matsunaga, S.; Nakao, Y.; Furihata, K.; West, L.; Faulkner, D.J.; Fusetani, N. Koshikamide B, a cytotoxic peptide lactone from a marine sponge Theonella sp. J. Org. Chem. 2008, 73, 7889–7894. [Google Scholar] [CrossRef] [PubMed]
  135. Festa, C.; De Marino, S.; Sepe, V.; Monti, M.C.; Luciano, P.; D’Auria, M.V.; Débitus, C.; Bucci, M.; Vellecco, V.; Zampella, A. Perthamides C and D, two new potent anti-inflammatory cyclopeptides from a Solomon Lithistid sponge Theonella swinhoei. Tetrahedron 2009, 65, 10424–10429. [Google Scholar] [CrossRef]
  136. Festa, C.; De Marino, S.; Sepe, V.; D’Auria, M.V.; Bifulco, G.; Andrés, R.; Terencio, M.C.; Payá, M.; Debitus, C.; Zampella, A. Perthamides C–F, potent human antipsoriatic cyclopeptides. Tetrahedron 2011, 67, 7780–7786. [Google Scholar] [CrossRef]
  137. Mohammed, R.; Peng, J.; Kelly, M.; Hamann, M.T. Cyclic heptapeptides from the Jamaican sponge Stylissa caribica. J. Nat. Prod. 2006, 69, 1739–1744. [Google Scholar] [CrossRef] [PubMed]
  138. Sun, J.; Cheng, W.; de Voogd, N.J.; Proksch, P.; Lin, W. Stylissatins B–D, cycloheptapeptides from the marine sponge Stylissa massa. Tetrahedron Lett. 2016, 57, 4288–4292. [Google Scholar] [CrossRef]
  139. Afifi, A.H.; El-Desoky, A.H.; Kato, H.; Mangindaan, R.E.P.; de Voogd, N.J.; Ammar, N.M.; Hifnawy, M.S.; Tsukamoto, S. Carteritins A and B, cyclic heptapeptides from the marine sponge Stylissa carteri. Tetrahedron Lett. 2016, 57, 1285–1288. [Google Scholar] [CrossRef]
  140. Ibrahim, S.R.M.; Edrada-Ebel, R.; Mohamed, G.A.; Youssef, D.T.A.; Wray, V.; Proksch, P. Callyaerin G, a new cytotoxic cyclic peptide from the marine sponge Callyspongia aerizusa. Arkivoc 2008, 2008, 164. [Google Scholar] [CrossRef]
  141. Zhan, K.X.; Jiao, W.H.; Yang, F.; Li, J.; Wang, S.P.; Li, Y.S.; Han, B.N.; Lin, H.W. Reniochalistatins A-E, cyclic peptides from the marine sponge Reniochalina stalagmitis. J. Nat. Prod. 2014, 77, 2678–2684. [Google Scholar] [CrossRef] [PubMed]
  142. Zhang, H.J.; Yi, Y.H.; Yang, G.J.; Hu, M.Y.; Cao, G.D.; Yang, F.; Lin, H.W. Proline-containing cyclopeptides from the marine sponge Phakellia fusca. J. Nat. Prod. 2010, 73, 650–655. [Google Scholar] [CrossRef] [PubMed]
  143. Valeria D’Auria, M.; Zampella, A.; Paloma, L.G.; Minale, L.; Debitus, C.; Roussakis, C.; Le Bert, V. Callipeltins B and C; bioactive peptides from a marine Lithistida sponge Callipelta sp. Tetrahedron 1996, 52, 9589–9596. [Google Scholar] [CrossRef]
  144. Randazzo, A.; Bifulco, G.; Giannini, C.; Bucci, M.; Debitus, C.; Cirino, G.; Gomez-Paloma, L. Halipeptins A and B: Two novel potent anti-inflammatory cyclic depsipeptides from the Vanuatu marine sponge Haliclona species. J. Am. Chem. Soc. 2001, 123, 10870–10876. [Google Scholar] [CrossRef] [PubMed]
  145. Capon, R.J.; Ford, J.; Lacey, E.; Gill, J.H.; Heiland, K.; Friedel, T. Phoriospongin A and B: Two new nematocidal depsipeptides from the Australian marine sponges Phoriospongia sp. and Callyspongia bilamellata. J. Nat. Prod. 2002, 65, 358–363. [Google Scholar] [CrossRef] [PubMed]
  146. Plaza, A.; Gustchina, E.; Baker, H.L.; Kelly, M.; Bewley, C.A. Mirabamides A-D, depsipeptides from the sponge Siliquariaspongia mirabilis that inhibit HIV-1 fusion. J. Nat. Prod. 2007, 70, 1753–1760. [Google Scholar] [CrossRef] [PubMed]
  147. Tran, T.D.; Pham, N.B.; Fechner, G.; Zencak, D.; Vu, H.T.; Hooper, J.N.; Quinn, R.J. Cytotoxic cyclic depsipeptides from the Australian marine sponge Neamphius huxleyi. J. Nat. Prod. 2012, 75, 2200–2208. [Google Scholar] [CrossRef] [PubMed]
  148. Coello, L.; Reyes, F.; Martin, M.J.; Cuevas, C.; Fernandez, R. Isolation and structures of pipecolidepsins A and B, cytotoxic cyclic depsipeptides from the Madagascan sponge Homophymia lamellosa. J. Nat. Prod. 2014, 77, 298–303. [Google Scholar] [CrossRef] [PubMed]
  149. Martin, M.J.; Rodriguez-Acebes, R.; Garcia-Ramos, Y.; Martinez, V.; Murcia, C.; Digon, I.; Marco, I.; Pelay-Gimeno, M.; Fernandez, R.; Reyes, F.; et al. Stellatolides, a new cyclodepsipeptide family from the sponge Ecionemia acervus: Isolation, solid-phase total synthesis, and full structural assignment of stellatolide A. J. Am. Chem. Soc. 2014, 136, 6754–6762. [Google Scholar] [CrossRef] [PubMed]
  150. Clark, D.P.; Carroll, J.; Naylor, S.; Crews, P. An Antifungal Cyclodepsipeptide, Cyclolithistide A, from the SpongeTheonella swinhoei. J. Org. Chem. 1998, 63, 8757–8764. [Google Scholar] [CrossRef]
  151. Okada, Y.; Matsunaga, S.; van Soest, R.W.; Fusetani, N. Nagahamide A, an antibacterial depsipeptide from the marine sponge Theonella swinhoei. Org. Lett. 2002, 4, 3039–3042. [Google Scholar] [CrossRef] [PubMed]
  152. Plaza, A.; Bifulco, G.; Keffer, J.L.; Lloyd, J.R.; Baker, H.L.; Bewley, C.A. Celebesides A–C and theopapuamides B-D, depsipeptides from an Indonesian sponge that inhibit HIV-1 entry. J. Org. Chem. 2009, 74, 504–512. [Google Scholar] [CrossRef] [PubMed]
  153. Ratnayake, A.S.; Bugni, T.S.; Feng, X.; Harper, M.K.; Skalicky, J.J.; Mohammed, K.A.; Andjelic, C.D.; Barrows, L.R.; Ireland, C.M. Theopapuamide, a cyclic depsipeptide from a Papua New Guinea lithistid sponge Theonella swinhoei. J. Nat. Prod. 2006, 69, 1582–1586. [Google Scholar] [CrossRef] [PubMed]
  154. Plaza, A.; Bifulco, G.; Masullo, M.; Lloyd, J.R.; Keffer, J.L.; Colin, P.L.; Hooper, J.N.; Bell, L.J.; Bewley, C.A. Mutremdamide A and koshikamides C-H, peptide inhibitors of HIV-1 entry from different Theonella species. J. Org. Chem. 2010, 75, 4344–4355. [Google Scholar] [CrossRef] [PubMed]
  155. Tan, K.C.; Wakimoto, T.; Abe, I. Sulfoureido lipopeptides from the marine sponge discodermia kiiensis. J. Nat. Prod. 2016, 79, 2418–2422. [Google Scholar] [CrossRef] [PubMed]
  156. Cabrita, M.T.; Vale, C.; Rauter, A.P. Halogenated compounds from marine algae. Mar. Drugs 2010, 8, 2301–2317. [Google Scholar] [CrossRef] [PubMed]
  157. Rudi, A.; Chill, L.; Aknin, M.; Kashman, Y. Didmolamide A and B, two new cyclic hexapeptides from the marine ascidian Didemnum molle. J. Nat. Prod. 2003, 66, 575–577. [Google Scholar] [CrossRef] [PubMed]
  158. Donia, M.S.; Wang, B.; Dunbar, D.C.; Desai, P.V.; Patny, A.; Avery, M.; Hamann, M.T. Mollamides B and C, Cyclic hexapeptides from the indonesian tunicate Didemnum molle. J. Nat. Prod. 2008, 71, 941–945. [Google Scholar] [CrossRef] [PubMed]
  159. Salib, M.N.; Molinski, T.F. Cyclic Hexapeptide Dimers, Antatollamides A and B, from the Ascidian Didemnum molle. A Tryptophan-Derived Auxiliary for l- and d-Amino Acid Assignments. J. Org. Chem. 2017, 82, 10181–10187. [Google Scholar] [CrossRef] [PubMed]
  160. Dalisay, D.S.; Rogers, E.W.; Edison, A.S.; Molinski, T.F. Structure elucidation at the nanomole scale. 1. Trisoxazole macrolides and thiazole-containing cyclic peptides from the nudibranch Hexabranchus sanguineus. J. Nat. Prod. 2009, 72, 732–738. [Google Scholar] [CrossRef] [PubMed]
  161. Lee, J.-S.; Satake, M.; Horigome, Y.; Oshima, Y.; Yasumoto, T. Gamakamide-E, a strongly bitter tasting cyclic peptide with a hydantoin structure from cultured oysters crassostrea gigas. Fish. Aquatic. Sci. 2012, 15, 15–19. [Google Scholar] [CrossRef]
  162. Goetz, G.; Yoshida, W.Y.; Scheuer, P.J. The absolute stereochemistry of Kahalalide F. Tetrahedron 1999, 55, 7739–7746. [Google Scholar] [CrossRef]
  163. Vervoort, H.; Fenical, W.; Epifanio, R.A. Tamandarins A and B: New cytotoxic depsipeptides from a Brazilian ascidian of the family Didemnidae. J. Org. Chem. 2000, 65, 782–792. [Google Scholar] [CrossRef] [PubMed]
  164. Dmitrenok, A.; Iwashita, T.; Nakajima, T.; Sakamoto, B.; Namikoshi, M.; Nagai, H. New cyclic depsipeptides from the green alga Bryopsis species; application of a carboxypeptidase hydrolysis reaction to the structure determination. Tetrahedron 2006, 62, 1301–1308. [Google Scholar] [CrossRef]
  165. Horgen, F.D.; delos Santos, D.B.; Goetz, G.; Sakamoto, B.; Kan, Y.; Nagai, H.; Scheuer, P.J. A new depsipeptide from the sacoglossan mollusk Elysia ornata and the green alga Bryopsis species. J. Nat. Prod. 2000, 63, 152–154. [Google Scholar] [CrossRef] [PubMed]
  166. Whitson, E.L.; Ratnayake, A.S.; Bugni, T.S.; Harper, M.K.; Ireland, C.M. Isolation, structure elucidation, and synthesis of eudistomides A and B, lipopeptides from a Fijian ascidian Eudistoma sp. J. Org. Chem. 2009, 74, 1156–1162. [Google Scholar] [CrossRef] [PubMed]
  167. Iwasaki, A.; Ohno, O.; Sumimoto, S.; Matsubara, T.; Shimada, S.; Sato, T.; Suenaga, K. Mebamamides A and B, Cyclic Lipopeptides Isolated from the Green Alga Derbesia marina. J. Nat. Prod. 2015, 78, 901–908. [Google Scholar] [CrossRef] [PubMed]
  168. Berthod, A.; Liu, Y.; Bagwill, C.; Armstrong, D.W. Facile liquid chromatographic enantioresolution of native amino acids and peptides using a teicoplanin chiral stationary phase. J. Chromatogr. A 1996, 731, 123–137. [Google Scholar] [CrossRef]
  169. Péter, A.; Török, G.; Armstrong, D. High-performance liquid chromatographic separation of enantiomers of unusual amino acids on a teicoplanin stationary phase. J. Chromatogr. A 1998, 793, 283–296. [Google Scholar] [CrossRef]
Figure 1. Schematic presentation of the methodologies generally used for determination of the configuration of amino acid residues of marine peptides. HPLC—High Performance Liquid Chromatography; CSP—Chiral Stationary Phase; FDAA—1-Fluoro-2-4-dinitrophenyl-5-d,l-alanine amide; FDLA—1-Fluoro-2-4-dinitrophenyl-5-d,l-leucine amide.
Figure 1. Schematic presentation of the methodologies generally used for determination of the configuration of amino acid residues of marine peptides. HPLC—High Performance Liquid Chromatography; CSP—Chiral Stationary Phase; FDAA—1-Fluoro-2-4-dinitrophenyl-5-d,l-alanine amide; FDLA—1-Fluoro-2-4-dinitrophenyl-5-d,l-leucine amide.
Molecules 23 00306 g001
Figure 2. Structure of cyclic peptides 116, isolated from marine cyanobacteria and other bacteria, whose stereochemistry determination of their amino acids was performed by Marfey’s method (compounds 14) and by a combination of both Marfey’s method and chiral HPLC (compounds 516).
Figure 2. Structure of cyclic peptides 116, isolated from marine cyanobacteria and other bacteria, whose stereochemistry determination of their amino acids was performed by Marfey’s method (compounds 14) and by a combination of both Marfey’s method and chiral HPLC (compounds 516).
Molecules 23 00306 g002
Figure 3. Structure of wewakazole B (17) isolated from a marine cyanobacteria.
Figure 3. Structure of wewakazole B (17) isolated from a marine cyanobacteria.
Molecules 23 00306 g003
Figure 4. Structure of cyclic depsipeptides 1846, isolated from marine cyanobacteria and other bacteria, whose stereochemistry of their amino acids was determined only by chiral HPLC.
Figure 4. Structure of cyclic depsipeptides 1846, isolated from marine cyanobacteria and other bacteria, whose stereochemistry of their amino acids was determined only by chiral HPLC.
Molecules 23 00306 g004
Figure 5. Structure of cyclic depsipeptides 4778, isolated from marine cyanobacteria and other bacteria, whose stereochemistry of their amino acids was determined by a combination of Marfey’s method and chiral HPLC.
Figure 5. Structure of cyclic depsipeptides 4778, isolated from marine cyanobacteria and other bacteria, whose stereochemistry of their amino acids was determined by a combination of Marfey’s method and chiral HPLC.
Molecules 23 00306 g005
Figure 6. Structure of cyclic depsipeptides 7994, isolated from marine cyanobacteria and other bacteria, whose stereochemistry of their amino acids was determined by Marfey’s method.
Figure 6. Structure of cyclic depsipeptides 7994, isolated from marine cyanobacteria and other bacteria, whose stereochemistry of their amino acids was determined by Marfey’s method.
Molecules 23 00306 g006
Figure 7. Structure of lipopeptides 9598, isolated from marine cyanobacteria.
Figure 7. Structure of lipopeptides 9598, isolated from marine cyanobacteria.
Molecules 23 00306 g007
Figure 8. Structure of cyclic peptides 99112, isolated from marine-derived fungi.
Figure 8. Structure of cyclic peptides 99112, isolated from marine-derived fungi.
Molecules 23 00306 g008
Figure 9. Structure of cyclic depsipeptides 113131, isolated from marine-derived fungi.
Figure 9. Structure of cyclic depsipeptides 113131, isolated from marine-derived fungi.
Molecules 23 00306 g009
Figure 10. Structure of cyclic peptides 132151, isolated from marine sponges.
Figure 10. Structure of cyclic peptides 132151, isolated from marine sponges.
Molecules 23 00306 g010
Figure 11. Structure of cyclic peptides 152164, isolated from marine sponges.
Figure 11. Structure of cyclic peptides 152164, isolated from marine sponges.
Molecules 23 00306 g011
Figure 12. Structure of cyclic depsipeptides 165179, isolated from marine sponges.
Figure 12. Structure of cyclic depsipeptides 165179, isolated from marine sponges.
Molecules 23 00306 g012
Figure 13. Structure of cyclic depsipeptides 180195, isolated from marine sponges.
Figure 13. Structure of cyclic depsipeptides 180195, isolated from marine sponges.
Molecules 23 00306 g013
Figure 14. Structure of cyclic lipopeptides 196198, isolated from marine sponges.
Figure 14. Structure of cyclic lipopeptides 196198, isolated from marine sponges.
Molecules 23 00306 g014
Figure 15. Structure of cyclic peptides 199206, isolated from marine invertebrates and marine algae.
Figure 15. Structure of cyclic peptides 199206, isolated from marine invertebrates and marine algae.
Molecules 23 00306 g015
Figure 16. Structure of cyclic depsipeptides 207217, isolated from marine invertebrates and marine algae.
Figure 16. Structure of cyclic depsipeptides 207217, isolated from marine invertebrates and marine algae.
Molecules 23 00306 g016aMolecules 23 00306 g016b
Figure 17. Structure of lipopeptides 218221, isolated from marine invertebrates and marine algae.
Figure 17. Structure of lipopeptides 218221, isolated from marine invertebrates and marine algae.
Molecules 23 00306 g017
Figure 18. Chromatograms of enantiomeric mixture of dl-Ala (A), dl-pipecolic acid (B), and dl-Val (C). Column, Chirobiotic T; Mobile phase, MeOH:H2O:acetic acid (70:30:0.02 v/v/v); Flow rate, 1.0 mL/min; UV detection, 210 nm.
Figure 18. Chromatograms of enantiomeric mixture of dl-Ala (A), dl-pipecolic acid (B), and dl-Val (C). Column, Chirobiotic T; Mobile phase, MeOH:H2O:acetic acid (70:30:0.02 v/v/v); Flow rate, 1.0 mL/min; UV detection, 210 nm.
Molecules 23 00306 g018
Figure 19. Chromatograms of enantiomeric mixture of dl-Ala (a), l-Ala (b), and d-Ala (c). Column, Chirobiotic T; Mobile phase, MeOH:H2O:acetic acid (70:30:0.02 v/v/v); Flow rate, 1.0 mL/min; UV detection, 210 nm.
Figure 19. Chromatograms of enantiomeric mixture of dl-Ala (a), l-Ala (b), and d-Ala (c). Column, Chirobiotic T; Mobile phase, MeOH:H2O:acetic acid (70:30:0.02 v/v/v); Flow rate, 1.0 mL/min; UV detection, 210 nm.
Molecules 23 00306 g019
Figure 20. Distribution of the reported studies concerning the determination of the stereochemistry of marine peptides according to the methods used.
Figure 20. Distribution of the reported studies concerning the determination of the stereochemistry of marine peptides according to the methods used.
Molecules 23 00306 g020
Figure 21. Distribution of the studies concerning the determination of the stereochemistry of marine peptides according to the method used before (A) and after 2007 (B).
Figure 21. Distribution of the studies concerning the determination of the stereochemistry of marine peptides according to the method used before (A) and after 2007 (B).
Molecules 23 00306 g021
Table 1. Cyclic peptides from marine cyanobacteria and other bacteria.
Table 1. Cyclic peptides from marine cyanobacteria and other bacteria.
PeptideSourceaa CompositionChromatographic ConditionsBiological ActivitiesRefs.
Tetrapeptide (1)Bacterium Nocardiopsis sp.l-Ile, l-Leu, l-ProMarfey’s method (FDAA) combined with HPLC
C18 (YMC-ODS-A) (4.5 × 250 mm)
Flow rate: 0.8 mL/min; UV detection at 340 nm
MP: ACN(aq) (0–50% (v/v)) with 0.1% TFA
Cytotoxicity toward the leukemia cell-line K-562[60]
Anabaenopeptins NZ825, NZ841, NZ857 (24)Cyanobacterium Anabaena sp.l-Ile, d-Lys, l-Phe;
2: l-Hph; 3: l-Hph, l-Hty;
4: l-Hty
Marfey’s method (FDAA) combined with HPLC
Merck Chromolith performance RP-18e, (4.6 × 100 mm)
MP: 50 mM TEAP buffer (pH 3)/ACN (9:1 to 1:1 v/v)
No inhibition of serine proteases[61]
Aurilides B (5) and C (6)Cyanobacterium Lyngbya majusculal-Val, N-Me-l-Ile, l-IleLigand Exchange Type CSP; Phenomenex Chirex 3126 (D) (4.6 × 250 mm);
Flow rate: 1.0 mL/min; UV detection at 254 nm;
MP: 2 mM CuSO4 in ACN/H2O (5/95 v/v) or 2 mM CuSO4 in ACN/H2O (15/85 v/v)
Cytotoxicity against NCl-H460 and neuro-2a mouse neuroblastoma cell lines
5: also active against leukemia, renal, and prostate cancer cell lines
[62]
N-Me-l-Ala
6: N-Me-l-allo-Ile,
d-Hiva
Marfey’s method (FDAA) combined with HPLC
Microsob-MV C18 (4.6 × 250 mm)
Flow rate: 1.0 mL/min; UV detection at 254 nm
MP: 50 mM TEAP buffer pH 3/ACN (9:1 to 1:1 v/v)
Urukthapelstatin A (7)Marine Derived Mechercharimyces asporophorigenens YM11-542l-AlaMarfey’s method (FDAA) combined with HPLC
ODS-80Ts column (4.6 × 150 mm)
Flow rate:1.0 mL/min; UV detection at 340 nm
MP: MeOH, 0.1% TFA containing ACN or H2O
Growth inhibition of human lung cancer A549 cells, cytotoxicity against a human cancer cell line panel[63,70]
d-allo-IleLigand Exchange Type CSPSumichiral OA-5000 column (4.6 × 150 mm)
Flow rate: 1.0 mL/min; UV detection at 254 nm
MP: 5% IPA containing 2 mM CuSO4
Pompanopeptins A (8) and B (9)Cyanobacterium Lyngbya confervoides8: l-Val, l-Thr, l-Met (O), S-Ahp, l-Ile, l-Arg
9: l-Ile
Ligand Exchange Type CSP; Phenomenex Chirex 3126 N,S-dioctyl-(d)-penicillamine, 5 µm (4.6 × 250 mm)
Flow rate: 1.0 mL/min; UV detection at 254 nm
MP: 2 mM CuSO4 or 2 mM CuSO4/ACN (95:5 v/v)
8: Trypsin inhibitory activity[64]
8: N,O-diMe-Br-l-TyrMarfey’s method (FDLA) combined with HPLC-MS
Phenomenex Synergi 4u Hydro RP 80A (2 × 340 nm)
Flow rate: 0.15 mL/min; UV detection at 254 nm
MP: ACN/HCOOH (10–90:0.1 v/v) in gradient
9: d-Lys, l-Val, d-GluMarfey’s method (FDLA) combined with HPLC-MS
Alltech Altima HP C18 HL 54 (250 × 4.6 mm)
Flow rate: 1.0 mL/min; PDA detection from 200–500 nm
MP: ACN/aq TFA (30–70:0.1 v/v) in gradient
Marthiapeptide A (10)Deep sea-derived Marinactinospora thermotolerans SCSIO 00652l-IleLigand Exchange Type CSP; MCIGELCR10W (4.6 × 150 mm);
Flow rate: 0.5 mL/min; UV detection at 254 nm;
MP: 2 mM CuSO4 solution
Antibacterial and cytotoxic activities [65]
d-Phe, l-IleMarfey’s method (FDAA) combined with HPLC
Zorbax SB-C8 column, 5 μm (2.1 × 30 mm)
Nocardiamides A (11) and B (12)Marine-derived Actinomycete Nocardiopsis sp. CNX037l-Tyr, d-Leu,
d- and l-Val
Marfey’s method (FDAA or FDLA) combined with HPLC; Conditions not describedAntimicrobial activity and no cytotoxicity against HCT-116 cell line[66]
11: l-IleLigand Exchange Type CSP; MCIGELCRS10W, (4.6 × 250 mm);
Flow rate: 0.5 mL/min; UV detection at 254 nm;
MP: 2 mM CuSO4/H2O
Destomides B–D (1315)Deep sea-derived Streptomyces scopuliridis SCSIO ZJ46l-Asn, d-Leu
13: l-Trp, l-Val, l-Leu;
14: l-Gly, l-Ile,
15: l-Gly, l-Ile, l-Leu
Marfey’s method (FDAA) combined with HPLC
Phenomenex ODS column, 5 µm (4.6 × 150 mm)
Flow rate: 1.0 mL/min; UV detection at 340 nm
MP: ACN:H2O:TFA (15:85:0.1 to 90:10:0.1)
13: Antimicrobial activity against staphylococcus aureus ATCC 29213, Streptococcus pneumoniae NCTC 7466 and MRSE shhs-E1
1315: no cytotoxicity
[67]
15: l-KynLigand Exchange Type CSP; MCIGELCRS10W column, 3 µm (4.6 × 50 mm);
Flow rate: 0.5 mL/min; UV detection at 254 nm;
MP: 2 mM CuSO4 aqueous solution
Janadolide (16)Cyanobacterium Okeania sp.N-Me-l-Leu, l-Pro,
l-Val
Ligand Exchange Type CSP; Diacel CHIRALPAK (MA+) (4.6 × 50 mm);
Flow rate: 1.0 mL/min; UV detection at 254 nm;
MP: 2.0 mM CuSO4
Antitrypanosomal activity [68]
N-Me-l-AlaMarfey’s method (FDAA) combined with HPLC
Cosmosil Cholester (4.6 × 50 mm);
Flow rate: 1.0 mL/min; UV detection at 340 nm
MP: 0.02 M NaOAc(aq)/MeOH (45/55 v/v)
Wewakazole B (17)Cyanobacterium Moorea producensl-Ala, l-Phe, l-ProMacrocyclic Antibiotic type CSP
Chirobiotic TAG (2.1 × 250 mm);
Flow rate: 0.3 mL/min; UV detection at 340 nm;
MP: 0.1% aq. HCOOH and 1% (v/v) NH4OAc in MeOH
Cytotoxicity against MCF7 and human 460 lung cancer cell lines[69]
l-IleLigand Exchange type CSP; Sumichiral OA-5000 (4.6 × 150 mm);
Flow rate: 1.0 mL/min; UV Detection at 254 nm;
MP: MeOH/2.0 mM CuSO4 in H2O (5/95 v/v)
aa—Amino acid; FDAA—1-Fluoro-2-4-dinitrophenyl-5-l-alanine amide; ESI—Electrospray Ionization; LC—Liquid Chromatography; MS—Mass spectrometry; HPLC—High Performance Liquid Chromatography; MP—Mobile Phase; TEAP—Triethylammonium phosphate; ACN—Acetonitrile; CPA—Carboxypeptidase A; TFA—Trifluoracetic acid; MeOH—Methanol; TEA—Triethylamine; IPA—Isopropyl alcohol; FDLA—1-fluoro-2-4-dinitrophenyl-5-d,l-leucine amide; NaOAc—Sodium acetate; NH4OAc—Ammonium acetate.
Table 2. Cyclic depsipeptides from marine cyanobacteria and other bacteria.
Table 2. Cyclic depsipeptides from marine cyanobacteria and other bacteria.
PeptideSourceaa CompositionChromatographic ConditionsBiological ActivitiesRefs.
Malevamides
B (18) and C (19)
Cyanobacterium
Symploca laete-viridis
l-Pro, N-Me-l-Val, N-Me-l-Phe
18: l-Ile, N-Me-l-Ala, N-Me-d-Val, l-Val, (R)-Hiva; 19: l-Ala, N-diMe-l-Ser, l-Leu, N-Me-d-Ala, N-Me-l-Ile, (S)-Hiva
Ligand Exchange Type CSP; Chirex (D) Penicillamine, Phenomenex 00G-3126E0 (4.6 × 250 mm)
MP: 1.7 mM CuSO4 in ACN/H2O (14:86 v/v), 1.9 mM CuSO4 in ACN/H2O (5:95 v/v) or 2.0 mM CuSO4 in H2O
Flow rate: 1.0 and 0.8 mL/min; UV detection at 245 nm
Inactive against P-388, A-549 and HT-29 cancer cells[71]
Lyngbyapeptin B (20)Cyanobacterium Lyngbya majusculaN-Me-l-Ile, N-Me-l-Leu,
N,O-diMe-l-Tyr
Ligand Exchange Type CSP; Chirex (D) Penicillamine, Phenomenex 00G-3126E0 (4.6 × 250 mm)
MP: 2 mM CuSO4
Flow: 0.8 mL/min; UV detection at 254 nm
Cytotoxicity against KB and LoVo cells[72]
Tasipeptins A (21) and B (22)Cyanobacterium
Symploca sp.
l-Thr, l-Val, l-Leu, l-Glu,
N-Me-l-Phe
Ligand Exchange Type CSP; Phenomenex Chirex Phase 3126 (D) (4.6 × 250 mm)
MP: 2 mM CuSO4; 2 mM CuSO4/ACN (95:5 or 85:15 v/v)
UV detection at 254 nm
Cytotoxicity toward KB cells[73]
Wewakamide A (23)Cyanobacteria
Lyngbya semiplena and Lyngbya majuscula
l-M-Ala, l-Pro, l-Val, l-Me-Leu, l-Phe, l-Me-ILe, l-HivLigand Exchange Type CSP; Phenomenex Chirex 3126 (D) (4.6 × 250 mm);
MP: 2 mM CuSO4 in H2O or 2 mM CuSO4 in ACN/H2O (15:85 or 5:95 v/v)
Flow rate: 0.7, 0.8, 1.0 mL/min; UV detection at 254 nm
Brine shrimp toxicity[74]
Cocosamide A (24) and B (25)Cyanobacterium
Lyngbya majuscula
l-Pro, l-Val, N-Me-l-PheLigand Exchange Type CSP; Phenomenex Chirex (D), Penicillamine, 5 µm (4.6 × 250 mm)
MP: 2.0 mM CuSO4/ACN (85:15 or 90:10 v/v)
Flow rate: 1.0 mL/min; UV detection at 254 nm
Cytotoxicity against MCF-7 (breast cancer) and HT-29 (colon cancer) cells[75]
Dudawalamides A–D
(2629)
Cyanobacterium
Moorea producens
l-Dhoya, l-Hiva, l-Val
29: d-allo-Hiva
Ligand Exchange Type CSP; Chirex Phase 3126 (D) 5 µm (4.6 × 250 mm);
MP: 2 mM CuSO4-ACN (95:5 or 85:15 v/v or 87.5:12.5 v/v/v), ACN-H2O-HCOOH (30:70:0.1 or 70:30:0.1 v/v/v)
Flow rate: 0.8 mL/min; UV detection at 340 nm
Antiparasitic activity [76]
Pitipeptolides
A (30) and B (31)
Cyanobacterium
Lyngbya majuscula
l-Gly, l-Pro, l-Val, l-Ile, N-Me-l-Phe, (2S,3S)-HmpLigand Exchange Type CSP; Chiralpak MA (+) (4.6 × 50 mm);
MP: 2 mM CuSO4: ACN (90:10 or 85:15 v/v)
Flow rate: 1.0 mL/min; UV detection at 254 nm
Cytotoxic, antimycobacterial and elastase inhibitory activities[77]
Kohamamides A–C
(3234)
Cyanobacterium
Okeania sp.
l-Pro, l-Ala, l-Val, N-Me-l-Val, l-Leu; 32: l-IleLigand Exchange Type CSP; Chiralpak MA (+) (4.6 × 250 mm);
MP: 2 mM CuSO4, ACN: 2 mM CuSO4 (15:85 v/v);
Flow rate: 1.0 mL/min; UV detection at 254 nm
No growth inhibition against HeLa and HL60 cells[78]
Marformycins A–F (3540)Deep sea-derived
Streptomyces drozdowiczii
35: d-allo-Ile, l-Val; 36: d-allo-Ile, l-allo-Ile; 37: d-Val, l-allo-Ile; 38: d-allo-Ile, l-allo-Ile, l-Leu; 39 and 40: l-Thr, l-Val, d-Val, l-LeuLigand Exchange Type CSP; MCIGELCRS10W (4.6 × 50 mm);
MP: 2 mM CuSO4 in H2O
Flow rate: 0.5 mL/min; UV detection at 254 nm
Anti-infective activity against Micrococcus luteus[79]
Pitiprolamide (41)Cyanobacterium
Lyngbya majuscula
l-Pro, l-ValMacrocyclic Antibiotic Type CSP; Chirobiotic TAG (4.6 × 250 mm);
MP: MeOH/10 mM NH4OAc (40:60 v/v) (pH 5.6)
Flow rate: 0.5 mL/min
Cytotoxicity against CT116 and MCF7 cancer cell lines and antibacterial activity [80]
Palau’amide (42)Cyanobacterium
Lyngbya sp.
l-Ala, l-Ile, N-Me-l-Ala, N-Me-d-Phe and d-hydroxyisocaproic acidLigand Exchange Type CSP; Phenomenex Chirex Phase 3126 (D) (4.6 × 250 mm)
MP: 1 mM CuSO4; 2 mM CuSO4/ACN (95:5 or 85:15 v/v)
Flow rate: 0.8 mL/min; UV detection at 254 nm
Cytotoxicity against KB cell line[81]
Pitipeptolides C–F (4346)Cyanobacterium
Lyngbya majuscula
l-Pro, l-Val, l-Ile, l-Phe,
N-Me-l-Phe
Macrocyclic Antibiotic Type CSP; Chirobiotic TAG (4.6 × 250 mm);
MP: MeOH/10 mM NH4OAc (40:60 v/v) (pH 5.6);
Flow rate: 0.5 mL/min
Detection by EIMS in positive ion mode (MRM scan)
46: Active against Mycobacterium tuberculosis[82]
Ulongapeptin (47)Cyanobacterium
Lyngbya sp.
l-lactic acid, l-Val, N-Me-l-Val,
N-Me-d-Val, N-Me-d-Phe
Ligand Exchange Type CSP; Phenomenex Chirex Phase 3126 (D), 4.6 × 250 mm
MP: 2 mM CuSO4; 2 mM CuSO4/ACN (95:5 or 85:15 v/v)
Flow rate: 1.00 mL/min; UV detection at 254 nm
Cytotoxicity against KB cells [83]
l-Val, N-Me-l-Val, N-Me-d-ValMarfey’s method (FDLA) combined with HPLC
YMC-Pack AQ-ODS (10 × 250 mm); MP: 50% ACN in 0.01 N TFA
Flow rate: 2.5 mL/min; UV detection at 254 nm
2-hydroxy-3-methylvaleric acid
N-Me-l-Ala
Ligand Exchange Type CSP; CHIRALPAK MA (+) (4.6 × 50 mm);
MP: 1 mM CuSO4; 2 mM CuSO4/ACN (95:5 or 85:15 v/v)
Flow rate: 0.7 mL/min; UV detection at 254 nm
Largamides A–H (4855)Cyanobacterium
Oscillatoria sp.
48: l-Val, l-Thr, l-Ala, l-Leu, d-Gln, d-Tyr; 49: l-Val, l-Thr, l-Ala, l-Ahppa, d-Gln, d-Tyr; 50: l-Val, l-Thr, l-Ala, l-Ahpha, d-Gln, d-Tyr; 51: l-Val, l-Thr, l-Ala, l-Leu, l-Ahp, N-MeBr-l-Tyr, l-Ahppa; 52: l-Val, l-Thr, l-Ala, l-Leu, l-Ahp, N-MeCl-l-Tyr; 53: l-Val, l-Thr, l-Ala, l-Tyr, l-Ahp, N-MeCl-l-Tyr; 54: l-Val, l-Thr, l-Ala, l-hTyr, l-Ahp, N-MeCl-l-Tyr; 55: l-Val, l-Thr, l-Ala, l-Amppa, l-Gln, N-Me-l-AsnMarfey’s method (FDLA) combined with HPLC
Phenomenex Jupiter Proteo C12 column, 4 µm (4.6 × 150 mm);
MP: ACN containing 0.01 M TFA
Flow 0.5 mL/min; UV detection at 254 nm
5154: Chymotrypsin inhibition [84]
d-Glyceric acidLigand Exchange Type CSP; Phenomenex Chirex 3126 (D) (4.6 × 150 mm);
MP: 2 mM CuSO4:ACN (90/10 v/v);
Flow 0.5 mL/min; UV detection at 254 nm
Trungapeptins
A–C (5658)
Cyanobacterium
Lyngbya majuscula
l-Val, l-N-MeVal, l-alloLeu,
l-Pro
Marfey’s method (FDLA) combined with HPLC. Alltech Econosil C18;
MP A:40% ACN with 0.04%TFA. MP B: 37.5% ACN with 0.05%TFA.
Flow rate: 1.0 mL/min; UV detection at 254 nm
Brine shrimp toxicity and ichthyotoxicity [85]
Phenyllactic acid (S)Ligand Exchange Type CSP; CHIRALPAK MA (+) (4.6 × 50 mm);
MP: 2 mM CuSO4/ACN (85:15)
Flow rate: 0.5 mL/min; UV detection at 254 nm
Carriebowmide (59)Cyanbacterium
Lyngbya polychroa
l-Ala, N-Me-l-Leu, N-Me-d-Phe, l-Phe, l-MetLigand Exchange Type CSP; Phenomenex, Chirex (D) Penicillamine, 5μm (4.6 × 250 mm)
MP: 2.0 mM CuSO4-ACN (95:5, 90:10, or 85:15 v/v)
Flow rate: 0.8 or 1.0 mL/min; UV detection at 254 nm
Lipophilic extract reduced feeding on agar food pellets[86]
R-HmbaLigand Exchange Type CSP; Chiralpak MA (+) (4.6 × 250 mm);
MP: 2.0 mM CuSO4-ACN (90:10 v/v)
Flow rate:1.0 mL/min; UV detection at 254 nm
l-AbaLigand Exchange Type CSP; Phenomenex, Chirex (D) Penicillamine, 5 μm (4.6 × 250 mm);
MP: 2.0 mM CuSO4
Flow rate:1.0 mL/min; UV detection at 254 nm
(2R,3R)-AmhaMarfey’s method (FDAA) combined with HPLC
Atlantis, C18, (3.0 × 250 mm);
MP: 50 mM NH4COOCH3(aq)-ACN (70:30 v/v)
Flow rate: 1.0 mL/min; UV detection at 254 nm
Symplocamide A (60)Cyanobacterium Symploca sp.l-Val, l-Thr, l-Ile, l-Cit, l-Gln, l-Btyr, l-ButMarfey’s method (FDAA) combined with HPLC
Phenomenex Jupiter C18 column (4.6 × 250 mm)
MP: ACN:H2O:HOAc (15:85:0.02 to 1:1:0.02 v/v/v)
Flow rate: 0.5 mL/min; UV detection at 340 nm
Cytotoxicity and antimicrobial
activities
Chymotrypsin inhibitor
[87]
Kempopeptins A (61) and B (62)Cyanobacterium
Lyngbya sp.
61: N-O-diMe-Br-l-TyrMarfey’s method (FDLA) combined with HPLC
Conditions not described
61: Elastase and chymotrypsin inhibition
62: Trypsin inhibition
[88]
61: N-Me-l-Tyr, l-Val, l-Thr-2, l-Pro, l-Phe, l-Ahp, l-Leu
62: l-Lys, l-Thr, l-Val, l-Ile
Ligand Exchange Type CSP; Phenomenex Chirex Phase 3126 N,S-dioctyl-(d)-penicillamine column, 5 μm (4.6 × 250 mm);
MP: 2 mM CuSO4 in H2O:ACN (95:5 v/v) or 2 mM CuSO4
Flow rate: 1.0 mL/min; UV detection at 254 nm
Tiglicamides A–C (6365)Cyanobacterium
Lyngbya confervoides
l-Ala, l-Thr, l-Val, d-Glu, d-Tyr; 63: l-Htyr; 65: l-Met (O)Ligand Exchange Type CSP; Phenomenex, Chirex 3126, 5 μm (4.6 × 250 mm); Mobile Phase: 2 mM CuSO4
Flow rate: 1.0 mL/min; UV detection at 254 nm
Porcine pancreatic elastase inhibition[89]
65: l-PheMarfey’s method (FDLA) combined with HPLC
Alltech Alltima HP C18, 5μm (4.6 × 250 mm)
MP: 50–100% MeOH in 0.1% (v/v) aqueous TFA
Flow rate: 0.8 mL/min; PDA detection at 200–500 nm
Hantupeptin B (66)Cyanobacterium
Lyngbya majuscula
l-Pro, l-Val, N-Me-l-Val,
N-Me-l-Ile
Marfey’s method (FDAA) combined with HPLC
Phenomenex, Luna, 5 µm, (2.0 × 150 mm);
MP: ACN in 0.1% (v/v) aqueous HCOOH;
Flow rate: 0.2 mL/min
Cytotoxicity against MOLT-4 (leukemic) and MCF-7 (breast cancer) cell lines[90]
l-3-phenyllactic acid (S)Ligand Exchange Type CSP; Chiralpak MA (+) (4.6 × 500 mm)
MP: 2 mM CuSO4/ACN (85:15 v/v)
Flow rate: 0.7 mL/min; UV detection at 218 nm
Palmyramide A (67)Cyanobacterium
(Lyngbya majuscula) and a red alga Centroceras sp. complex
l-Val, N-Me-l-Val, l-ProMarfey’s method (FDAA) combined with HPLC/MS on a Merck LiChrospher 100 RP-18 (4.0 × 125 mm)
MP: ACN:H2O:HCOOH (30:70:0.1 to 70:30:0.1 v/v/v) or 2.0 mM CuSO4 in H2O
Flow rate: 0.7 mL/min; UV detection at 254 nm
Sodium channel blocking activity in neuro-2a cells and cytotoxic activity in H-460 (human lung carcinoma) cells[91]
l-Lac, l-PlaLigand Exchange Type CSP; Phenomenex Chirex 3126 (4.6 × 250 mm); Conditions not described
Veraguamides A–G (6874)Cyanobacterium
Symploca cf. hydnoides
6871, 73 and 74: l-Val, N-Me-l-Val, l-Pro; 70: (2S,3R) Br-Hmoya; 71: N-Me-l-Ile; 72: l-Ile, N-Me-l-Val, N-Me-l-Ile, l-ProMacrocyclic Antibiotic Type CSP; Chirobiotic TAG (4.6 × 250 mm);
MP: MeOH/10 mM NH4OAc (40:60 v/v) (pH 5.6);
Flow rate: 0.5 mL/min
Cytotoxic activity against HT29 (colorectal adenocarcinoma) and HeLa (cervical carcinoma) cell lines[92]
74: 2S:3R dpv
2R:3R Dml
Marfey’s method (FDAA) combined with HPLC-MS
Phenomenex Synergi Hydro-RP (4.6 × 150 mm)
MP: MeOH:H2O:HCOOH (40–100% MeOH: 0.1% HCOOH);
Flow rate: 0.5 mL/min
Porpoisamides A (75) and B (76)Cyanobacterium
Lyngbya sp.
75 and 76: l-Ala, l-Pro, N-Me-d-Phe, (2S,3S)-HmpaLigand Exchange Type CSP; Phenomenex Chirex 3126 (4.6 × 250 mm);
MP: 5% or 15% ACN in 2 mM CuSO4 in H2O;
Flow rate:1.0 mL/min
Cytotoxicity against HCT 116 (colorectal carcinoma) and U2OS (osteosarcoma) cells[93]
75: (2S,3R)-Amoa
76: (2R,3R)-Amoa
Ligand Exchange Type CSP; Chiralpak MA (+) (4.6 × 50 mm);
MP: 15% ACN in 2 mM CuSO4 in H2O
Flow rate: 1.0 mL/min
(2R,3R) 3-NH2-2-Me-octanoic acidMarfey’s method (FDAA) combined with HPLC
YMC-Pack AQ-ODS (10 × 250 mm)
MP: ACN:H2O: N-TFA (57:43:0.1 v/v/v)
Flow rate: 2.5 mL/min; UV detection at 340 nm
76: (2S)-HivaLigand Exchange Type CSP; CHIRALPAK MA (+) (4.6 × 50 mm);
MP: ACN/2 mM CuSO4 (10:90 v/v)
Flow: 1.0 mL/min; UV detection at 254 nm
Companeramides A (77) and B (78)Cyanobacterial assemblage collected from Coiba National Park, Panama77: l-Ala, N-Me-l-Ala, l-Pro, l-Ile, N-Me-l-Leu, and N-Me-l-Val; 78: l-Pro, N-Me-l-Val, l-Val, l-Ile, d- and N-Me-l-AlaMarfey’s method (FDAA) combined with HPLC
C18 column (3.9 × 150 mm)
MP: 40 mM NH4OAc (pH 5.2):ACN (9:1 to 1:1 v/v)
Flow rate: 1.0 mL/min; UV detection at 340 nm
Antiplasmodial activity against Plasmodium falciparum[94]
S-HivaLigand Exchange Type CSP; Phenomenex Chirex 3126 (D) (4.6 × 250 mm);
MP: CuSO4/ACN
Flow: 1.0 mL/min; UV detection at 254 nm
Piperazimycins A–C (7981)Fermentation broth of a Streptomyces sp.(S)-AMNA, (S,S)-OHPip1, (R,R)-γOHPip2, 79: (S)-αMeSerMarfey’s method (FDAA) combined with HPLC
C18;
MP: ACN in H2O (10–100%)
Flow rate: 1.0 mL/min; UV detection: 210, 254, 340 nm
79: Active against diverse cancer cell lines [95]
Grassypeptolides D (82) and E (83)Red sea cyanobacterium
Leptolyngbya sp.
d-allo-Thr, N-Me-d-Leu, l-Thr, N-Me-l-LeuMarfey’s method (FDAA) combined with HPLC
Gemini C18 110 A, 5 µm (4.6 × 250 mm)
Cytotoxicity against HeLa and mouse neuro-2a blastoma cells[96]
l-PLa, N-Me-l-Val, l-Pro, N-Me-l-Phe, (2S)-MeCysA, d-Aba, l-Cya, (2R,3R)-MabaMarfey’s method (FDAA) combined with HPLC
Kinetex XB-C18, 110 A, 2.6 µm (4.6 × 100 mm)
MP: ACN:H2O:HCOOH (30:70:0.1 to 70:30:0.1 v/v/v) or ACN:H2O:TFA (30:70:0.1 to 70:30:0.1 v/v/v);
Flow rate: 0.2 mL/min; UV detection at 340 nm and ESIMS
Fijimycins A–C (8486)Fermentation broth of Streptomyces sp. strain CNS-57584: d-PhSar, l-Ala, l-DiMe-Leu, Sar, d-Hyp, d-Leu, l-Thr; 85: l-N-MeLeu, l-Ala, l-DiMeLeu, Sar, d-Hyp, d-Leu, l-Thr; 86: l-PhSar, l-Ser, l-DiMeLeu, Sar, d-Hyp, d-Leu, l-ThrMarfey’s method (FDAA) combined with HPLC
C18 column, Luna (4.6 × 100 mm)
MP: ACN:H2O:TFA (10:90:1 to 50:50:1 v/v/v)
Flow rate: 0.7 mL/min; UV detection at 340 nm
Antibacterial activity against three MRSA strains of Staphylococcus aureus[97]
Itralamides A (87) and B (88), and Carriebowmide sulfone (89)Cyanobacterium
Lyngbya majuscula
87: l-Ala, d-Ala, N-Me-l-Ala,
N-Me-d-Phe, N-Me-l-Thr,
N-Me-l-Val
Marfey’s method (FDLA) combined with HPLC
Eclipse XDB-18, Agilent (4.6 × 150 mm)
MP: ACN:H2O:HCOOH (20:80:0.1 to 80:20:0.1 v/v/v)
Flow rate: 0.8 mL/min; Detection by ESI-MS
88: Cytotoxicity against HEK293 (human embryonic kidney) cell line[98]
88: N-Me-l-Ala, N-Me-d-Phe, N-Me-l-Thr, d-ValMarfey’s method (FDLA) combined with HPLC
Luna C18, Phenomenex, 5 µm (4.6 × 250 mm)
MP: ACN:H2O:HCOOH (20:80:0.1 to 90:10:0.1 v/v/v)
Flow rate: 0.8 mL/min
89: (2S,3R)-AMHAMarfey’s method (FDLA) combined with HPLC-PDA
dC18, 5 µm (3.0 × 250 mm); MP: ACN:H2O:HCOOH (0:100:0.1 to 50:50:0.1 v/v/v); Flow rate: 0.3 mL/min
Viequeamide A (90)Marine button cyanobacterium
Rivularia sp.
l-Val, l-Thr, N-Me-l-Val, l-ProMarfey’s method (FDLA) combined with HPLC
Conditions not described
Highly toxic to H460 (human lung cancer) cells[99]
Ngercheumicin
F–I (9194)
Photobacterium related to P. halotoleransl-Ser, l-allo-Thr, d-Ser, d-Thr, l-Leu, d-LeuMarfey’s method (FDLA) combined with HPLC
Dionex RSLC Ultimate 300 with a diode array detector
Kinetex C18 column, 2.6 µm at 60 °C (2.1 × 150 mm)
ACN:H2O:TFA (0:100:0.1 to 50:50:0.1 v/v/v)
Flow rate: 0.8 mL/min
9193: rnaIII inhibiting activities[100]
aa—Amino acid; FDAA—1-Fluoro-2-4-dinitrophenyl-5-l-alanine amide; LC—Liquid Chromatography; MS—Mass Spectrometry; HPLC—High Performance Liquid Chromatography; MP—Mobile Phase; TEAP—Triethylammonium phosphate; ACN—Acetonitrile; TFA—Trifluoracetic acid; MeOH—Methanol; TEA—Triethylamine; IPA—Isopropyl alcohol; FDLA—1-Fluoro-2-4-dinitrophenyl-5-d,l-leucine amide; NaOAc—Sodium acetate; NH4OAc—Ammonium acetate.
Table 3. Lipopeptides from marine cyanobacteria.
Table 3. Lipopeptides from marine cyanobacteria.
PeptideSourceaa CompositionChromatographic ConditionsBiological ActivitiesRef.
Antillatoxin B (95)Cyanobacterium Lyngbya majusculaN-Me-l-HphMarfey’s method (FDAA) combined with HPLC Waters Nova-Pak C18 (3.9 × 150 mm),
MP: 10 to 50% ACN in H2O with 0.05% TFA, UV detection at 340 nm
Sodium channel-activating and ichthyotoxic activities[101]
Lobocyclami-des A–C (9698)Cyanobacterium Lyngbya confervoides96: S-Ile, S-allo-Ile, S-Leu, R-β-Aoa, S-Ser, R-Tyr, S-Hse, R-HprLigand Exchange Type CSP Chirex 3126 (d)-penicillamine column;
MP: 2 mM aq CuSO4/ACN (1:99, 95:5 or 86:14 v/v);
Flow rate: 1.15–1.20 mL/min, UV detection at 254 nm
Antifungal activity against a panel of Candida sp.[102]
97: S-Ala, S-Thr, N-Me-S-Ile, R-Aoa, R-Ada, 2R,3R-4-OH-Hth, 2R,3S-3-OH-Leu, trans-3-OH-ProMarfey’s method (FDAA) combined with HPLCC18 column (4.8 × 250 mm);
MP: ACN: 0.1% aq. TFA buffer (pH 3) (1:9 to 1:1 v/v)
Flow rate: 1.0 mL/min; UV detection at 340 nm
aa—Amino acid; FDAA—1-Fluoro-2-4-dinitrophenyl-5-l-alanine amide; HPLC—High Performance Liquid Chromatography; MP—Mobile Phase; ACN—Acetonitrile; TFA—Trifluoracetic acid.
Table 4. Cyclic peptides from marine-derived fungi.
Table 4. Cyclic peptides from marine-derived fungi.
PeptideSourceaa CompositionChromatographic ConditionsBiological ActivitiesRef.
Cyclo-(l-leucyl-trans-4-hydroxyl-l-prolyl-d-leucyl-trans-4-hydroxy-l-proline) (99)Marine mangrove-derived fungi Phomopsis sp. K38 and Alternaria sp. E334-OH-l-Pro, d-Leu, l-LeuMarfey’s method (FDAA) combined with LC/MS
Alltima C18 column, 5 μm; (4.6 × 250 mm)
MP: MeOH:H2O:HCOOH (60:40:0.05 to 10:90:0.05 v/v/v);
Flow rate: 0.6 mL/min
Inhibition against four crop-threatening fungi [105]
Scytalidamides A (100) and B (101)Marine Fungus of the genus Scytalidiuml-Phe, N-Me-l-Phe, l-Leu, N-Me-l-Leu, l-Pro, 3-Me-l-ProMarfey’s method (FDLA) combined with HPLC
Agilent Hypersil ODS column, 5 μm (4.6 × 100 mm);
MP: ACN 25 to 65%;
Flow rate: 0.7 mL/min
Cytotoxicity against HCT-116 and NCI 60 cell lines [106]
Clonostachysins A (102) and B (103)Marine sponge-derived fungus Clonostachys rogersoniana strain HJK9N-Me-l-Ile, N-Me-l-Leu, l-Pro, l-Gly, N-Me-l-Tyr, N-Me-l-Ala
102: N-Me-l-Val; 103: N-Me-l-Ile
Marfey’s method (FDLA) combined with LC-ESI MS/MS; Conditions not describedInhibitory effect on dinoflagellate Prorocentrum micans[107]
Asperterrestide A (104)Marine-derived fungus Aspergillus terreus SCSGAF0162d-AlaMarfey’s method (FDAA) combined with HPLC
Alltima C18 column, 5 μm (4.6 × 250 mm);
MP: ACN:H2O:TFA (15:85:0.1 to 90:10:0.1 v/v/v);
Flow rate: 0.5 mL/min; UV detection at 254 nm
Cytotoxicity against U937 and MOLT4 human carcinoma cell lines and inhibitory effects on influenza virus [108]
Ligand Exchange Type CSP; MCI GELCRS 10 W (4.6 × 50 mm);
MP: 2 mM CuSO4:H2O solution
Flow rate: 1.0 mL/min; UV detection at 254 nm
Sclerotides A (105) and B (106)Marine-derived fungus, Aspergillus sclerotiorum PT06-1l-Thr, l-Ala, d-Phe, d-SerCrown Ether CSP; Crownpak CR (+);
MP: aq HClO4 pH 2.0;
Flow rate: 0.4 mL/min; UV detection at 200 nm
105 and 106: Antifungal activity
106: Cytotoxicity and antibacterial activity
[109]
Cordyheptapeptides C–E (107109)Marine-derived fungus Acremonium persicinum SCSIO 115N-Me-l-Tyr, l-Phe, l-Pro, l-Leu
107109: N-Me-d-Phe, l-Val
109: N-Me-l-Gly, N-Me-d-Tyr, l-allo-Ile
Crown Ether Chiral CSP; Crownpak CR (+)
MP: 2.0 mM CuSO4:ACN (95:5 v/v)
Flow rate: 1.0 mL/min; UV detection at 254 nm
107 and 109: Cytotoxicity against SF-268, MCF-7, and NCI-460 tumor cell lines[110]
Similanamide (110)Marine sponge-associated fungus Aspergillus similanensis KUFA 0013l-Ala, d-Leu, l-Val, N-Me-l-Leu, d-pipecolic acidMacrocyclic Antibiotic Type CSP; Chirobiotic T, 5 μm (4.6 × 150 mm);
MP: MeOH:H2O:CH3COOH (70:30:0.02 v/v/v);
Flow rate: 1.0 mL/min; UV detection at 210 nm
Cytotoxicity against
MCF-7, NCI-H460 and A373 tumor cell lines
[111]
Sartoryglabramide A (111) and B (112)Marine sponge-associated fungus Neosartorya glabra KUFA 0702l-Phe, l-Pro
112: l-Trp
Macrocyclic Antibiotic Type CSP; Chirobiotic T, 5 μm (4.6 × 150 mm);
MP: MeOH:H2O (80:20 v/v)
Flow rate: 1.0 mL/min; UV detection at 210 nm
Neither antibacterial nor antifungal activity[112]
aa—Amino acid; FDAA—1-Fluoro-2-4-dinitrophenyl-5-l-alanine amide; HPLC—High Performance Liquid Chromatography; MP—Mobile Phase; ACN—Acetonitrile; TFA—Trifluoracetic acid; MeOH—Methanol; FDLA—1-fluoro-2-4-dinitrophenyl-5-d,l-leucine amide.
Table 5. Cyclic depsipeptides from marine-derived fungi.
Table 5. Cyclic depsipeptides from marine-derived fungi.
PeptideSourceaa CompositionChromatographic ConditionsBiological ActivitiesRef.
Exumolides A (113) and B (114)Fungus of the genus Scytalidium sp.l-Pro, l-Phe,
N-Me-l-Leu
Marfey’s method (FDAA) combined with HPLC
Hewlett Packard 1090 Diode Array, 5 µm (10 × 250 mm);
MP: 10–50% aq ACN (0.1% TFA)
Flow rate: 1.0 mL/min; UV detection at 340 nm
Antimicroalgal activity against unicellular chlorophyte Dunaliella sp[113]
Guangomide A (115)Sponge-derived fungusN-Me-d-PheMarfey’s method (FDAA) combined with HPLC
Alltech Altima C18 column, 5 µm (10 × 250 mm)
MP: ACN:H2O (4:1 to 1:1 v/v);
Flow rate: 1.0 mL/min; UV detection at 340 nm
Antibacterial activity against Staphylococcus epidermidis and Enterococcus durans[114]
Destruxin E chlorohydrin (116) and pseudodestruxin C (117)Marine-derived fungus Beauveria felinaN-Me-l-Val
116: N-Me-l-Ala, l-Ile
117: l-Phe
Marfey’s method (FDAA) combined with HPLC
C18 column, 5 µm (4.6 × 250 mm);
MP: 10–20% ACN in 0.1 M NH4OAc (pH = 5)
Flow rate: 1.0 mL/min; UV detection at 340 nm
Cytotoxicity in NCI’s 60 cell line panel[115]
Zygosporamide (118)Marine-derived fungus Zygosporium masoniil-Phe, l-Leu, d-LeuMarfey’s method (FDAA) combined with HPLC
C18, Agilent column, 5 µm (4.6 × 250 mm)
MP: 10–50% ACN (0.1% TFA)
Flow rate: 1.0 mL/min; UV detection at 340 nm
Cytotoxicity in RXF 393 and SF-268 cancer cell lines[116]
Petriellin A (119)Coprophilous fungus Petriella sordidaN-Me-l-Ile,
N-Me-l-Thr
d-Phenyllactate
Marfey’s method (FDAA) combined with HPLC
C18 column (4.6 × 250 mm); Conditions not described; UV detection at 260 nm
Antifungal activity[117]
Alternaramide (120)Marine derived fungus Alternaria sp. SF-5016l-Pro, d-PheMarfey’s method (FDAA) combined with HPLC
Capcell Pak C18 column; MP: 30–60% ACN in H2O (0.1% HCOOH); Flow rate: 1.0 mL/min
Antibacterial activity against Bacillus subtilis and Staphylococcus aureus[118]
Petrosifungins A (121) and B (122)Penicillum brevicompac-tuml-Val, l-Pro, l-Thr, l-pipecolinic acidsMarfey’s method (FDAA) combined with HPLC
C18 column, Waters, 5 µm (2.1 × 150 mm);
MP: H2O or ACN (0.05% TFA);
Flow rate: 1.0 mL/min
Not described[119]
Oryzamides A–E (123127)Sponge-Derived fungus Nigrospora oryzae PF18l-Ala, d-Leu, l-Val
123: l-Leu; 124: l-Tyr
125 and 126: l-Met; 127: l-Phe
Marfey’s method (FDLA) combined with UHPLC
Acquity UHPLC BEH column, 1.7 µm (2.1 × 250 mm);
MP: 10–100% ACN in H2O with 0.1% HCOOH;
Flow rate: 0.5 mL/min; UV detection at 360 nm
No cytotoxicity, antibacterial, antiparasitic, and NF-kB activities[122]
Spicellamide A (128) and B (129)Marine-derived fungus Spicellum roseumN-Me-d-Phe,
N-Me-l-Ala, l-Ala
Marfey’s method (FDAA) combined with HPLC
C18 column; Macherey-Nagel Nucleodur 100, 5 µm (2.0 × 125 mm);
MP: MeOH:H2O (10:90 v/v to 100% MeOH) or 100% MeOH with NH4Ac, 2 mmol
129: Cytotoxicity [123]
l-2-hydroxyisocaproic acidLigand Exchange Type CSP; Phenomenex Chirex 3126 N,S-dioctyl-(d)-penicillamine (4.6 × 50 mm)
MP: 2 mM CuSO4 in ACN:H2O (15:85 v/v)
Flow rate: 1.0 mL/min; UV detection at 254 nm
Depsipeptides 1962A (130) and 1962B (131)Endophytic fungus Kandelia candell-Tyr, l-Val, d-Leu, (S)-O-LeuCrown Ether CSP; Crownpak CR (+) column (0.4 × 150 mm),
MP: 2 mM CuSO4 aq. solutions
Flow rate: 0.5 mL/min; UV detection at 200 nm
131: Activity against MCF-7 tumor cell line[124]
aa—Amino acid; FDAA—1-Fluoro-2-4-dinitrophenyl-5-l-alanine amide; HPLC—High Performance Liquid Chromatography; MP—Mobile Phase; ACN—Acetonitrile; TFA—Trifluoracetic acid; MeOH—Methanol; TEA—Triethylamine; FDLA—1-Fluoro-2-4-dinitrophenyl-5-d,l-leucine amide; NaOAc—Sodium acetate; NH4OAc—Ammonium acetate.
Table 6. Cyclic peptides from marine sponges.
Table 6. Cyclic peptides from marine sponges.
PeptideSourceaa CompositionChromatographic ConditionsBiological ActivitiesRef.
Haliclamide (132)Vanuatu marine sponge Haliclona sp.N-Me-l-PheMarfey’s method (FDAA) combined with HPLC
Vydac C18;
MP: ACN in H2O with 0.1% TFA (9:1 to 1:1 v/v); UV detection at 340 nm
Cytotoxicity against NSCLC-N6 carcinoma cell line[129]
Microsclerodermins J (133) and K (134)Deep water sponge Microscleroderma herdmanil-Ile, l-Thr
133: l-Phe, l-Gly
134: l-Val, l-Ala
Marfey’s method (FDAA) combined with HPLC
C18 column, 5 µm (4.6 × 150 mm)
Flow rate: 1.0 mL/min; UV detection at 340 nm
Activity against opportunistic pathogenenic fungi[130]
Euryjanicins E–G (135137)The Caribbean Sponge Prosuberites laughlinil-Pro, l-Ile, l-Phe
135: l-Asp
Marfey’s method (FDAA) combined with HPLC
C18 column, 5 µm (4.6 × 150 mm)
Flow rate: 1.0 mL/min; UV detection at 340 nm
No significant activity cytotoxicity against the National Cancer Institute 60 tumor cell line panel[131]
Chujamide A (138)Marine sponge Suberites waedoensisl-Pro, l-Tyr, l-Cys,
l-Leu, l-Phe
l-Ile (S)
Marfey’s method (FDAA) combined with HPLC
ESI-LC/MS YMC ODS-A column, 5 µm (4.6 × 250 mm)
MP: H2O:ACN (80:20 to 30:70 v/v)
Flow rate: 0.7 mL/min; UV detection at 360 nm
Weak cytotoxicity against A549 and K562 cell lines[132]
Kapakahines A–D (139142)Marine Sponge Cribrochalina olemdal-Val, l-Ile, l-Leu,
l-Trp, l-Phe, l-Ala,
l-Pro, l-Try
Marfey’s method (FDAA) combined with HPLC
Cosmosil C18-MS column, 5 µm (4.6 × 250 mm)
MP 37.5% ACN in 0.05% TFA or 20% or 38% ACN in 50 mM NH4OAc
139141: Cytotoxicity against P388 cell line
139: Inhibition against protein phosphatase
[133]
Koshikamide B (143)Marine sponge Theonella sp.d-Phe, l-Thr,
N-Me-l-Val, N-Me-l-Asn, N-Me-l-Leu
Marfey’s method (FDAA) combined with HPLC
ODS HPLC (10 × 250 mm);
MP: ACN:H2O:TFA (25:75:0.05 to 55:45:0.05 v/v/v);
Flow rate: 1.0 mL/min; UV detection at 340 nm
Cytotoxicity against P388 and HCT-116 tumor cell lines[134]
Perthamides C (144) and D (145)Solomon Lithistid sponge Theonella swinhoeil-Asp, l-ThrOMe, (2R,3S)-βOHAsp, l-PheMarfey’s method (FDAA) combined with HPLC/MS Proteo C18 column (1.8 × 25 mm)
MP: 10–50% aq ACN with 5% HCOOH and 0.05% TFA
Flow rate: 0.15 mL/min
Anti-inflammatory activity[135]
Perthamides E (146) and F (147)Polar extracts of the sponge Theonella swinhoei146: l-ThrOMe
147: l-Phe
Marfey’s method (FDAA) combined with HPLC
Proteo C18 column, (1.8 × 25 mm);
MP: 10–50% aq ACN with 5% HCOOH and 0.05% TFA
Flow rate: 0.15 mL/min
147: IL-8 release inhibition[136]
Stylisins 1 (148) and 2 (149)Jamaican sponge Stylissa caribical-Pro, l-Tyr, l-Ile
148: l-Leu, l-Phe
Marfey’s method (FDAA) combined with HPLC
HPLC water Nova Pack column (3.9 × 150 mm)
MP: TEAP buffer (pH 3.0 ± 0.02):ACN (90 to 60% TEAP) UV detection at 340 nm
No antimicrobial, antimalarial, anticancer, anti-HIV-1, anti-Mtb and anti-inflammatory activities[137]
Carteritins A (150) and B (151)Marine sponge Stylissa carteri150: l-Pro, l-Phe, l-Ile, l-Pro (trans), l-Pro (cis), l-Glu, l-Tyr; 151: l-Pro (trans), l-Leu,
l-Tyr, l-Pro(cis)
Marfey’s method (FDAA) combined with HPLC
Cosmosil C18 MS (4.6 × 250 mm);
MP: H2O:TFA (100:0.1) to ACN:H2O:TFA (50:50:0.1 v/v/v);
Flow rate: 1.0 mL/min; UV detection at 340 nm
150: Cytotoxicity against HeLa, HCT116, and RAW264 cells [139]
Stylissatins B–D (152154)Marine sponge Stylissa massal-Pro, l-Phe, l-Leu
152: l-His
153154: l-Asp, l-Val
Marfey’s method (FDAA) combined with HPLC
Thermo BDS Hypersil C18 column, 5 μm (4.6 × 150 mm);
MP: 30–70% MeOH:H2O (H3PO4)
Flow rate: 1.0 mL/min; UV detection at 340 nm
152: Inhibitory effects against a panel of human tumor cell lines including HCT-116, HepG2, BGC-823, NCI-H1650, A2780, and MCF7[138]
Callyaerin G (155)Indonesian sponge Callyspongia aerizusal-Pro, l-Leu, l-Phe,
l-FGly
Marfey’s method (FDAA) combined with HPLC/MS; Conditions not describedCytotoxicity against L5178Y, Hela, and PC12[140]
Reniochalistatins A–E (156160)Marine sponge Reniochalina stalagmitisl-Pro, l-Phe, l-Val,
l-Leu, l-Ile, l-Tyr
Ligand Exchange Type CSP; MCI GELCRS 10 W (4.6 × 50 mm);
MP: 2 mM CuSO4:H2O solution
Flow rate: 1.0 mL/min; UV detection at 254 nm
160: Cytotoxicity against RPMI-8226, MGC-803, HL-60, HepG2, and HeLa [141]
156: l-Asn
160: l-Trp
Marfey’s method (FDAA) combined with HPLC YMC-Park Pro C18, 5 µm (4.6 × 250 mm)
MP: 2 mM CuSO4:H2O solution
Flow rate: 1.0 mL/min; UV detection at 254 nm
Phakellistatins 15–18 (161164)South china sea sponge Phakellia fuscal-Pro
161: l-Trp, l-Ile, l-Leu, l-Thr; 162: l-Phe, l-Asp, l-Ser, l-Arg, l-Ala, l-Val, l-Thr, l-Tyr; 163: l-Trp, l-Val, l-Leu, l-Ile; 164: l-Tyr, l-Ile, l-Phe
Ligand-exchange type CSP; Chirex 3126 (d)-penicillamine column (4.6 × 150 mm)
MP: aq 2 mM CuSO4:MeOH (85:15 to 70:30 v/v) or aq 1 mM/0.5 mM CuSO4;
Flow rate: 0.5 or 1.0 mL/min
161: Cytotoxicity against P388 cancer cell line
162: Cytotoxicity against P388 and BEL-7402 cancer cell lines
[142]
aa—Amino acid; FDAA—1-Fluoro-2-4-dinitrophenyl-5-l-alanine amide; LC—Liquid Chromatography; MS—Mass spectrometry; HPLC—High Performance Liquid Chromatography; MP—Mobile Phase; TEAP—Triethylammonium phosphate; ACN—Acetonitrile; TFA—Trifluoracetic acid; MeOH—Methanol; TEA—Triethylamine; IPA—Isopropyl alcohol; NaOAc—Sodium acetate; NH4OAc—Ammonium acetate.
Table 7. Cyclic depsipeptides from marine sponges.
Table 7. Cyclic depsipeptides from marine sponges.
PeptideSourceaa CompositionChromatographic conditionsBiological activitiesRef.
Callipeltins B (165) and C (166)Callipelta sp.l-Ala, d-Arg, l-Thr,
N-Me-l-Ala, l-Leu
Marfey’s method (FDAA) combined with HPLC; Column not described; MP: TEAP (50 nM, pH 3.0):ACN 90–50% TEAP
Flow rate: 2.0 mL/min; UV detection at 340 nm
Cytotoxicity
166: Growth inhibitory activity against Candida albicans
[143]
Halipeptiins A (167) and B (168)Haliclona speciesl-AlaMarfey’s method (FDAA) combined with HPLC
Vydac C18 column; MP: H2O (0.1% TFA):ACN (0:1 to 1:1 v/v)
UV detection at 340 nm
168: Anti-inflammatory activity [144]
Phoriospongin A (169) and B (170)Phoriospongia sp. and Callyspongia bilamellatad-Asp, d-allo-Thr, d-Ala, l-Phe,
d-Leu, d-nor-Val,
N-Me-d-nor-Val
170: N-Me-l-Leu
Marfey’s method (FDAA) combined with HPLC
C18 column, 5 µm (4.6 × 250 mm)
Flow rate: 1.0 mL/min; UV detection at 340 nm
Nematocidal activity against the parasite Haemonchus contortus[145]
Mirabamides A–D (171174)Siliquarias-pongia mirabilisN-Me-l-Thr, l-Thr, l-Ala, d-3-OMeAla, (2R,3R)-3-OH-Leu (3S,4R)-diMe-l-Glu, (2S,3R)-diaminobutanoic acid; 174: l-HPrMarfey’s method (FDAA) combined with HPLC
Phenomenex Jupiter Proteo C12 column, 4 µm (4.6 × 150 mm)
MP: 25–70% ACN; Flow rate: 0.5 mL/min
171: Anti-HIV activity
173 and 174: Antibacterial activity
171173: Antifungal activity
[146]
Neamphamides B (175), C (176) and D (177)Neamphius huxleyid-Arg, l-Asn, l-Hpr, l-Leu, d-allo-Thr
175 and 177: N-Me-l-Gln
176: N-Me-l-Glu
Marfey’s method (FDAA) combined with HPLC
Phenomenex Luna Column C18, 3 µm (2.0 × 150 mm)
MP: H2O:ACN:HCOOH (100:0:0.1 to 0:100:0.1 v/v/v)
UV detection at 340 nm
Growth inhibition of human cell lines: A549, HeLa, LNCaP, PC3, and NFF[147]
Pipecolidepsins A (178) and B (179)Homophymia lamellosad-Asp, l-Leu, d-Lys, d-allo-Thr, (3S,4R) diMe-l-Glu, (2S,3S)-EtO-Asp, N-Me-l-Glu, l-PipMarfey’s method (FDAA) combined with HPLC
Symmetry C18, 5 µm (4.6 × 150 mm); MP: 20–50% ACN (0.04% TFA) in H2O (0.04% TFA); Flow rate: 0.8 mL/min
Cytotoxicity against three human tumor cell lines (A-549, HT-29, and MDA-MB-231)[148]
Stellatolide A (180)Ecionemia acervusN-Me-l-Ala, l-Leu,
N-Me-l-Gln, N-Me-d-Ser,
d-allo-Thr
Marfey’s method (FDAA) combined with HPLC Hewlett-Packard Hypersil BDS-C18, 4 µm (4.0 × 100 mm); MP: H2O (0.1% TFA):ACN (90:10 to 50:50 v/v); Flow rate: 1.0 mL/minIn in vitro antiproliferative activity[149]
Cyclolithistide A (181)Theonella swinhoeinor-S-Val, S-Phe, S-Gln,
N-Me-S-Leu, S-Ala,
S-Allo-S-Thr
Marfey’s method (FDAA) combined with HPLC
ODS (4.6 × 250 mm); MP: 100% H2O
Flow rate: 2.0 mL/min; UV detection at 210 nm
Antifungal activity against Candida albicans (ATCC 24433)[150]
Nagahamide A (182)Theonella swinhoeil-Val, l-Ser, 3S-AHBAMarfey’s method (FDAA) combined with HPLC
ODS column (4.6 × 250 mm); Conditions not described
Antibacterial activity[151]
Theopapuamides B (183) and C (184), Celebesides A–C (185187)Siliquarias-pongia mirabilis185: l-βMeAsnMarfey’s method (FDAA) combined with HPLC/MS
Phenomenex Jupiter Proteo C12 column, 4 µm (4.6 × 150 mm)
MP: 25–70% ACN with 0.01 M TFA; Flow rate: 0.5 mL/min
185: Inhibits HIV-1 Entry
183185: Cytotoxic to human colon tumor cell line (HCT-116)
183 and 185: Antifungal activity against Candida albicans
[152]
Ligand Exchange Type CSP Phenomenex column, Chirex Phase 3126 (D) (4.6 × 150 mm); MP: 1 mM CuSO4:ACN (95:5 v/v)Flow rate: 0.5 mL/min; UV detection at 254 nm
Theopapuamide (188)Lithistid sponge Theonella swinhoeiN-Me-l-Leu, d-Asp, l-Leu, N-Me-l-GluLigand Exchange Type CSP Chirex Phase 3126 (D), 5 µm
(4.6 × 250 mm); MP: IPA: 2 mM CuSO4 (5:95 v/v)
Flow rate: 1.0 mL/min; UV detection at 254 nm
Cytotoxicity against CEM-TART and HCT-cell lines[153]
d-allo-ThrMarfey’s method (FDAA) combined with HPLCPhenomenex C18, 5 µm (4.6 × 250 mm); MP: 10–50% ACN in H2O (0.05% TFA); Flow rate: 1.0 mL/min; UV detection at 340 nm
Mutremdamide A (189) and Koshikamides C–H (190195)Theonella swinhoei and Theonella cupola189: N-Me-l-Val; 190: N-Me-l-Val, N-Me-l-Asn, l-Asn, N-Me-l-Leu, l-Pro, N-Me-allo-l-Ile, d-PheMarfey’s method (FDAA) combined with HPLC
LC-MS analysis using a C12 column, 4 µm (4.6 × 250 mm); MP: ACN with 0.01% TFA; Flow rate: 0.5 mL/min
189195: Anti-HIV-1 activity [154]
191 and 192: N-Me-allo-l-Ile,
N-Me-l-Val; 192194: N-Me-allo-l-Ile, l-Ala 1, d-Ala2, l-Asn
Marfey’s method (FDAA) combined with HPLC
LC-MS, C18 column, 4 µm (4.6 × 250 mm); MP: 20 mM buffer (AF):ACN (3:1 to 3:7 v/v); Flow rate: 0.5 mL/min
195: N-Me-allo-l-IleChiral HPLC (column not described); MP: 1 mM CuSO4:ACN (95:5 v/v); Flow rate: 0.5 mL/min; UV detection at 254 nm
aa—Amino acid; FDAA—1-Fluoro-2-4-dinitrophenyl-5-l-alanine amide; LC—Liquid Chromatography; MS—Mass spectrometry; HPLC—High Performance Liquid Chromatography; MP—Mobile Phase; TEAP—Triethylammonium phosphate; ACN—Acetonitrile; TFA—Trifluoracetic acid; MeOH—Methanol; TEA—Triethylamine; NaOAc—Sodium acetate; NH4OAc—Ammonium acetate.
Table 8. Lipopeptides from marine sponge.
Table 8. Lipopeptides from marine sponge.
PeptideSourceaa CompositionChromatographic ConditionsBiological ActivityRef.
Sulfolipo-discamides A–C
(196198)
Sponge
Discoderma kiiensis
l-Uda, l-GlyMarfey’s method (FDAA) combined with HPLC Cosmosil C18-MSII column (4.6 × 250 mm); MP: 100 mM NaClO4 in 60% ACN
Flow rate: 0.8 mL/min
196: Cytotoxicity against P388 cell line[155]
aa—Amino acid; FDAA—1-Fluoro-2-4-dinitrophenyl-5-l-alanine amide; HPLC—High Performance Liquid Chromatography; MP—Mobile Phase; ACN—Acetonitrile.
Table 9. Cyclic peptides from marine invertebrates and algae.
Table 9. Cyclic peptides from marine invertebrates and algae.
PeptideSourceaa CompositionChromatographic ConditionsBiological ActivitiesRef.
Didomolamides A (199) and B (200)Ascidian Didemnum mollel-Thr, l-Ala, l-Phe
200: l-Tzl
Marfey’s method (FDAA) combined with HPLC; MP: 50 mM (TEAP) buffer pH 3: ACN (9:1 to 1:1 v/v); Flow rate: 1.0 mL/min; UV detection at 340 nmCytotoxicity against A549, HT29 MEL28 tumor cell lines [157]
Mollamides B (201) and C (202)Tunicate Didemnum mollel-Thr, l-Ile, l-Pro
201: l-Val, l-Phe
202: l-Ser, l-Leu
Marfey’s method (FDAA) combined with HPLC; MP: 50 mM TEAP, pH 3.0: ACN (90:10 to 60:40 v/v) or 40 mM NH4OAc, 70% ACN, and 30% MeOH (98:2 to 66:34 v/v)
Flow rate: 1.0 or 0.8 mL/min; UV detection at 340 nm
201: Activity against HIV, Plasmodium falciparum, Lieshmania donovan, and cytotoxicity against H460, MCF7, SF-268 cell lines[158]
Antatollamides A (203) and B (204)Ascidian Didemnum-mollel-Ile, l-Phe, l-Val, l-Pro, d-AlaMarfey’s method (FDLA) combined with HPLC/MSHypersil Gold C18 column, 1.9 µm (2.1 × 50 mm); MP: H2O 0.1%; HCOOH:ACN (85:15 to 55:45 v/v)
Flow rate: 0.5 mL/min
203: Weak cytotoxicity against a chronic lymphocytic leukemia cell line[159]
Sanguinamide A (205)Nudibranch Hexabranchs sanguineusl-Pro, l-Ile, l-Ala, l-PheMarfey’s method (FDLA) combined with HPLC Agilent Zorbax SB-Aq C18 column, 5 µm (4.6 × 250 mm)
MP: 80% (H2O: 0.1% HCOOH), 20% (ACN)
Antifungal activity[160]
Gamakamide E (206)Oysters Crassostrea gigal-Met(O),
N-Me-l-Phe, l-Leu, d-Lys, l-Phe
Marfey’s method (FDLA) combined with HPLC
Conditions not described
No growth inhibition abilities[161]
aa—Amino acid; FDAA—1-Fluoro-2-4-dinitrophenyl-5-l-alanine amide; HPLC—High Performance Liquid Chromatography; MP—Mobile Phase; ACN—Acetonitrile; TEAP—Triethylammonium phosphate; FDLA—1-Fluoro-2-4-dinitrophenyl-5-d,l-leucine amide; TFA—Trifluoracetic acid; MeOH—Methanol; TEA—Triethylamine; NH4OAc—Ammonium acetate.
Table 10. Cyclic depsipeptides from marine invertebrates and algae.
Table 10. Cyclic depsipeptides from marine invertebrates and algae.
PeptideSourceaa CompositionChromatographic ConditionsBiological ActivitiesRef.
Kahalalides A–F (207212)Mollusk Elysia rufescens207: d-Val-5; 208: l-Val-1, d-Val-2, d-allo-Thr-1; 209: l-Val-3, d-Val-4, l-Thr-2; 210: d-Val-2, d-allo-Thr-1Marfey’s method (FDLA) combined with HPLC
COSMOSIL 5C18-AR
MP: ACN:H2O:TFA (42:48:0.05 v/v/v) or ACN:H2O:50 mM NH4OAc (20:80:0.01 v/v/v)
207: Antimalarial activity
211: Activity against RSV II virus
[162]
Table 10. Cont.Tamandarins A (213) and B (214)Ascidian of the family Didemni-dae213: S-Lac, l-Pro, N-Me-d-Leu, l-Thr, (3S,4R,5S)-Ist
214: S-Lac, l-Pro, N-Me-d-Leu, l-Thr 3S,4R)-Nst
Marfey’s method (FDAA) combined with HPLC
Hewlett-Packard ODS Hypersil 5 µm (4.6 × 200 mm); MP: 0.1% TFA in H2O or MeOH; Flow rate: 1.0 mL/min; UV detection at 340 nm
213: Cytotoxicity against various human cancer cell lines[163]
KahalalidesP
(215) and Q (216)
Green alga Bryopsis sp.l-Asp, l-Val,
d-Leu, l-Ser,
l-Hyp, l-Pro,
l-Lys
Marfey’s method (FDAA) combined with HPLC COSMOSIL 5C18-AR-II (4.6 × 250 mm); MP: 0.1 M NH4OAc pH 3 or 90% aq ACNNo antimicrobial and no hemolytic activities[164]
Kahalalide O (217)Mollusk Elysia ornata and green alga Bryopsis sp.l-Ile, l-Thr, d-allo-Thr, d-Tyr,
l-Val
Ligand Exchange Type CSP Chirex (D) Penicillamine Column (4.6 × 250 mm); MP: 1.9 mM CuSO4 in ACN:H2O (5:95) or 2.0 mM CuSO4 in H2O; UV detection at 254 nmNo growth inhibition of P-388, A549, HT29 and MEL28 cancer cell lines[165]
d-TrpMarfey’s method (FDAA) combined with HPLC
COSMOSIL 5C18-AR; MP: ACN:H2O:TFA (37.5:62.5:0.05 v/v/v); Flow rate: 1.0 mL/min
UV detection at 254 nm
aa—Amino acid; FDAA—1-Fluoro-2-4-dinitrophenyl-5-l-alanine amide; HPLC—High Performance Liquid Chromatography; MP—Mobile Phase; ACN—Acetonitrile; FDLA—1-fluoro-2-4-dinitrophenyl-5-d,l-leucine amide; TFA—Trifluoracetic acid; MeOH—Methanol; TEA—Triethylamine; NH4OAc—Ammonium acetate.
Table 11. Lipopeptides from marine invertebrates and algae.
Table 11. Lipopeptides from marine invertebrates and algae.
PeptideSourceaa CompositionChromatographic ConditionsBiological ActivitiesRef.
Eudistomides A (218) and B (219)Ascidian Eudistoma sp.l-Pro, l-Ala, l-Leu
219: l-Cyp
Ligand Exchange Type CSP
Phenomenex Chirex 3126 (D) (4.6 × 250 mm); MP: 2 mM CuSO4, 2 mM CuSO4:ACN (95:5 or 85:15 v/v); Flow rate: 1.0 mL/min
UV detection at 254 nm
No activity reported[166]
Mebamamides A (220) and B (221)Green algae Derbesia marinal-Leu, l-Pro, d-Ala, l-Thr, l-Val, d-Phe, d-SerLigand Exchange Type CSP
Diacel CHIRALPAK (MA+) (4.6 × 50 mm); MP: 2.0 mM CuSO4, Flow rate: 1.0 mL/min; UV detection at 254 nm
No growth inhibitory activity against HeLa and HL60 cell lines[167]
aa—Amino acid; MP—Mobile Phase; ACN—Acetonitrile.
Table 12. Chiral HPLC analysis of the acidic hydrolysates of 110, 111 and 112 by co-injection with amino acids standards.
Table 12. Chiral HPLC analysis of the acidic hydrolysates of 110, 111 and 112 by co-injection with amino acids standards.
Retention Time (min) Retention Time (min)
d-Trp (A)5.20d-pipecolic acid (B)14.67
l-Trp (A)4.51Acidic hydrolysate of 110 (B)6.59, 7.20, 8.09, 8.83, 9.67, 10.57, 14.69
l-Val (B)6.60Acidic hydrolysate of 110 +
dl-Val (co-injection) (B)
6.61, 7.31, 8.30, 8.10, 8.84, 9.70, 10.50, 14.95
d-Val (B)8.32Acidic hydrolysate of 110 +
dl-Ala (co-injection) (B)
6.59, 7.19, 8.04, 8.81, 9.37, 9.70, 10.50, 14.90
l-Ala (B)7.16Acidic hydrolysate of 110+
dl-Leu (co-injection) (B)
6.60, 6.76, 7.26, 8.04, 8.83, 9.67, 10.54, 15.02
d-Ala (B)9.36Acidic hydrolysate of 110 +
dl-pipecolic acid (co-injection) (B)
6.58, 7.20, 8.09, 8.64, 8.84, 9.77, 10.64, 14.64
l-Leu (B)6.78Acidic hydrolysate of 110 +
N-Me-l-Leu (co-injection) (B)
6.59, 7.20, 8.09, 8.83, 9.67, 10.57, 14.69
d-Leu (B)9.67Acidic hydrolysate of 111 (A)1.91, 2.55, 2.86, 3.49, 3.89, 6.79
N-Me-l-Leu (B)8.09Acidic hydrolysate of 111 +
dl-Phe (co-injection) (A)
1.87, 2.50, 2.89, 3.68, 5.01, 6.82
l-Phe (A)3.81Acidic hydrolysate of 111 +
dl-Pro (co-injection) (A)
1.96, 2.60, 2.96, 3.52, 3,92, 6.70, 21.09
d-Phe (A)5.00Acidic hydrolysate of 112 (A)1.93, 3.07, 3.80, 4.29, 4.60, 6.62
l-Pro (A)6.72Acidic hydrolysate of 112 +
dl-Phe (co-injection) (A)
1.90, 3.10, 3.78, 4.39, 5.04, 6.70
d-Pro (A)20.10Acidic hydrolysate of 112 +
dl-Pro (co-injection) (A)
2.04, 3.02, 3.72, 4.30, 4.60, 6.66, 19.40
l-pipecolic acid (B)8.68Acidic hydrolysate of 112 +
dl-Trp (co-injection) (A)
1.93, 2.99, 3.70, 4.29, 4.60, 5.07, 6.33
Column, Chirobiotic T; Mobile phase, MeOH:H2O (80:20 v/v) (A) or MeOH:H2O:acetic acid (70:30:0.02 v/v/v) (B); Flow rate, 1.0 mL/min (A) or 0.5 mL/min (B); UV detection, 210 nm.

Share and Cite

MDPI and ACS Style

Phyo, Y.Z.; Ribeiro, J.; Fernandes, C.; Kijjoa, A.; Pinto, M.M.M. Marine Natural Peptides: Determination of Absolute Configuration Using Liquid Chromatography Methods and Evaluation of Bioactivities. Molecules 2018, 23, 306. https://doi.org/10.3390/molecules23020306

AMA Style

Phyo YZ, Ribeiro J, Fernandes C, Kijjoa A, Pinto MMM. Marine Natural Peptides: Determination of Absolute Configuration Using Liquid Chromatography Methods and Evaluation of Bioactivities. Molecules. 2018; 23(2):306. https://doi.org/10.3390/molecules23020306

Chicago/Turabian Style

Phyo, Ye’ Zaw, João Ribeiro, Carla Fernandes, Anake Kijjoa, and Madalena M. M. Pinto. 2018. "Marine Natural Peptides: Determination of Absolute Configuration Using Liquid Chromatography Methods and Evaluation of Bioactivities" Molecules 23, no. 2: 306. https://doi.org/10.3390/molecules23020306

Article Metrics

Back to TopTop