ArticlePDF AvailableLiterature Review

Marine peptides in breast cancer: Therapeutic and mechanistic understanding

Authors:

Abstract and Figures

Breast cancer is the most prevalent invasive form of cancer in females and posing a great challenge for overcoming the disease burden. The growth in global cancer deaths mandates the discovery of new efficacious natural anti-tumor treatments. In this regard, aquatic species offer a rich supply of possible drugs. Studies have shown that several marine peptides damage cancer cells by a broad range of pathways, including apoptosis, microtubule balance disturbances, and suppression of angiogenesis. Traditional chemotherapeutic agents are characterized by a plethora of side effects, including immune response suppression. The discovery of novel putative anti-cancer peptides with lesser toxicity is therefore necessary and timely, especially those able to thwart multidrug resistance (MDR). This review addresses marine anti-cancer peptides for the treatment of breast cancer.
Content may be subject to copyright.
Biomedicine & Pharmacotherapy 142 (2021) 112038
Available online 16 August 2021
0753-3322/© 2021 Published by Elsevier Masson SAS. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Marine peptides in breast cancer: Therapeutic and
mechanistic understanding
Salman Ahmed
a
, Hamed Mirzaei
b
, Michael Aschner
c
, Ajmal Khan
d
, Ahmed Al-Harrasi
d
,
*
,
Haroon Khan
e
,
*
a
Department of Pharmacognosy, Faculty of Pharmacy and Pharmaceutical Sciences, University of Karachi, Karachi 75270, Pakistan
b
Research Center for Biochemistry and Nutrition in Metabolic Diseases, Institute for Basic Sciences, Kashan University of Medical Sciences, Kashan, Iran
c
Department of Molecular Pharmacology, Albert Einstein College of Medicine, Bronx, NY 10461, USA
d
Natural and Medical Sciences Research Center, University of Nizwa, P.O Box 33, Postal Code, 616, Birkat Al Mauz, Nizwa, Oman
e
Department of Pharmacy, Abdul Wali Khan University Mardan, 23200, Pakistan
ARTICLE INFO
Keywords:
Marine peptides
Protein hydrolysate
Apoptosis
Metastasis
Cell cycle arrest
ABSTRACT
Breast cancer is the most prevalent invasive form of cancer in females and posing a great challenge for over-
coming disease burden. The growth in global cancer deaths mandates the discovery of new efcacious natural
anti-tumor treatments. In this regard, aquatic species offer a rich supply of possible drugs. Studies have shown
that several marine peptides damage cancer cells by a broad range of pathways, including apoptosis, microtubule
balance disturbances, and suppression of angiogenesis. Traditional chemotherapeutic agents are characterized by
a plethora of side effects, including immune response suppression. The discovery of novel putative anti-cancer
peptides with lesser toxicity is therefore necessary and timely, especially those able to thwart multi drug
resistance (MDR). This review addresses marine anti-cancer peptides for the treatment of breast cancer.
1. Introduction
Breast cancer is among the most prevalent cancers in women.
Around 0.62 million deaths in 2018 were caused by breast cancer, ac-
cording to Globocan 2018 gures from the International Organization
for Research on Cancer (I.A.R.C.). At the present rate, the number of
incidents is forecasted to increase to 3.05 million, and the death toll will
approximate 7 million by 2040 [1]. Cancer is conventionally treated
with chemotherapy, radiation, and surgery [2]. Conventional cancer
chemotherapy has many side effects, and it often targets multiple or-
gans. It is subject to MDR caused by over-expression of membrane
transporters, which may expel intracellular anti-cancer drugs, thus
decreasing drug accumulation and efcacy [3]. Furthermore, radiation
therapy and surgery increase the probability of cancer invasion [2].
Accordingly, developing new efcacious anti-cancer drugs is required
[4,5].
Extensive development initiatives have been underway to acquire
effective compounds of natural origin [6,7]. Approximately 71% of
Earths atmosphere is aquatic, rendering it an enormous reservoir of
novel bioactive substances of rare and special chemical characteristics.
Sea species are a treasure trove of anti-cancer drugs. Over the last
decade, numerous studies have shown that marine products could act as
anti-tumor agents and play a preventive role in tumor management,
underlying their promise in the discovery of novel and efcient phar-
maceuticals [4].
More than 50% of the FDA-approved medications in the 1980s and
Abbreviations: Akt, Serine / threonine protein-kinase family protein-kinase B; APAF-1, Apoptotic Peptidase-Activating Factor-1; BAD, BCl-2 / Bcl-X associated
death-domain protein; BAK, BCl-2 homologous antagonist - killer protein; BAX, BCl-2-associated x protein; Bcl-2, B-Cell lymphoma 2; Bcl-xL, B-Cell lymphoma-extra
large; Casp, Caspase; CXCR4, C-X-C chemokine-receptor type-4; Cyt c, cytochrome c; ErbB3, V-erb-b2 Erythroblastic Leukemia Viral Oncogene Homolog 3 (avian);
ERK1/2, Extracellular Signal Regulated kinase ½; FoxO3a, Forkhead Box Protein O3; HIF1
α
, Hypoxia inducible factor 1alpha; MMP, matrix metalloproteinase; Mcl-1,
Myeloid Cell Leukemia-1; MDR, Multidrug Resistance; MRP1, Multidrug Resistance-associated Protein 1; PARP, Poly ADP-Ribose Polymerase; P-gp, P-glycoprotein;
PI3K, Phosphatidylinositol-3-Kinase; ROS, Reactive Oxygen Species; STAT3, Signal transducer and activator of transcription 3; TNBC, Triple-negative Breast Cancer
Cell; VEGF, Vascular Endothelial Growth Factor.
* Corresponding authors.
E-mail addresses: salmanahmed@uok.edu.pk (S. Ahmed), h.mirzaei2002@gmail.com (H. Mirzaei), michael.aschner@einsteinmed.org (M. Aschner), aharrasi@
unizwa.edu.om (A. Al-Harrasi), haroonkhan@awkum.edu.pk, hkdr2006@gmail.com (H. Khan).
Contents lists available at ScienceDirect
Biomedicine & Pharmacotherapy
journal homepage: www.elsevier.com/locate/biopha
https://doi.org/10.1016/j.biopha.2021.112038
Received 19 June 2021; Received in revised form 1 August 2021; Accepted 7 August 2021
Biomedicine & Pharmacotherapy 142 (2021) 112038
2
1990s have aquatic life origin. Indeed, sea-derived medicines are a
signicant source of anti-cancer drugs [8]. Several marine-derived drugs
have been approved as anti-cancer drugs, since the original approval of
cytarabine in 1969. The discovery of Didemnin B, obtained from Tridi-
demnum solidum in 1981 and dolastatin 10 from Dolabella auricularia in
1987, represents a critical point in the synthesis of marine-derived
cancer chemotherapy peptides [9]. Bioactive peptides from natural
aquatic products, including small marine peptides [10], have shown
efcacy as potential drug candidates with distinct modulations of
several molecular pathways [11].
There are numerous reasons as to why marine peptides have drawn
attention in the search for anti-cancer drugs. They have some signicant
benets over proteins or antibodies, such as small size, simple
manufacturing, readily modied, cell membrane-crossing capability,
low drug-drug interaction, chemical, and biological versatility. An
additional value is the fewer number of adverse side effects due to lack
of accumulation in the kidneys or liver [1215]. The current review is
predominantly focused on the underlying mechanisms of marine pep-
tides in the treatment of breast cancer coupled with their therapeutic
potential.
2. Marine anti-cancer peptides
Marine peptides represent one of the most versatile sources of ther-
apeutically effective drug molecules [16]. Marine anti-cancer peptides
have been extracted from cyanobacteria, sponges, mollusks, ascidians,
algae, fungi, bacteria (actinomycete and streptomyces), and protein
hydrolysates from sh, amphibians, crocodiles, and turtles. Marine
peptides can be categorized as linear and cyclic peptides. Linear peptides
are formed by a straight amino acid chain connected with amide bonds
[17]. Hemiasterlin A-B (tripeptides) [18], Belamide A (tetrapeptides)
[19], Symplostatin, Dolastatin 10 and 15 (pentapeptides) [20] are
reported from sponges, cyanobacteria and mollusks.
Cyclic penta, hexa, hepta, decapeptides, depsipetides and oligopep-
tides from marine organisms are reported to possess anti-breast cancer
properties [17]. Galaxamide, A1- A5 (cyclic pentapeptides) [21], Mol-
lamide B (cyclic hexapeptides) [22]; Rolloamide A, Stylissatin B (cyclic
heptapeptides) [23,24], Stylopeptide 2 (cyclodecapeptides) [25], Lax-
aphycin B5, B6 and Wewakazole B (cyclic dodecapeptides) [26,27] are
obtained from algae, ascidia, sponge and cyanobacteria. Cordyhepta-
peptide C-E (cyclic heptapeptides) is obtained from marine derived
fungus [28]. Cyclic depsipeptides have a more complicated structure,
with ester bonds substitutes for amide bonds secondary to the presence
of hydroxy acid in the peptide framework [9,17].
Numerous anti-cancer cyclic depsipepides have been recorded from
marine organisms. Desmethoxymajusculamide C [29], Cocosamides A-B
[30], Hantupeptin A-C [31], Malyngamide 3 [30], Pitiprolamide [32],
Pitipeptolide A-F [33], Isomalyngamide A and A-1 [34], Largazole [35];
Cryptophycin [36] Cryptophycin 1 [37], Coibamide A [38,39] are ob-
tained from cyanobacteria. Jaspamide / Jasplakinolide A-P [40], Geo-
diamolide A-E, H, I [4143], Pipecolidepsin A-B [44] from sponge;
Kahalalide F and Elisidepsin [45,46], Kulokekahilide-2 [47] from
mollusk; Dehydrodidemnin B (Aplidin / Plitidepsin) from ascidia Apli-
dium albicans [48] and Ohmyungsamycin A-B from streptomyces [49]
are cyclic depsipeptides with antibreast cancer properties.
Cyclic pentadepsipeptide Sansalvamide has been isolated from fun-
gus Fusarium sp [50]. Kailuin A-D (cyclic acyldepsipeptides) [51]. Thi-
ocoraline (cyclic thiodepsipeptide) [52] and Marthiapeptide A
(polythiazole cyclic peptide) [53] have been isolated from bacteria.
Macrocyclic peptide, Diazonamide A has been isolated from ascidia
Diazona angulata [54]. Cyclic oligopeptides are cyclic peptides consist-
ing of 220 amino acids formed by nonribosomal peptide synthesis [17].
Efrapeptin G from fungus Acremonium sp. is one such example [55].
Lipopeptides are linear, or cyclic lipid acylated peptides, typically
Fig. 1. Scheme of two of the involved pathways in apoptosis.
S. Ahmed et al.
Biomedicine & Pharmacotherapy 142 (2021) 112038
3
with the fatty acid side chain [9]. Several anti-cancer Lipopeptides have
been isolated from marine sources, including Curacin A-C, Somocysti-
namide A from cyanobacteria Lyngbya majuscula, and Schizothrix sp [56,
57]. Iturin A from bacteria Bacillus megaterium [58]. In triple negative
breast cancer, the Cyclo (L-Leucyl-L-Prolyl) a marine peptide has shown
propensity to target the EGFR and CD151 signaling pathway [59].
Marine Protein hydrolysates represent a family of nutraceuticals that
can prevent cancer [60]. Protein hydrolysates are characterized as oli-
gopeptides and free amino acid complex mixtures with antioxidant,
antiproliferative, antihypertensive, and antimicrobial effects [60,61].
Enzymatic hydrolysis is more likely to improve free radical scavenging
activity [62]. Protein hydrolysates obtained from sh [60], amphibians
[63] and turtles [64] have also been shown to possess anti-breast cancer
properties.
Marine anti-cancer peptides modulate/regulate a number of cellular
and molecular pathways, like DNA defense, cell-cycle control, apoptosis
initiation, angiogenesis suppression, migration, invasion, and metastasis
inhibition [4,5,8,65,66]. The current topic emphasizes the importance
of large marine peptides as an essential tool for discovering novel
anti-breast cancer treatments addressing their mechanistic effects.
3. Mechanistic insights
3.1. Apoptosis
A successful anti-cancer agent should have multidimensional
apoptotic targets [67,68]. Intrinsic and extrinsic pathways are linked to
caspase-3 (Casp-3) activation and induce DNA damage, nuclear and
cytoskeletal protein destruction, protein cross-linking, apoptosis body
development, and nally, phagocytic cell uptake (Fig. 1). The intrinsic
pathway is controlled by the Bcl-2 protein, releases Cyt c, and reacts
with APAF-1 to produce a platform for Casp-3, -7, and -9 activation.
Extrinsic pathways do not include mitochondria and are activated by
cell-surface death receptors [6971].
Cyt c discharge from mitochondria plays a pivotal role in apoptosis
induction, causing a sequence of biochemical reactions culminating in
activation of Casps and subsequent cell death [72,73]. C-phycocyanin
induces apoptosis in BT-474, HBL 100, MCF7, MDA MB-231, and
SKBR-3 cells in vitro by Cyt c release and the ensuing activation of Casps
-9 [74,75]. Mere15, a linear polypeptide formed by Meretrix meretrix,
mediates Cyt c discharge and Casp-3,-9 and PARP cleavage [76]. Iturin
A, a lipopeptide formed by Bacillus megaterium mediates the release of
Cyt c, Casp-3,-9 and PARP in MCF7, T47D and MDA-MB-231, -468 [77].
Casps act as core apoptosis executors and are activated upon pro-
teolytic cleavage [78]. For example, symplostatin 1 treatment increases
the activity of Casp-3 in MDA-MB-435 and NCI/ADR leading to cell
death with IC
50s
of 0.15 and 2.9 nM, respectively [20],. Jaspamide A-P
induces apoptosis by increasing Casp-3 in MCF7 [40]. Activation of
Casp-3,-7,-8 and-9 has been detected in response to dolastatin 10 and 15
treatment [54]. Curacin A-C [56], Mere15 [76] Somocystinamide A
[57], Dehydrodidemnin B [48] Coibamide A [39], Cryptophycin [79]
has shown the same behavior in MCF7, MCF-15, MDA-MB-231 and -435.
Galaxamide and Galaxamide analogs A1-A5 have shown cytotoxicity
against MCF7 cell-lines by activating Casp-3, -9 and PARP [21]. Simi-
larly, Keyhole limpet hemocyanin (KLH) marine peptides enhances
apoptotic activity in MCF7 cells by 250 ng mL
-1
[48,80].
The combination of Bcl-2 inhibition and Bax induction is a successful
means for initiating apoptosis [22]. Jaspamide A-P induces apoptosis in
MCF7 by stimulating Casp 3, increasing Bax, decreasing Bcl-2 and PARP
protein proteolysis [40]. Symplostatin shows an analogous effect in
MDA-MB-435 and NCI/ADR with IC
50 s
of 0.15 and 2.9 nM, respectively
[20,81]. Similarly, C-phycocyanin induced apoptosis in BT-474,HBL
100, MCF7, MDA-MB-231 and SKBR-3 cells [74,75]. Dolastatin 10 and
15 [8,82]; Mere15 leads to the same response in MCF-15, triggered by
p53 [76]. BAX upregulation and Bcl-2, Mcl-1, Bcl-xL, downregulation in
MDA-MB-231 and MCF7 cells has been noted upon Iturin A [83] and
Dehydrodidemnin B treatments [48,84]. Tuna hydrolysate protein
(Thunnus tonggol) has shown cytotoxicity in MCF7 cells with IC
50 s
of
1.39 mg mL
-1
causing apoptosis by increased Casp-3, -9, PARP, Bcl-2,
Bax, and p53 expressions [85].
The PI3K/AKT pathways plays an important role in controlling cell
cycle and survival. AKT inhibition reduces the level of phosphorylated
Bad, Bax, Bak, triggers Cyt c release, and activates Casp-9 and regulates
p53-dependent apoptosis [40]. PI3K/AKT inhibition and decreased
ErbB3 have been shown to lead to cell cycle arrest and Bax and Bak
activation [86]. Kahalalide F causes PI3K/AKT inhibition and ErbB3
depletion in MCF7, SKBR3 and BT474 cells [45,87,88]. Elisidepsin
(PM02734, Irvalec®), a Kahalalide F synthetic derivative, has shown
cytotoxic activity in MDA-MB-231, -361, -435 SKBR3 cells through
PI3K/AKT inhibition and ErbB3 depletion [46,87,89]. Dolastatins has
shown an analogous response [90]. FoxO3a is a tumor suppressor, and
has been shown to downregulate Akt transcription factor and induce
apoptosis [83]. Iturin A has been shown to downregulate AKT phos-
phorylation and upregulate FoxO3a in MCF7, MDA-MB-231,-468 and
T47D cells [58].
3.2. Antimitotic
Antimitotic drugs act through stabilization, destabilization of
microtubule dynamics [9193]. The microtubule system is essential for
mitosis and cell division, making it an effective anti-cancer drugstarget.
Microtubules and microtubule-associated proteins are the major con-
stituents of the mitotic spindle having a crucial role in cell division.
Alterations in the tubulin-microtubule equilibrium leads to degradation
of the mitotic spindle, interrupting the cell cycle during the
meta-anaphase transformation, ultimately resulting in cell death [94].
Belamide A [19], Diazonamide A [54,95] and MML from mollusk Mer-
etrix meretrix [96] have been shown to exhibit cytotoxicity in MCF7 cells
by increasing cell membrane permeability and tubulin depolymeriza-
tion. Hemiasterlin, Hemiasterlin A, and B interrupt the G2-M phase
through disrupting microtubule dynamics in MCF 7 cells with IC50
0.57 nM [18]. Hemiasterlin and Hemiasterlin C arrest the G2-M phase
secondary to microtubule depolymerization in MDA-MB-435 cells with
IC
50 s
of 0.0154 and 0.4002
μ
g mL
1
, respectively [97]. Similarly,
Microcionamide A, B [98] and Milnamide A - D [42,99] cause micro-
tubules depolymerization in MCF7, SKBR3, and MDA-MB-435 cells.
Scleritodermin A, a cyclic peptide, has been shown to be active in SKBR3
cells with IC
50 s
0.67 µM, and inhibited microtubule polymerization and
G2-M phase arrest [100]. Symplostatin 1 (dolastatin 10 analog) from
cyanobacteria Symploca sp. Has shown signicant anti-cancer effects in
murine mammary 16/C mouse xenograft model (early-stage mammary
adenocarcinoma). Increased Casp 3 increment, decreased Bcl-2 decre-
ment and microtubular depolymerization have been suggested as po-
tential anti-cancer mechanisms in mediating these effects [81].
Microtubule stabilizing agents accelerate microtubule polymeriza-
tion and damage the cytoskeleton and spindle cancer cells framework,
thus disrupting mitosis [91,94]. Cryptophycin has shown marked cyto-
toxicity in MCF7 (IC
50
of 0.016 nM), MCF-7/ADR (IC
50
value of
0.017 nM) and MDA-MB-435 (IC
50
of 50 pM) cells by stabilizing spindle
microtubules [36,37,101].
3.3. Antimetastatic
Microlaments have an essential role in cell migration. Actin poly-
merization inhibition results in microlament disruption, reducing cell
motility, and mitigates metastatic progression of neoplastic cells [102].
Desmethoxymajusculamide C is potent against MDA-MB-435 (IC
50
0.22 µM) cells, via actin microlament disruption [29]. Geodiamolides
A, B, D, E, H, and I have been shown to disrupt the actin laments in
T47D, MCF7, and MDA-MB-435 cells [42,43]. By positively affecting the
STAT3 pathway, MMP2 and MMP9 are subjected to upregulation and
facilitate cancer invasion. STAT3 inhibition leads to apoptosis and
S. Ahmed et al.
Biomedicine & Pharmacotherapy 142 (2021) 112038
4
suppresses metastasis in cancer cells [103]. C-phycocyanin inhibits
STAT3 in MCF7 and inhibits metastasis [75].
3.4. Antiangiogenic
Angiogenesis plays a central function in carcinogenesis (Fig. 2)
[104]. VEGF is abundantly expressed in cancer cells and is the key to
angiogenesis activation [105]. The angiogenesis mechanisms include
ERK1/2, CXCR4, HIF1
α
and Akt [106,107]. MMP2 and MMP9 also play
an essential role in tumor invasion and metastases [108]. Petrosaspongia
sponge Mycothiazole blocked HIF1
α
in T47D (IC
50
1 nM) along with
repression of the HIF1 reference gene VEGF expression [109].
C-phycocyanin [74], Isomalyngamide A and A-1 [34], and Coibamide A
[38,39] inhibit MCF7 and MDA-MB-231 cell migration by decreasing
VEGFR2 expression and MMP-9.
3.5. Cell cycle arrest
Disruption of the cell cycle is closely linked to apoptosis [110114].
Activation of the cyclin-dependent kinase (Cyclin D1 and Cyclin E) in-
hibitors, p21 and p53 suppresses tumor growth and protects against
DNA damage by halting the cell cycle and regulates apoptosis
[115117]. Coibamide A has shown G
1
phase arrest in MDA-MB-231
cells at GI
50
value of 2.8 nM [38,39]. Thiocoraline, a cyclic thiodep-
sipeptide from Micromonospora sp. displayed G1 phase arrest in SKBR3
cells with a GI
50
value of 2.2 nM [52]. C-phycocyanin [74,75] and
Dehydrodidemnin B (Aplidin / Plitidepsin) [118] induced G1-G2 phase
arrest by decreasing cyclin D1, cyclin E; and CDK2 and increasing p21 in
MCF7 and MDA-MB-231 cells, respectively. Hemiasterlin and Hemi-
asterlin A- C [18,97] and HTI-286 [119] induced G2-M phase arrest in
MCF7, MDA-MB-435 and MX-1W cells. Similarly, Mere15 induced p53
upregulation and cell cycle arrest in MCF-15 cells [76]. Peptides from
Porphyra haitanesis have shown cells cytotoxicity and G0/G1 phase ar-
rest in MCF7 with IC
50
of 200.97
μ
g mL
-1
[120].
3.6. Oxidative stress
Oxidative stress triggered by mitochondrial disorders is induced by
ROS (Reactive Oxygen Species) generation. A variety of factors such as a
hypoxia, antioxidants, ER stresses, and NADPH contribute or inhibit
ROS generation (Fig. 3) [121]. Aplidin induces apoptosis in
MDA-MB-231 cells by promoting increased GSSG/GSH ratio [122]. DNA
fragmentation, the most common DNA damage, is directly associated
with oxidative stress [123]. Cryptophycin 1 in MDA-MB-435 cells has
been shown to induce DNA fragmentation [37]. Iturin A has also been
shown to induce DNA fragmentation in MCF7, T47D, MDA-MB-231 and
-468 cells [58]. C-phycocyanin scavenges ROS and increases γ-H2AX in
MCF7 cells, an indicator of DNA damage [74,124]. Binucleated cells are
seen as a consequence of oxidative stress in response to symplostatin
[20].
Fig. 2. A schema of angiogenesis and its involved factors.
Fig. 3. Various factors that affect ROS generation.
S. Ahmed et al.
Biomedicine & Pharmacotherapy 142 (2021) 112038
5
3.7. Destruction of cancer cell membrane
Anti-cancer peptides (ACPs) cause cell membrane depolarization,
leading to tumor cellsfailure to sustain normal osmotic pressure and a
massive leakage of cytoplasmic material [125,126]. Anti-proliferative
peptides destroy cancer cells by necrotic mechanisms that trigger cell
membrane lysis [60,127,128]. Anti-cancer peptides cause membrane
destabilization, cell lyses, and the ensuing death of cancer cells [129,
130]. Anti-proliferative peptides with high ROS reduced activity can
prevent cancer incidence [60]. Protein hydrolysate from Thunnus tonggol
muscle by-product showed an anti-proliferative activity in MCF7 cells
with IC
50 s
of 8.1 and 8.8
μ
M, respectively [131]. Gadus morhua Atlantic
cod, Pleuronectes platessa plaice and Micromesistius poutassou sh hy-
drolysates induced anti-proliferative activity of MCF7/6 and
MDA-MB-231 cells at 1 g L
1
[132]. Loach (Misgurnus anguillicaudatus)
muscle hydrolysate obtained from the papain enzyme exhibits
anti-proliferative activity at 40 mg mL-1 MCF7 cancer cell line [133]. A
hydrolysate from Dosidicus gigas has shown toxicity in MCF7 cells, with
IC
50
value of 0.13 mg mL
-1
[134]. Antitumor peptides (RGVKGPR,
KLGPKGPR, and SSPGPPVH) from Cuora trifasciata turtle have been
shown to inhibit MCF7 cancer cells [64].
3.8. Unknown mechanism for anti-cancer activity of marine peptides
Ascidians mollamide B [22] and Spongian Callyptide A [122],
Criamide B [11,135], Pipecolidepsin A-B [44], Rolloamide A [23],
Stylissatin B [24] and Pembamide [136] induced potent cytotoxicity
with unidentied mechanism. Kulokekahilide-2 from mollusks [47] and
malyngamide 3 [30] cocosamides A and B [30], wewakazole B [27]
from cyanobacteria are other examples. Anti-cancer marine peptides
also include Kailuins AD from bacteria [51,137], Marthiapeptide A
from actinomycetes [53], Ohmyungsamycin A and Ohmyungsamycin B
from streptomyces [49], Cordyheptapeptide C -E [28], Sansalvamide
[50] and Efrapeptin G [138] from the marine sponge-derived fungus.
The peptide from saltwater clam Ruditapes philippinarum has shown
cytotoxicity in MDA-MB-231 cells with IC
50
of 1.58 mg mL
-1
[139,140].
4. Marine peptides in MDR cancers
The leading causes of chemotherapy failure are inherent or acquired
drug resistance due to overexpression of P-gp [141143]. MDA-MB-231,
-436,-468, BT20, BT-549, SKBR3, and Hs578T are useful for studying
molecular aberrations and mechanisms inuenced by these aberrations
in TNBC [144].
Hemiasterlin and Hemiasterlin C [97], HTI-286 [119,145] and Mil-
namide A - D [42,99] depolymerize MDA-MB-435 breast cancer mi-
crotubules showed their less interaction with MDR P-glycoprotein
(P-gp). Preclinical experiments have demonstrated that HTI-286 induces
the degradation of the tumor and reduces human MX-1W (MRP1 over-
expressed) breast carcinoma xenografts in mice paclitaxel and vincris-
tine were unsuccessful due to P-gp associated resistance [119].
Symplostatin 1 (dolastatin 10 analog) showed Bcl-2 inhibition and
microtubule depolymerization in MDR breast cancer MDA-MB-435 (IC
50
0.15 nM) and NCI/ADR (IC
50
2.9 nM) cells [20]. However, Symplostatin
1 showed marginal antitumor activity against MDR tumors mammary
adenocarcinoma 17/Adr and mammary adenocarcinoma 16/C/Adr
(adriamycin-resistant murine early stage solid tumors) [81]. Crypto-
phycin, a cytotoxic macrocyclic depsipeptide isolated from cyanobac-
teria Nostoc sp., is an antimicrotubule agent that tends to be a weaker
P-gp substrate than Vinca alkaloids.
Breast carcinoma cells are abundantly drug resistant due to increased
expression of P-gp and are signicantly less cryptophycin resistant than
colchicine, vinblastine and taxol. Cryptophycin has shown antimitotic
activity in breast cancer MCF7 and MCF7/ADR cells with IC
50 s
of 0.016
and 0.017 nM respectively through microtubule stabilization [36].
Cryptophycin (50 pM) also induces mitotic arrest in MDA-MB-435 by
developing distorted mitotic spindles without affecting the interphase
microtubules [37]. Geodiamolide D-E demonstrated antiproliferative
action toward MDA-MB-435, a P-gp upregulating MDR cell line, with
actin lament interruption [42]. Kulokekahilide-2 is also active towards
MDA-MB-435 cells, with an IC
50
of 14.6 nM [47]. Stylopeptide 2 had
antiproliferative effects on BT-549 and HS 578 T cells at a dosage of
10
5
M [25]. Ohmyungsamycin A and Ohmyungsamycin B has been
shown to exhibit antiproliferative effects against MDA-MB-231 with
IC50s of 0.688 and 12.7
μ
M, respectively [49]. Interestingly, there was
no cytotoxicity (IC50 >40
μ
M) in normal epithelial MRC-5 cells [49].
Largazole inhibits MDA-MB-231 with GI
50
value of 7.7 nM over normal
mammary NMuMG cells with IC
50
of 122 nM [35]. C-phycocyanin af-
fects cancer cell cycle propagation by G0/G1 step arrest in
MDA-MB-231, indicating weak P-gp transport substrate [146]. Pipeco-
lidepsin A, Pipecolidepsin B, Pembamide, and Callyptide A inhibit
MDA-MB-231 cellular growth with IC
50 s
of 0.7, 0.02, 3.35, and 29 µM,
respectively [44,122,136]. DZ-2384 has antitumor activity by inhibition
of mitotic spindle formation in metastatic TNBC (MDA-MB-231-LM2),
lacks neurotoxicity in rats, and signicantly increases survival [147].
Frog-derived peptide Hymenochirin-1B [63] and Alyteserin-2a [148]
have been shown to be cytotoxic to MDA-MB-231 cells. Micro-
cionamides A and B are cytotoxic to SKBR3 cells [98]. Kahalalide F
induced cytotoxicity in SKBR3 via PI3K-AKT inhibition [45,87]. Eli-
sidepsin has shown similar response in MDA-MB-231, -361, -435 and
SKBR3 [46,87].
5. Marine peptides in clinical trial status
Up to now, Hemiasterlin (E7974) [149] and Eribulin mesylate
(Halaven®) [87] have been approved by FDA for breast cancer. Pliti-
depsin (Aplidin®) [87], and Keyhole Limpet Hemocyanin (Immuco-
thel®) [149] have been approved by FDA and ATGA for different
cancers, but clinical trials for breast cancer are needed.
Dolastatin 10 have not advance to clinical trials due to its insigni-
cant therapeutic index and substantial toxic side effects. Thus, further
clinical trials are discontinued, and structural modications have been
established to enhance therapeutic efcacy, especially against TNBC
[13,150]. LU 103793 was evaluated in patients for efcacy and tolera-
bility in metastatic breast cancer but experienced neutropenia, asthenia,
stomatitis, myalgia, and hypertension resulting in cessation of further
assessment [151]. Soblidotin (TZT-1027) was engineered to retain
potent antitumor efcacy while reducing the parent drugs toxicity,
dolastatin 10. In human MX-1 mammary carcinoma xenografts models,
soblidotin showed very efcient outcomes [152]. Soblidotin has reached
phase I clinical trials and has demonstrated less neurotoxicity than other
tubulin inhibitors with a suggested dosage of 1.8 mg m
2
[12,153].
High toxicity, low solubility, and limited life span resulted in the
Didemnin B clinical trialswithdrawal in support of the second gener-
ation didemnin, plitidepsin [9,15]. Dehydrodidemnin B, commonly
known as (Aplidin or Plitidepsin), is more active than Didemnin pre-
clinical models against breast cancer cell lines and so far has not shown
evidence of life-threatening neuromuscular toxicity [15]. Plitidepsin is
in phase III clinical research for breast, melanoma, and non-small cell
lung cancers [154]. Elisidepsin (PM02734, Irvalec®), one of the most
potent analogs of Kahalalide F, was chosen for phase II clinical research
owing to its benecial therapeutic index and non-toxic prole [8].
Taltobulin or SPA-110 (HTI-286) is a synthetic hemiasterlin analog
with promising antimitotic action. This compound has progressed to
clinical trials for to its potential role in suppressing colchicine-like
tubulin polymerization and prevent cell division in MCF7 and MX-1W.
The terminations of HTI286 phase I clinical trials due to its toxicity
mandate the development of new synthetic formulations with lesser
adverse reactions [119,155].
S. Ahmed et al.
Biomedicine & Pharmacotherapy 142 (2021) 112038
6
Table 1
Anticancer effects of Marine peptides in the different reported studies.
Peptides Marine sources
(Species name)
Active derivative In vitro In vivo Anticancer
Mechanisms
References
Human
breast
cancer
cell lines
IC50s Experimental
model
Dose
Desmethoxymajusculamide
C
Cyanobacteria
(Lyngbya majuscula)
Cyclic depsipeptide MDA-
MB-435
0.22 µM Actin microlament
disruption
[29]
Isomalyngamide A and A-1 MDA-
MB-231
A, 0.06; A-1: 0.337
μ
M VEGFR2 and MMP-9
[34]
Cocosamides A-B MCF7 A: 30; B: 39
μ
M cell viability [30]
Hantupeptin A-C A: 4; B:0.5; C: 1.0 µM [31,163]
Malyngamide 3 29
μ
M [30]
Pitiprolamide 33
μ
M [32]
Pitipeptolide A-F A:13; B: 11; C: 73; D and
E: >100; F: 83
μ
M
[33]
Wewakazole B Cyclic
dodecapeptide
MCF7 0.58
μ
M [27]
Curacin A-C Lipopeptide A: 0.72; B: 0.82; C:
2.3
μ
M
Caspase 3;
Microtubule
depolymerization
[56]
Somocystinamide A Cyanobacteria
(Lyngbya majuscula
and Schizothrix sp.)
210 nM Caspase 8 [57]
Largazole Cyanobacteria
(Symploca sp.)
Cyclic depsipeptide MDA-
MB-231
7.7 nM cancer cell growth [35]
Belamide A Linear tetrapeptide MCF7 1.6
μ
M Microtubule
destabilization
[19]
Symplostatin Linear
Pentapeptide
MDA-
MB-435;
NCI/ADR
0.15 nM / 2.9 nM Caspase 3; Bcl2;
Bax; PARP ;
Microtubules
depolymerization
[20]
murine
mammary 16/C
mouse xenograft
model
1.25 mg/
kg i.v.
[81]
Cryptophycin Cyanobacteria
(Nostoc sp.)
Cyclic depsipeptide MCF7 0.016 nM Microtubule
stabilization
[36]
MCF-7/
ADR
0.017 nM
Cryptophycin 1 MDA-
MB-435
50 pM DNA fragmentation;
Microtubule
stabilization; Caspase
3
[37]
Laxaphycin B5, B6 Cyanobacteria
(Phormidium sp.)
Cyclic
dodecapeptide
MDA-
MB-231
GI
50
(µM) =B5: 2.2; B6:
0.81
cell viability [26]
MDA-
MB-435
GI
50
(µM) B5: 1.2; B6:
0.58
Coibamide A Cyanobacteria
(Leptolyngbya sp.)
Cyclic depsipeptide MDA-
MB-231
2.8 nM Caspase 3, -7 ;
VEGF; G
1
phase
arrest
[38,164]
C-phycocyanin Cyanobacteria
(Limnothrix sp. NS01
and Spirulina
platensis)
Peptide MCF7 15.43 µM DNA fragmentation;
Caspase-9 ; cyt c ;
Bcl2; Bax; PARP ;
Stat3
[74,75]
G1 - G2 phase arrest
(cyclin D1, cyclin E;
p21)
Antiangiogenic
(VEGFR2 and MMP-
9 )
AKT inhibition;
γ-H2AX ; Production
of ROS and singlet
oxygen radicals
HBL 100 8.31 µM
BT-474 8.45 µM
SKBR3 15.73 µM
MDA-
MB- 231
5.98 µM
Galaxamide, A1- A5 Algae (Galaxaura
lamentosa)
Cyclic
pentapeptide
MCF7 Galaxamide: 14.09; A1:
9.4; A2: 9.64; A3: 7.23;
A4: 6.56; A5:
4.18
μ
g mL
1
Caspase-9, -3 ; PARP
[21]
Callyptide A Sponge
(Callyspongia sp.)
Cyclic peptide MDA-
MB-231
GI
50
=29 µM cell viability [122]
Criamide B Sponge (Cymbastela
sp.)
Peptide MCF7 6.8
μ
g mL
1
[11,135]
Jaspamide A-P Sponge (Jaspis
splendans)
Cyclic depsipeptide A: 0.019; B: 3.41; C: 2; D:
0.05; E: 0.02; F: 30; G:
0.6; H: 30; J: 5; K: 0.48; L:
Caspase-3 ; Bcl2;
Bax; PARP
[40]
(continued on next page)
S. Ahmed et al.
Biomedicine & Pharmacotherapy 142 (2021) 112038
7
Table 1 (continued )
Peptides Marine sources
(Species name)
Active derivative In vitro In vivo Anticancer
Mechanisms
References
Human
breast
cancer
cell lines
IC50s Experimental
model
Dose
0.61; M: 0.1; N: 33; O:
0.38; P: 12
μ
M
Geodiamolide A-B, H, I Sponge (Auletta sp.
and Geodia
corticostylifera)
A: 17.83; B: 9.82; H:
89.96; I: 65.70 nM
Actin lament
disruption
[43]
T47D A: 18.82; B: 113.90; H:
38.36; I: 115.30 nM
[43]
Geodiamolide H MDA-
MB-231
4.33 ×10
-7
M cancer cell growth [41]
Geodiamolide H HS 578 T 2.45 ×10
-7
M
Geodiamolide D-E MDA-
MB-435
D: 0.08; E: 0.25
μ
g mL
1
Actin lament
disruption
[42]
Pipecolidepsin A-B Sponge
(Homophymia
lamellosa)
MDA-
MB-231
GI
50
=A: 0.7; B: 0.02
μ
M cell viability [44]
Rolloamide A Sponge (Eurypon
laughlini)
Cyclic
heptapeptides
MCF7 0.88 µM [23]
BT549 1.3 µM
MDA-
MB-231
2.2 µM
MDA-
MB-361
5.8 µM
MDA-
MB-435
0.40 µM
MDA-
MB-468
0.38 µM
Stylissatin B Sponge (Stylissa
massa)
MCF7 4.8
μ
M [24]
Stylopeptide 2 Sponge (Stylotella
sp.)
Cyclic decapeptide BT-549;
HS 578T
10
5
M Antiproliferative
effect
[25]
Scleritodermin A Sponge
(Scleritoderma
nodosum)
Cyclic peptide SKBR3 0.67 µM Microtubules
depolymerization;
G2/M phase arrest
[100]
Hemiasterlin and
Hemiasterlin A-B
Sponge
(Hemiasterella minor,
Auletta sp.,
Cymbastela sp., and
Siphonochalina sp.)
Linear tripeptide MCF7 Hemiasterlin: 0.5; A: 2; B:
7 nM
[18]
Hemiasterlin and
Hemiasterlin C
MDA-
MB-435
Hemiasterlin: 0.0154;
Hemiasterlin C:
0.4002
μ
g mL
1
[97]
HTI-286 MCF7 7.3 nM [119]
MX-1W 1.8 nM
MCF7 mouse
xenograft model
1 mg/kg
i.v. for 9
days
Antiproliferative
effect
[119]
MX-1W mouse
xenograft model
1.6 mg/
kg i.v.
Pembamide Sponge
(Cribrochalina sp.)
N-methylated
linear peptide
MDA-
MB-231
GI
50
=3.35
μ
M cell viability [136]
Microcionamide A-B Sponge (Clathria
(Thalysias) abietina)
MCF7 A: 125; B: 177 nM Microtubules
depolymerization
[98]
SKBR3 A: 98; B: 172 nM
Milnamide A and D Sponge (Auletta sp.
and Cymbastela sp.)
MDA-
MB-435
A: 6.02; D: 16.9 µM [99]
Milnamide B-C B: 1.48 ×10
4
; C:
0.32 ±0.02
μ
g mL
1
[42]
Mycothiazole Sponge
(Petrosaspongia
mycojiensis)
Mixed polyketide/
peptide-derived
compound
T47D 1 nM HIF1
α
inhibition;
VEGF
[109]
Dolastatin 10 Mollusk (Dolabella
auricularia)
Linear
Pentapeptide
MCF7 0.06 nM Caspase 3 ; Bcl2;
Bax; PARP ; p53;
Microtubule
depolymerization
[54]
Dolastatin 15 0.9 nM
Kahalalide F Mollusk (Elysia
rufescens)
Cyclic depsipeptide 0.28 µM PI3K-AKT inhibition;
ErbB3 depletion
[45]
SKBR3 0.23 µM
BT474 0.26 µM
Elisidepsin MCF7 8 µM [46]
MDA-
MB-231
4.7 µM
MDA-
MB-361
1.25 µM
MDA-
MB-435
4.4 µM
SKBR3 6 µM
ZR-751 0.4 µM
Kulokekahilide-2 Mollusk (Philinopsis
speciosa)
MDA-
MB-435
14.6 nM cell viability [47]
(continued on next page)
S. Ahmed et al.
Biomedicine & Pharmacotherapy 142 (2021) 112038
8
Table 1 (continued )
Peptides Marine sources
(Species name)
Active derivative In vitro In vivo Anticancer
Mechanisms
References
Human
breast
cancer
cell lines
IC50s Experimental
model
Dose
Mere15 Mollusk (Meretrix
meretrix)
Polypeptide MCF-15 57.43
μ
g mL
1
G2-M phase arrest;
Caspase 3, 9 ; cyt c ;
Bcl2; Bax; PARP ;
p53
[76]
Diazonamide A Ascidia (Diazona
angulata)
Macrocyclic
peptide
MCF7 1.9 nM Microtubule
depolymerization
[54]
DZ-2384 (synthetic
diazonamide)
BT-549 0.66 nmol [147]
GCRC
1735
7 nmol
GCRC
1915
17 nmol
MDA-
MB-231
7.7 nmol
MDA-
MB-436
2.8 nmol
MDA-MB-231-
LM2 mouse
xenograft model
4.5 mg/
m
2
i.v. for
28 days
Dehydrodidemnin B Ascidia (Aplidium
albicans)
Cyclic depsipeptide MCF7 50 nM G1 - G2 phase arrest
(cyclin D1; cyclin E;
p21), Caspase-9, 3
; Bcl2; Bax; PARP
[48]
MDA-
MB-231
5 nM [84]
Mollamide B Ascidia (Didemnum
molle)
Cyclic
hexapeptides
MCF7 100 µM cell viability [22]
Cordyheptapeptide C-E Fungus
(Acremonium
persicinum SCSIO
115)
Cyclic
heptapeptide
MCF7 C: 3; D: 82.7; E: 2.7
μ
M [28]
Efrapeptin G Fungus
(Acremonium sp.)
Oligopeptide 0.027
μ
M MCF7 mouse
xenograft model
0.15 mg/
kg i.p. for
28 days
cancer cell growth [55]
MDA-
MB-231
3.430
μ
M MDA-MB-231
mouse xenograft
model
0.3 mg/
kg i.p. for
28 days
T47D 0.057
μ
M
MDA-
MB-453
0.132
μ
M
Sansalvamide Fungus (Fusarium
sp.)
Cyclic
pentadepsipeptide
MAXF
401
0.02
μ
g mL
1
cell viability [50]
Iturin A Bacteria (Bacillus
megaterium)
Lipopeptide MCF7 12.16 ±0.24 µM DNA fragmentation;
AKT inhibition;
FoxO3a; Bcl2; Bax;
PARP ; Bcl-xL ; cyt c
[58]
T47D 26.29 ±0.78 µM
MDA-
MB-231
7.98 ±0.19 µM
MDA-
MB-468
13.30 ±0.97 µM
MDA-MB-231
mouse xenograft
model
10 mg/kg
i.v.
cell viability [58]
Kailuin A-D Bacteria (BH-107) Cyclic
acyldepsipeptide
MCF7 GI
50
(
μ
g mL
1
) =A: 3; B:
2; C: 4; D: 3
[51]
Kailuin D 39
μ
M [137]
Marthiapeptide A Bacteria
Marinactinospora
thermotolerans
SCSIO 00652)
Polythiazole cyclic
peptide
0.43
μ
M [53]
Proximicin A-C Bacteria
(Verrucosispora sp.)
Polyamide A: 24.6
μ
M; B: 12.1
μ
M;
C: 1.8
μ
M
[165]
Thiocoraline Bacteria
(Micromonospora sp.
L-13-ACM2092)
Cyclic
thiodepsipeptide
SKBR3 GI
50
=2.2 nM G
1
phase arrest [52]
Ohmyungsamycin A-B Bacteria
(Streptomyces strain
SNJ042)
Cyclic
depsipeptides
MDA-
MB-231
A: 0.68 and B: 12.7
μ
M cell viability [49]
S. Ahmed et al.
Biomedicine & Pharmacotherapy 142 (2021) 112038
9
6. Conclusions and future perspectives
The most prevalent and fatal illness in women is breast cancer. The
available treatments are effective for cancer treatment but have side
effects. There is still an urgent need for novel medications to be effective
for the cancer therapy. The discovery of novel clinical chemotherapeutic
peptides from diverse aquatic life can be incorporated into breast cancer
prevention and care [156]. The lack of ethnomedicinal background,
technical difculties in collecting marine animals, particularly deep-sea
organisms; isolation and purication problems are obstacles in
anti-cancer peptides research [157]. Thanks to modern technology, it is
increasingly possible to extract samples from the sea and different
peptides from aquatic materials [158]. Marine peptides have demon-
strated possible anti-cancer activities against various forms of cancer,
such as cell growth inhibition, antimitotic activity (anti-tubulin effects),
apoptosis induction, and migration, invasion or metastasis inhibition.
These marine peptides have proven to be a valuable and exciting
resource for developing anti-cancer drugs and a platform for discovering
new cellular targets for therapeutic action [159]. Therefore, it is highly
relevant to deepen the study of marine peptidesanti-cancer mecha-
nisms to develop new candidate compounds [160]. The tolerance of
cancer cells to chemotherapy is indeed one of the sources of modern
pharmacotherapys inefciency. Marine peptides act efciently as
MDR-threatening proteins. The more signicant part of the exploration
led to marine peptides; anti-cancer power is in vitro, making it difcult to
give the right determination on its helpfulness.
Of these compounds, only a few progressed to clinical studies, and a
relatively small number of peptides have successfully entered the
pharmaceutical pipeline and have been used clinically.
Numerous marine peptides are undergoing clinical trials, although
there is a widely unexplored area of marine protein hydrolysates [157].
Short half-life, low bioavailability, processing and manufacturing
problems, and protease susceptibility are signicant drawbacks of
therapeutic peptides [1214,83]. For low cell membrane permeability,
cell penetrating peptides are used. Metabolic instability and short
half-life in circulation may be overcome by using D-amino acid substi-
tution, peptide cyclization, encapsulation with nanoparticles, pegyla-
tion, and XTEN conjugation. D-amino acid substitution reduces
immunogenicity [126,161,162]. Protein hydrolysates are an alternative
source of anti-cancer, antioxidant, and antiproliferative bioactive com-
pounds. However, further investigation is required on the cell cycle
mode or apoptosis of cancer cell lines. in vivo and in silico studies are also
necessary to identify and characterize the mechanism of action and
safety of marine peptides and protein hydrolysates to achieve complete
anti-cancer drug efcacy [8]. In particular, further analysis of the
variability of marine peptides in structural modication and modes of
action would provide a rich resource for creating unique and potent new
pharmaceuticals (Table 1).
Ethics approval and consent to participate
Not applicable.
Funding
The project was supported by grant from The Oman Research
Council (TRC) through the funded project (BFP/RGP/HSS/19/198).
CRediT authorship contribution statement
Salman Ahmed and Ajmal Khan written the initial draft. Hamed
Mirzaei drawn the gures. Michael Aschner proof read the article for
English and grammatic corrections. Ahmed Al-Harrasi and Haroon Khan
designed and supervised the overall review.
Competing interests
The authors of this article have no nancial conict.
Consent for publication
Not applicable.
Availability of data and materials
All datasets on which the conclusions of the manuscript rely are
presented in the paper
References
[1] F. Bray, J. Ferlay, I. Soerjomataram, R.L. Siegel, L.A. Torre, A. Jemal, Global
cancer statistics 2018: GLOBOCAN estimates of incidence and mortality
worldwide for 36 cancers in 185 countries, CA Cancer J. Clin. 68 (2018)
394424.
[2] S. Senapati, A.K. Mahanta, S. Kumar, P. Maiti, Controlled drug delivery vehicles
for cancer treatment and their performance, Signal Transduct. Target. Ther. 3
(2018) 7, https://doi.org/10.1038/s41392-017-0004-3.
[3] D. Waghray, Q. Zhang, Inhibit or evade multidrug resistance p-glycoprotein in
cancer treatment, J. Med. Chem. 61 (2018) 51085121, https://doi.org/
10.1021/acs.jmedchem.7b01457.
[4] S. Khalifa, N. Elias, M.A. Farag, L. Chen, A. Saeed, M.F. Hegazy, M.S. Moustafa,
A. Abd El-Wahed, S.M. Al-Mousawi, S.G. Musharraf, F.R. Chang, A. Iwasaki,
K. Suenaga, M. Alajlani, U. G¨
oransson, H.R. El-Seedi, Marine natural products: a
source of novel anticancer drugs, Mar. Drugs 17 (2019) 491, https://doi.org/
10.3390/md17090491.
[5] P.C. Jimenez, D.V. Wilke, L.V. Costa-Lotufo, Marine drugs for cancer: surfacing
biotechnological innovations from the oceans, Clinics 73 (2018) 482, https://doi.
org/10.6061/clinics/2018/e482s.
[6] S. Ahmed, H. Khan, M. Aschner, H. Mirzae, E. Küpeli Akkol, R. Capasso,
Anticancer potential of furanocoumarins: mechanistic and therapeutic aspects,
Int. J. Mol. Sci. 21 (2020), https://doi.org/10.3390/ijms21165622.
[7] S. Ahmed, H. Khan, D. Fratantonio, M.M. Hasan, S. Shari, N. Fathi, H. Ullah,
L. Rastrelli, Apoptosis induced by luteolin in breast cancer: Mechanistic and
therapeutic perspectives, Phytomedicine 59 (2019), 152883, https://doi.org/
10.1016/j.phymed.2019.152883.
[8] V. Ruiz-Torres, J.A. Encinar, M. Herranz-L´
opez, A. P´
erez-S´
anchez, V. Galiano,
E. Barraj´
on-Catal´
an, V. Micol, An updated review on marine anticancer
compounds: the use of virtual screening for the discovery of small-molecule
cancer drugs, Molecules 22 (2017) 1037, https://doi.org/10.3390/
molecules22071037.
[9] M. Barreca, V. Span`
o, A. Montalbano, M. Cueto, A.R. Díaz Marrero, I. Deniz,
A. Erdo˘
gan, L. Luki´
c Bilela, C. Moulin, E. Tafn-de-Givenchy, F. Spriano,
G. Perale, M. Mehiri, A. Rotter, O. P Thomas, P. Barraja, S.P. Gaudˆ
encio,
F. Bertoni, Marine anticancer agents: an overview with a particular focus on their
chemical classes, Mar. Drugs 18 (2020) 619, https://doi.org/10.3390/
md18120619.
[10] Q.-T. Zhang, Z.D. Liu, Z. Wang, T. Wang, N. Wang, N. Wang, B. Zhang, Y.F. Zhao,
Recent advances in small peptides of marine origin in cancer therapy, Mar. Drugs
19 (2021) 115143.
[11] V. Gogineni, M.T. Hamann, Marine natural product peptides with therapeutic
potential: chemistry, biosynthesis, and pharmacology, Biochim. Biophys. Acta
(BBA) - Gen. Subj. 1862 (2018) 81196, https://doi.org/10.1016/j.
bbagen.2017.08.014.
[12] Hayashi, M.A., Ducancel, F. & Konno, K. (Hindawi, 2012).
[13] S. Marqus, E. Pirogova, T.J. Piva, Evaluation of the use of therapeutic peptides for
cancer treatment, J. Biomed. Sci. 24 (2017) 21.
[14] D.J. Craik, D.P. Fairlie, S. Liras, D. Price, The future of peptide-based drugs,
Chem. Biol. Drug Des. 81 (2013) 136147.
[15] H.K. Kang, M.-C. Choi, C.H. Seo, Y. Park, Therapeutic properties and biological
benets of marine-derived anticancer peptides, Int. J. Mol. Sci. 19 (2018) 919,
https://doi.org/10.3390/ijms19030919.
[16] K. Sridhar, B.S. Inbaraj, B.-H. Chen, Recent developments on production,
purication and biological activity of marine peptides, Food Res. Int. 147 (2021),
110468.
[17] Y. Lee, C. Phat, S.-C. Hong, Structural diversity of marine cyclic peptides and their
molecular mechanisms for anticancer, antibacterial, antifungal, and other clinical
applications, Peptides 95 (2017) 94105, https://doi.org/10.1016/j.
peptides.2017.06.002.
[18] H.J. Anderson, J.E. Coleman, R.J. Andersen, M. Roberge, Cytotoxic peptides
hemiasterlin, hemiasterlin A and hemiasterlin B induce mitotic arrest and
abnormal spindle formation, Cancer Chemother. Pharmacol. 39 (1996) 223226.
[19] T.L. Simmons, K.L. McPhail, E. Ortega-Barría, S.L. Mooberry, W.H. Gerwick,
Belamide A, a new antimitotic tetrapeptide from a Panamanian marine
cyanobacterium, Tetrahedron Lett. 47 (2006) 33873390, https://doi.org/
10.1016/j.tetlet.2006.03.082.
S. Ahmed et al.
Biomedicine & Pharmacotherapy 142 (2021) 112038
10
[20] S.L. Mooberry, R.M. Leal, T.L. Tinley, H. Luesch, R.E. Moore, T.H. Corbett, The
molecular pharmacology of symplostatin 1: a new antimitotic dolastatin 10
analog, Int. J. Cancer 104 (2003) 512521.
[21] X. Xiao, X. Liao, S. Qiu, Z. Liu, B. Du, S. Xu, Synthesis, cytotoxicity and apoptosis
induction in human tumor cells by galaxamide and its analogues [corrected],
Mar. Drugs 12 (2014) 45214538, https://doi.org/10.3390/md12084521.
[22] M.S. Donia, B. Wang, D.C. Dunbar, P.V. Desai, A. Patny, M. Avery, M.T. Hamann,
Mollamides B and C, cyclic hexapeptides from the Indonesian tunicate Didemnum
molle, J. Nat. Prod. 71 (2008) 941945.
[23] D.E. Williams, K. Yu, H.W. Behrisch, R. Van Soest, R.J. Andersen, Rolloamides A
and B, cytotoxic cyclic heptapeptides isolated from the caribbean marine sponge
Eurypon laughlini, J. Nat. Prod. 72 (2009) 12531257, https://doi.org/10.1021/
np900121m.
[24] J. Sun, W. Cheng, N.J. de Voogd, P. Proksch, W. Lin, Stylissatins BD,
cycloheptapeptides from the marine sponge Stylissa massa, Tetrahedron Lett. 57
(2016) 42884292, https://doi.org/10.1016/j.tetlet.2016.08.024.
[25] M.R. Brennan, C.E. Costello, S.D. Maleknia, G.R. Pettit, K.L. Erickson,
Stylopeptide 2, a proline-rich cyclodecapeptide from the sponge Stylotella sp,
J. Nat. Prod. 71 (2008) 453456, https://doi.org/10.1021/np0704856.
[26] P. Sullivan, A. Krunic, J.E. Burdette, J. Orjala, Laxaphycins B5 and B6 from the
cultured cyanobacterium UIC 10484, J. Antibiot. 73 (2020) 526533, https://doi.
org/10.1038/s41429-020-0301-x.
[27] J.A. Lopez, S.S. Al-Lihaibi, W.M. Alarif, A. Abdel-Lateff, Y. Nogata, K. Washio,
M. Morikawa, T. Okino, Wewakazole B, a cytotoxic cyanobactin from the
cyanobacterium moorea producens collected in the red sea, J. Nat. Prod. 79
(2016) 12131218, https://doi.org/10.1021/acs.jnatprod.6b00051.
[28] Z. Chen, Y. Song, Y. Chen, H. Huang, W. Zhang, J. Ju, Cyclic heptapeptides,
cordyheptapeptides CE, from the marine-derived fungus Acremonium
persicinum SCSIO 115 and their cytotoxic activities, J. Nat. Prod. 75 (2012)
12151219, https://doi.org/10.1021/np300152d.
[29] T.L. Simmons, L.M. Nogle, J. Media, F.A. Valeriote, S.L. Mooberry, W.H. Gerwick,
Desmethoxymajusculamide C, a cyanobacterial depsipeptide with potent
cytotoxicity in both cyclic and ring-opened forms, J. Nat. Prod. 72 (2009)
10111016.
[30] S.P. Gunasekera, C.S. Owle, R. Montaser, H. Luesch, V.J. Paul, Malyngamide 3
and cocosamides A and B from the marine cyanobacterium Lyngbya majuscula
from Cocos Lagoon, Guam, J. Nat. Prod. 74 (2011) 871876, https://doi.org/
10.1021/np1008015.
[31] A. Tripathi, J. Puddick, M.R. Prinsep, P.P.F. Lee, L.T. Tan, Hantupeptin A, a
cytotoxic cyclic depsipeptide from a singapore collection of Lyngbya majuscula,
J. Nat. Prod. 72 (2009) 2932, https://doi.org/10.1021/np800448t.
[32] R. Montaser, K.A. Abboud, V.J. Paul, H. Luesch, Pitiprolamide, a proline-rich
dolastatin 16 analogue from the marine cyanobacterium Lyngbya majuscula from
Guam, J. Nat. Prod. 74 (2011) 109112, https://doi.org/10.1021/np1006839.
[33] R. Montaser, V.J. Paul, H. Luesch, Pitipeptolides CF, antimycobacterial
cyclodepsipeptides from the marine cyanobacterium Lyngbya majuscula from
Guam, Phytochemistry 72 (2011) 20682074, https://doi.org/10.1016/j.
phytochem.2011.07.014.
[34] T.T. Chang, S.V. More, I.H. Lu, J.C. Hsu, T.J. Chen, Y.C. Jen, C.K. Lu, W.S. Li,
Isomalyngamide A, A-1 and their analogs suppress cancer cell migration in vitro,
Eur. J. Med. Chem. 46 (2011) 38103819, https://doi.org/10.1016/j.
ejmech.2011.05.049.
[35] K. Taori, V.J. Paul, H. Luesch, Structure and activity of largazole, a potent
antiproliferative agent from the oridian marine Cyanobacterium Symploca sp,
J. Am. Chem. Soc. 130 (2008) 18061807, https://doi.org/10.1021/ja7110064.
[36] C.D. Smith, X. Zhang, S.L. Mooberry, G.M. Patterson, R.E. Moore, Cryptophycin: a
new antimicrotubule agent active against drug-resistant cells, Cancer Res. 54
(1994) 37793784.
[37] S.L. Mooberry, L. Busquets, G. Tien, Induction of apoptosis by cryptophycin 1, a
new antimicrotubule agent, Int. J. Cancer 73 (1997) 440448.
[38] R.A. Medina, D.E. Goeger, P. Hills, S.L. Mooberry, N. Huang, L.I. Romero,
E. Ortega-Barría, W.H. Gerwick, K.L. McPhail, Coibamide A, a potent
antiproliferative cyclic depsipeptide from the Panamanian marine
cyanobacterium Leptolyngbya sp, J. Am. Chem. Soc. 130 (2008) 63246325,
https://doi.org/10.1021/ja801383f.
[39] J.D. Serrill, X. Wan, A.M. Hau, H.S. Jang, D.J. Coleman, A.K. Indra, A.W. Alani, K.
L. McPhail, J.E. Ishmael, Coibamide A, a natural lariat depsipeptide, inhibits
VEGFA/VEGFR2 expression and suppresses tumor growth in glioblastoma
xenografts, Investig. N. Drugs 34 (2016) 2440, https://doi.org/10.1007/s10637-
015-0303-x.
[40] F. Gala, M.V. DAuria, S. De Marino, V. Sepe, F. Zollo, C.D. Smith, S.N. Keller,
A. Zampella, Jaspamides MP: new tryptophan modied jaspamide derivatives
from the sponge Jaspis splendans, Tetrahedron 65 (2009) 5156, https://doi.org/
10.1016/j.tet.2008.10.076.
[41] W.F. Tinto, A.J. Lough, S. McLean, W.F. Reynolds, M. Yu, W.R. Chan,
Geodiamolides H and I, further cyclodepsipeptides from the marine sponge
Geodia sp, Tetrahedron 54 (1998) 44514458, https://doi.org/10.1016/S0040-
4020(98)00157-4.
[42] R.N. Sonnenschein, J.J. Farias, K. Tenney, S.L. Mooberry, E. Lobkovsky, J. Clardy,
P. Crews, A further study of the cytotoxic constituents of a milnamide-producing
sponge, Org. Lett. 6 (2004) 779782.
[43] M. Rangel, M.P. Prado, K. Konno, H. Naoki, J.C. Freitas, G.M. Machado-Santelli,
Cytoskeleton alterations induced by Geodia corticostylifera depsipeptides in
breast cancer cells, Peptides 27 (2006) 20472057, https://doi.org/10.1016/j.
peptides.2006.04.021.
[44] L. Coello, F. Reyes, M. a J. s Martín, C. Cuevas, R. Fernández, Isolation and
structures of pipecolidepsins A and B, cytotoxic cyclic depsipeptides from the
Madagascan sponge Homophymia lamellosa, J. Nat. Prod. 77 (2014) 298303.
[45] Y. Su´
arez, L. Gonz´
alez, A. Cuadrado, M. Berciano, M. Lafarga, A. Mu˜
noz,
Kahalalide F, a new marine-derived compound, induces oncosis in human
prostate and breast cancer cells, Mol. Cancer Ther. 2 (2003) 863872.
[46] M. Serova, A. de Gramont, I. Bieche, M.E. Riveiro, C.M. Galmarini, M. Aracil,
J. Jimeno, S. Faivre, E. Raymond, Predictive factors of sensitivity to elisidepsin, a
novel Kahalalide F-derived marine compound, Mar. Drugs 11 (2013) 944959,
https://doi.org/10.3390/md11030944.
[47] Y. Nakao, W.Y. Yoshida, Y. Takada, J. Kimura, L. Yang, S.L. Mooberry, P.
J. Scheuer, Kulokekahilide-2, a cytotoxic depsipeptide from a cephalaspidean
mollusk Philinopsis speciosa, J. Nat. Prod. 67 (2004) 13321340.
[48] D.R. Riggs, B. Jackson, L. Vona-Davis, D. McFadden, In vitro anticancer effects of
a novel immunostimulant: keyhole limpet hemocyanin, J. Surg. Res. 108 (2002)
279284.
[49] S. Um, T.J. Choi, H. Kim, B.Y. Kim, S.H. Kim, S.K. Lee, K.B. Oh, J. Shin, D.C. Oh,
Ohmyungsamycins a and b: cytotoxic and antimicrobial cyclic peptides produced
by Streptomyces sp. from a volcanic Island, J. Org. Chem. 78 (2013)
1232112329, https://doi.org/10.1021/jo401974g.
[50] G.N. Belofsky, P.R. Jensen, W. Fenical, Sansalvamide: a new cytotoxic cyclic
depsipeptide produced by a marine fungus of the genus Fusarium, Tetrahedron
Lett. 40 (1999) 29132916, https://doi.org/10.1016/S0040-4039(99)00393-7.
[51] G.G. Harrigan, B.L. Harrigan, B.S. Davidson, Kailuins AD, new cyclic
acyldepsipeptides from cultures of a marine-derived bacterium, Tetrahedron 53
(1997) 15771582, https://doi.org/10.1016/S0040-4020(96)01136-2.
[52] A. Negri, E. Marco, V. García-Hern´
andez, A. Domingo, A.L. Llamas-Saiz, S. Porto-
Sand´
a, R. Riguera, W. Laine, M.H. David-Cordonnier, C. Bailly, L.F. García-
Fern´
andez, J.J. Vaquero, F. Gago, Antitumor activity, X-ray crystal structure, and
DNA binding properties of thiocoraline A, a natural bisintercalating
thiodepsipeptide, J. Med. Chem. 50 (2007) 33223333, https://doi.org/10.1021/
jm070381s.
[53] X. Zhou, H. Huang, Y. Chen, J. Tan, Y. Song, J. Zou, X. Tian, Y. Hua, J. Ju,
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. 75 (2012) 22512255, https://doi.
org/10.1021/np300554f.
[54] Z. Cruz-Monserrate, H.C. Vervoort, R. Bai, D.J. Newman, S.B. Howell, G. Los, J.
T. Mullaney, M.D. Williams, G.R. Pettit, W. Fenical, E. Hamel, Diazonamide A and
a synthetic structural analog: disruptive effects on mitosis and cellular
microtubules and analysis of their interactions with tubulin, Mol. Pharmacol. 63
(2003) 12731280, https://doi.org/10.1124/mol.63.6.1273.
[55] C. Burz, I. Berindan-Neagoe, O. Balacescu, A. Irimie, Apoptosis in cancer: key
molecular signaling pathways and therapy targets, Acta Oncol. 48 (2009)
811821.
[56] P. Verdier-Pinard, J.Y. Lai, H.D. Yoo, J. Yu, B. Marquez, D.G. Nagle, M. Nambu, J.
D. White, J.R. Falck, W.H. Gerwick, B.W. Day, E. Hamel, Structure-activity
analysis of the interaction of curacin A, the potent colchicine site antimitotic
agent, with tubulin and effects of analogs on the growth of MCF-7 breast cancer
cells, Mol. Pharmacol. 53 (1998) 6276.
[57] W. Wrasidlo, A. Mielgo, V.A. Torres, S. Barbero, K. Stoletov, T.L. Suyama, R.
L. Klemke, W.H. Gerwick, D.A. Carson, D.G. Stupack, The marine lipopeptide
somocystinamide A triggers apoptosis via caspase 8, Proc. Natl. Acad. Sci. USA
105 (2008) 23132318.
[58] G. Dey, R. Bharti, G. Dhanarajan, S. Das, K.K. Dey, B.N. Kumar, R. Sen,
M. Mandal, Marine lipopeptide Iturin A inhibits Akt mediated GSK3β and FoxO3a
signaling and triggers apoptosis in breast cancer, Sci. Rep. 5 (2015) 114.
[59] K. Deepak, S. Kumari, G. Shailender, R.R. Malla, Marine natural compound cyclo
(L-leucyl-L-prolyl) peptide inhibits migration of triple negative breast cancer cells
by disrupting interaction of CD151 and EGFR signaling, Chem.-Biol. Interact. 315
(2020), 108872.
[60] M.I. Shaik, N.M. Sarbon, A review on purication and characterization of anti-
proliferative peptides derived from sh protein hydrolysate, Food Rev. Int.
(2020) 121.
[61] A. Neklyudov, A. Ivankin, A. Berdutina, Properties and uses of protein
hydrolysates, Appl. Biochem. Microbiol. 36 (2000) 452459.
[62] H. Korhonen, A. Pihlanto, Bioactive peptides: production and functionality, Int.
Dairy J. 16 (2006) 945960.
[63] S. Attoub, H. Arafat, M. Mechkarska, J.M. Conlon, Anti-tumor activities of the
host-defense peptide hymenochirin-1B, Regul. Pept. 187 (2013) 5156.
[64] S. He, X. Mao, T. Zhang, X. Guo, Y. Ge, C. Ma, X. Zhang, Separation and
nanoencapsulation of antitumor peptides from Chinese three-striped box turtle
(Cuora trifasciata), J. Microencapsul. 33 (2016) 344354.
[65] A.F. Wali, S. Majid, S. Rasool, S.B. Shehada, S.K. Abdulkareem, A. Firdous,
S. Beigh, S. Shakeel, S. Mushtaq, I. Akbar, H. Madhkali, M.U. Rehman, Natural
products against cancer: review on phytochemicals from marine sources in
preventing cancer, Saudi Pharm. J. 27 (2019) 767777, https://doi.org/10.1016/
j.jsps.2019.04.013.
[66] H. Malve, Exploring the ocean for new drug developments: marine pharmacology,
J. Pharm. Bioallied Sci. 8 (2016) 8391, https://doi.org/10.4103/0975-
7406.171700.
[67] M. Hassan, H. Watari, A. AbuAlmaaty, Y. Ohba, N. Sakuragi, Apoptosis and
molecular targeting therapy in cancer, Biomed. Res. Int. 2014 (2014), 150845,
https://doi.org/10.1155/2014/150845.
[68] S. Elmore, Apoptosis: a review of programmed cell death, Toxicol. Pathol. 35
(2007) 495516.
S. Ahmed et al.
Biomedicine & Pharmacotherapy 142 (2021) 112038
11
[69] G. Kroemer, Mitochondrial control of apoptosis: an introduction, Biochem.
Biophys. Res. Commun. 304 (2003) 433435.
[70] L. Oliver, F.M. Vallette, The role of caspases in cell death and differentiation,
Drug Resist. Updates 8 (2005) 163170.
[71] S. Cory, J.M. Adams, The Bcl2 family: regulators of the cellular life-or-death
switch, Nat. Rev. Cancer 2 (2002) 647656.
[72] X. Jiang, X. Wang, Cytochrome C-mediated apoptosis, Annu. Rev. Biochem. 73
(2004) 87106, https://doi.org/10.1146/annurev.biochem.73.011303.073706.
[73] S. Gupta, G.E. Kass, E. Szegezdi, B. Joseph, The mitochondrial death pathway: a
promising therapeutic target in diseases, J. Cell. Mol. Med. 13 (2009) 10041033.
[74] Z. Su, Z. Yang, Y. Xu, Y. Chen, Q. Yu, Apoptosis, autophagy, necroptosis, and
cancer metastasis, Mol. Cancer 14 (2015) 48.
[75] D. Heras-Sandoval, J.M. P´
erez-Rojas, J. Hern´
andez-Dami´
an, J. Pedraza-Chaverri,
The role of PI3K/AKT/mTOR pathway in the modulation of autophagy and the
clearance of protein aggregates in neurodegeneration, Cell. Signal. 26 (2014)
26942701.
[76] C. Wang, M. Liu, L. Cheng, J. Wei, N. Wu, L. Zheng, X. Lin, A novel polypeptide
from Meretrix meretrix Linnaeus inhibits the growth of human lung
adenocarcinoma, Exp. Biol. Med. 237 (2012) 442450.
[77] G. Dey, R. Bharti, G. Dhanarajan, S. Das, K.K. Dey, B.N. Kumar, R. Sen,
M. Mandal, Marine lipopeptide Iturin A inhibits Akt mediated GSK3β and FoxO3a
signaling and triggers apoptosis in breast cancer, Sci. Rep. 5 (2015) 10316,
https://doi.org/10.1038/srep10316.
[78] Y. Shi, Caspase activation, inhibition, and reactivation: a mechanistic view,
Protein Sci. 13 (2004) 19791987, https://doi.org/10.1110/ps.04789804.
[79] M. Abotaleb, P. Kubatka, M. Caprnda, E. Varghese, B. Zolakova, P. Zubor,
R. Opatrilova, P. Kruzliak, P. Stefanicka, D. Büsselberg, Chemotherapeutic agents
for the treatment of metastatic breast cancer: an update, Biomed. Pharmacother.
101 (2018) 458477, https://doi.org/10.1016/j.biopha.2018.02.108.
[80] D.R. Riggs, B.J. Jackson, L. Vona-Davis, A. Nigam, D.W. McFadden, In vitro
effects of keyhole limpet hemocyanin in breast and pancreatic cancer in regards
to cell growth, cytokine production, and apoptosis, Am. J. Surg. 189 (2005)
680684.
[81] H. Luesch, R.E. Moore, V.J. Paul, S.L. Mooberry, T.H. Corbett, Isolation of
dolastatin 10 from the marine cyanobacterium symploca species VP642 and total
stereochemistry and biological evaluation of its analogue symplostatin 1, J. Nat.
Prod. 64 (2001) 907910, https://doi.org/10.1021/np010049y.
[82] N.R. Wall, R.M. Mohammad, A.M. Al-Katib, Bax:Bcl-2 ratio modulation by
bryostatin 1 and novel antitubulin agents is important for susceptibility to drug
induced apoptosis in the human early pre-B acute lymphoblastic leukemia cell
line, Reh, Leuk. Res. 23 (1999) 881888, https://doi.org/10.1016/S0145-2126
(99)00108-3.
[83] W. Chiangjong, S. Chutipongtanate, S. Hongeng, Anticancer peptide:
physicochemical property, functional aspect and trend in clinical application
(Review), Int. J. Oncol. 57 (2020) 678696, https://doi.org/10.3892/
ijo.2020.5099.
[84] A. Cuadrado, L.F. Garcia-Fernandez, L. Gonzalez, Y. Suarez, A. Losada, V. Alcaide,
T. Martinez, J.M. Fernandez-Sousa, J.M. Sanchez-Puelles, A. Munoz, AplidinTM
induces apoptosis in human cancer cells via glutathione depletion and sustained
activation of the epidermal growth factor receptor, Src, JNK, and p38 MAPK*,
J. Biol. Chem. 278 (2003) 241250, https://doi.org/10.1074/jbc.M201010200.
[85] C.-C. Hung, Y.-H. Yang, P.-F. Kuo, K.-C. Hsu, Protein hydrolysates from tuna
cooking juice inhibit cell growth and induce apoptosis of human breast cancer
cell line MCF-7, J. Funct. Foods 11 (2014) 563570.
[86] H. Lee, H. Lee, H. Chin, K. Kim, D. Lee, ERBB3 knockdown induces cell cycle
arrest and activation of Bak and Bax-dependent apoptosis in colon cancer cells,
Oncotarget 5 (2014) 51385152, https://doi.org/10.18632/oncotarget.2094.
[87] E. Wang, M.A. Sorolla, P.D.G. Krishnan, A. Sorolla, From seabed to bedside: a
review on promising marine anticancer compounds, Biomolecules 10 (2020) 248,
https://doi.org/10.3390/biom10020248.
[88] M.L. Janmaat, J.A. Rodriguez, J. Jimeno, F.A.E. Kruyt, G. Giaccone, Kahalalide F
induces necrosis-like cell death that involves depletion of ErbB3 and inhibition of
akt signaling, Mol. Pharmacol. 68 (2005) 502510, https://doi.org/10.1124/
mol.105.011361.
[89] Y.-H. Ling, M. Aracil, Y. Zou, Z. Yuan, B. Lu, J. Jimeno, A.M. Cuervo, R. Perez-
Soler, PM02734 (elisidepsin) induces caspase-independent cell death associated
with features of autophagy, inhibition of the Akt/mTOR signaling pathway, and
activation of death-associated protein kinase, Clin. Cancer Res. 17 (2011)
53535366.
[90] H. Piplani, C. Rana, V. Vaish, K. Vaiphei, S. Sanyal, Dolastatin, along with
Celecoxib, stimulates apoptosis by a mechanism involving oxidative stress,
membrane potential change and PI3-K/AKT pathway down regulation, Biochim.
Biophys. Acta (BBA)-Gen. Subj. 1830 (2013) 51425156.
[91] R.A. Stanton, K.M. Gernert, J.H. Nettles, R. Aneja, Drugs that target dynamic
microtubules: a new molecular perspective, Med. Res. Rev. 31 (2011) 443481,
https://doi.org/10.1002/med.20242.
[92] J.A. Hadeld, S. Ducki, N. Hirst, A.T. McGown, Tubulin and microtubules as
targets for anticancer drugs, Prog. Cell Cycle Res. 5 (2003) 309326.
[93] E. Mukhtar, V.M. Adhami, H. Mukhtar, Targeting microtubules by natural agents
for cancer therapy, Mol. Cancer Ther. 13 (2014) 275284, https://doi.org/
10.1158/1535-7163.mct-13-0791.
[94] D. Fanale, G. Bronte, F. Passiglia, V. Cal`
o, M. Castiglia, F. Di Piazza, N. Barraco,
A. Cangemi, M.T. Catarella, L. Insalaco, A. Listì, R. Maragliano, D. Massihnia,
A. Perez, F. Toia, G. Cicero, V. Bazan, Stabilizing versus destabilizing the
microtubules: a double-edge sword for an effective cancer treatment option?
Anal. Cell. Pathol. 2015 (2015), 690916.
[95] M. Lachia, C.J. Moody, The synthetic challenge of diazonamide A, a macrocyclic
indole bis-oxazole marine natural product, Nat. Prod. Rep. 25 (2008) 227253.
[96] X. Ning, J. Zhao, Y. Zhang, S. Cao, M. Liu, P. Ling, X. Lin, A novel anti-tumor
protein extracted from Meretrix meretrix Linnaeus induces cell death by
increasing cell permeability and inhibiting tubulin polymerization, Int. J. Oncol.
35 (2009) 805812, https://doi.org/10.3892/ijo_00000393.
[97] W.R. Gamble, N.A. Durso, R.W. Fuller, C.K. Westergaard, T.R. Johnson, D.
L. Sackett, E. Hamel, J.H. Cardellina, M.R. Boyd, Cytotoxic and tubulin-
interactive hemiasterlins from Auletta sp. and Siphonochalina spp. sponges,
Bioorg. Med. Chem. 7 (1999) 16111615.
[98] R.A. Davis, G.C. Mangalindan, Z.P. Bojo, R.R. Antemano, N.O. Rodriguez, G.
P. Concepcion, S.C. Samson, D. de Guzman, L.J. Cruz, D. Tasdemir, M.K. Harper,
X. Feng, G.T. Carter, C.M. Ireland, Microcionamides A and B, bioactive peptides
from the Philippine sponge Clathria (Thalysias) abietina, J. Org. Chem. 69 (2004)
41704176.
[99] C. Chevallier, A.D. Richardson, M.C. Edler, E. Hamel, M.K. Harper, C.M. Ireland,
A new cytotoxic and tubulin-interactive milnamide derivative from a marine
sponge Cymbastela sp, Org. Lett. 5 (2003) 37373739.
[100] E.W. Schmidt, C. Raventos-Suarez, M. Bifano, A.T. Menendez, C.R. Fairchild, D.
J. Faulkner, Scleritodermin A, a cytotoxic cyclic peptide from the lithistid sponge
Scleritoderma nodosum, J. Nat. Prod. 67 (2004) 475478, https://doi.org/
10.1021/np034035z.
[101] D. Panda, V. Ananthnarayan, G. Larson, C. Shih, M.A. Jordan, L. Wilson,
Interaction of the antitumor compound cryptophycin-52 with tubulin,
Biochemistry 39 (2000) 1412114127.
[102] M. Trendowski, Exploiting the cytoskeletal laments of neoplastic cells to
potentiate a novel therapeutic approach, Biochim. Biophys. Acta (BBA) - Rev.
Cancer 1846 (2014) 599616, https://doi.org/10.1016/j.bbcan.2014.09.007.
[103] H. Lee, A.J. Jeong, S.-K. Ye, Highlighted STAT3 as a potential drug target for
cancer therapy, BMB Rep. 52 (2019) 415423, https://doi.org/10.5483/
BMBRep.2019.52.7.152.
[104] D.R. Bielenberg, B.R. Zetter, The contribution of angiogenesis to the process of
metastasis, Cancer J. 21 (2015) 267273.
[105] M. Shibuya, Vascular endothelial growth factor (VEGF) and its receptor (VEGFR)
signaling in angiogenesis: a crucial target for anti- and pro-angiogenic therapies,
Genes Cancer 2 (2011) 10971105, https://doi.org/10.1177/
1947601911423031.
[106] S. Nakamura, Y. Chikaraishi, K. Tsuruma, M. Shimazawa, H. Hara, Ruboxistaurin,
a PKCβ inhibitor, inhibits retinal neovascularization via suppression of
phosphorylation of ERK1/2 and Akt, Exp. Eye Res. 90 (2010) 137145.
[107] M. Ushio-Fukai, Redox signaling in angiogenesis: role of NADPH oxidase,
Cardiovasc. Res. 71 (2006) 226235.
[108] A. Winer, S. Adams, P. Mignatti, Matrix metalloproteinase inhibitors in cancer
therapy: turning past failures into future successes, Mol. Cancer Ther. 17 (2018)
11471155, https://doi.org/10.1158/1535-7163.mct-17-0646.
[109] J.B. Morgan, F. Mahdi, Y. Liu, V. Coothankandaswamy, M.B. Jekabsons, T.
A. Johnson, K.V. Sashidhara, P. Crews, D.G. Nagle, Y.D. Zhou, The marine sponge
metabolite mycothiazole: a novel prototype mitochondrial complex I inhibitor,
Bioorg. Med. Chem. 18 (2010) 59885994, https://doi.org/10.1016/j.
bmc.2010.06.072.
[110] M. Malumbres, M. Barbacid, Mammalian cyclin-dependent kinases, Trends
Biochem. Sci. 30 (2005) 630641.
[111] R. Suryadinata, M. Sadowski, B. Sarcevic, Control of cell cycle progression by
phosphorylation of cyclin-dependent kinase (CDK) substrates, Biosci. Rep. 30
(2010) 243255.
[112] H.C. Hwang, B.E. Clurman, Cyclin E in normal and neoplastic cell cycles,
Oncogene 24 (2005) 27762786.
[113] L.J. Hardwick, A. Philpott, Nervous decision-making: to divide or differentiate,
Trends Genet. 30 (2014) 254261.
[114] R. Visconti, R. Della Monica, D. Grieco, Cell cycle checkpoint in cancer: a
therapeutically targetable double-edged sword, J. Exp. Clin. Cancer Res. 35
(2016) 153, https://doi.org/10.1186/s13046-016-0433-9.
[115] O. Cazzalini, A.I. Scovassi, M. Savio, L.A. Stivala, E. Prosperi, Multiple roles of the
cell cycle inhibitor p21CDKN1A in the DNA damage response, Mutat. Res. /Rev.
Mutat. Res. 704 (2010) 1220.
[116] B. Shamloo, S. Usluer, p21 in cancer research, Cancers 11 (2019) 1178, https://
doi.org/10.3390/cancers11081178.
[117] K. Hientz, A. Mohr, D. Bhakta-Guha, T. Efferth, The role of p53 in cancer drug
resistance and targeted chemotherapy, Oncotarget 8 (2017) 89218946.
[118] M. Pelay-Gimeno, Y. García-Ramos, M. Jesús Martin, J. Spengler, J.M. Molina-
Guijarro, S. Munt, A.M. Francesch, C. Cuevas, J. Tulla-Puche, F. Albericio, The
rst total synthesis of the cyclodepsipeptide pipecolidepsin A, Nat. Commun. 4
(2013) 110.
[119] F. Loganzo, C.M. Discafani, T. Annable, C. Beyer, S. Musto, M. Hari, X. Tan,
C. Hardy, R. Hernandez, M. Baxter, T. Singanallore, G. Khazova, M.
S. Poruchynsky, T. Fojo, J.A. Nieman, S. Ayral-Kaloustian, A. Zask, R.J. Andersen,
L.M. Greenberger, HTI-286, a synthetic analogue of the tripeptide hemiasterlin, is
a potent antimicrotubule agent that circumvents P-glycoprotein-mediated
resistance in vitro and in vivo, Cancer Res. 63 (2003) 18381845.
[120] X. Fan, L. Bai, X. Mao, X. Zhang, Novel peptides with anti-proliferation activity
from the Porphyra haitanesis hydrolysate, Process Biochem. 60 (2017) 98107,
https://doi.org/10.1016/j.procbio.2017.05.018.
[121] E.B. Kurutas, The importance of antioxidants which play the role in cellular
response against oxidative/nitrosative stress: current state, Nutr. J. 15 (2016) 71,
https://doi.org/10.1186/s12937-016-0186-5.
S. Ahmed et al.
Biomedicine & Pharmacotherapy 142 (2021) 112038
12
[122] L.A. Shaala, D.T. Youssef, S.R. Ibrahim, G.A. Mohamed, Callyptide A, a new
cytotoxic peptide from the Red Sea marine sponge Callyspongia species, Nat.
Prod. Res. 30 (2016) 27832790.
[123] C. Guo, L. Sun, X. Chen, D. Zhang, Oxidative stress, mitochondrial damage and
neurodegenerative diseases, Neural Regen. Res. 8 (2013) 20032014, https://doi.
org/10.3969/j.issn.1673-5374.2013.21.009.
[124] M. Podhorecka, A. Skladanowski, P. Bozko, H2AX phosphorylation: its role in
DNA damage response and cancer therapy, J. Nucleic Acids 2010 (2010), 920161,
https://doi.org/10.4061/2010/920161.
[125] D. Gaspar, A.S. Veiga, M.A.R.B. Castanho, From antimicrobial to anticancer
peptides. A review, Front. Microbiol. 4 (2013) 294, https://doi.org/10.3389/
fmicb.2013.00294.
[126] M. Xie, D. Liu, Y. Yang, Anti-cancer peptides: classication, mechanism of action,
reconstruction and modication, Open Biol. 10 (2020), 200004.
[127] Y.-b Huang, X.-f Wang, H.-y Wang, Y. Liu, Y. Chen, Studies on mechanism of
action of anticancer peptides by modulation of hydrophobicity within a dened
structural framework, Mol. Cancer Ther. 10 (2011) 416426.
[128] E. Teerasak, P. Thongararm, S. Roytrakul, L. Meesuk, P. Chumnanpuen,
Prediction of anticancer peptides against MCF-7 breast cancer cells from the
peptidomes of Achatina fulica mucus fractions, Comput. Struct. Biotechnol. J. 14
(2016) 4957.
[129] D. Raucher, J.S. Ryu, Cell-penetrating peptides: strategies for anticancer
treatment, Trends Mol. Med. 21 (2015) 560570.
[130] F. Harris, S.R. Dennison, J. Singh, D.A. Phoenix, On the selectivity and efcacy of
defense peptides with respect to cancer cells, Med. Res. Rev. 33 (2013) 190234.
[131] K.-C. Hsu, E.C. Li-Chan, C.-L. Jao, Antiproliferative activity of peptides prepared
from enzymatic hydrolysates of tuna dark muscle on human breast cancer cell line
MCF-7, Food Chem. 126 (2011) 617622.
[132] L. Picot, S. Bordenave, S. Didelot, I. Fruitier-Arnaudin, F. Sannier, G. Thorkelsson,
J.P. Berg´
e, F. Gu´
erard, A. Chabeaud, J.M. Piot, Antiproliferative activity of sh
protein hydrolysates on human breast cancer cell lines, Process Biochem. 41
(2006) 12171222, https://doi.org/10.1016/j.procbio.2005.11.024.
[133] L. You, M. Zhao, R.H. Liu, J.M. Regenstein, Antioxidant and antiproliferative
activities of loach (Misgurnus anguillicaudatus) peptides prepared by papain
digestion, J. Agric. Food Chem. 59 (2011) 79487953, https://doi.org/10.1021/
jf2016368.
[134] A. Alem´
an, E. P´
erez-Santín, S. Bordenave-Juchereau, I. Arnaudin, M.C. G´
omez-
Guill´
en, P. Montero, Squid gelatin hydrolysates with antihypertensive, anticancer
and antioxidant activity, Food Res. Int. 44 (2011) 10441051.
[135] J.E. Coleman, E. Dilip de Silva, F. Kong, R.J. Andersen, T.M. Allen, Cytotoxic
peptides from the marine sponge Cymbastela sp, Tetrahedron 51 (1995)
1065310662, https://doi.org/10.1016/0040-4020(95)00646-P.
[136] C. Urda, M. P´
erez, J. Rodríguez, C. Jim´
enez, C. Cuevas, R. Fern´
andez,
Pembamide, a N-methylated linear peptide from a sponge Cribrochalina sp,
Tetrahedron Lett. 57 (2016) 32393242, https://doi.org/10.1016/j.
tetlet.2016.05.054.
[137] C.M. Theodore, N. Lorig-Roach, P.C. Still, T.A. Johnson, M. Draˇ
skovi´
c, J.
A. Schwochert, C.N. Naphen, M.S. Crews, S.A. Barker, F.A. Valeriote, R.S. Lokey,
P. Crews, Biosynthetic products from a nearshore-derived gram-negative
bacterium enable reassessment of the kailuin depsipeptides, J. Nat. Prod. 78
(2015) 441452, https://doi.org/10.1021/np500840n.
[138] C.M. Boot, K. Tenney, F.A. Valeriote, P. Crews, Highly N-methylated linear
peptides produced by an atypical sponge-derived Acremonium sp, J. Nat. Prod. 69
(2006) 8392.
[139] E.-K. Kim, Y.S. Kim, J.W. Hwang, J.S. Lee, S.H. Moon, B.T. Jeon, P.J. Park,
Purication and characterization of a novel anticancer peptide derived from
Ruditapes philippinarum, Process Biochem. 48 (2013) 10861090, https://doi.
org/10.1016/j.procbio.2013.05.004.
[140] Z.-G. Sun, Z.-N. Li, L.-H. Zhao, J.-M. Zhang, Research progress of anti-breast
cancer peptides, Int. J. Innov. Res. Med. Sci. 4 (2019) 504506, https://doi.org/
10.23958/ijirms/vol04-i08/734.
[141] J. Kasaian, F. Mosaffa, J. Behravan, M. Masullo, S. Piacente, M. Ghandadi,
M. Iranshahi, Reversal of P-glycoprotein-mediated multidrug resistance in MCF-
7/Adr cancer cells by sesquiterpene coumarins, Fitoterapia 103 (2015) 149154,
https://doi.org/10.1016/j.tote.2015.03.025.
[142] S. Dewanjee, T.K. Dua, N. Bhattacharjee, A. Das, M. Gangopadhyay, R. Khanra,
S. Joardar, M. Riaz, V. Feo, M. Zia-Ul-Haq, Natural products as alternative choices
for p-glycoprotein (P-gp) inhibition, Molecules 22 (2017), https://doi.org/
10.3390/molecules22060871.
[143] J. Jiang, X. Wang, K. Cheng, W. Zhao, Y. Hua, C. Xu, Z. Yang, Psoralen reverses
the P-glycoprotein-mediated multidrug resistance in human breast cancer MCF-7/
ADR cells, Mol. Med. Rep. 13 (2016) 47454750, https://doi.org/10.3892/
mmr.2016.5098.
[144] L. Billen, A. Shamas-Din, D. Andrews, Bid: a Bax-like BH3 protein, Oncogene 27
(2008) S93S104.
[145] A. Yamashita, E.B. Norton, J.A. Kaplan, C. Niu, F. Loganzo, R. Hernandez, C.
F. Beyer, T. Annable, S. Musto, C. Discafani, A. Zask, S. Ayral-Kaloustian,
Synthesis and activity of novel analogs of hemiasterlin as inhibitors of tubulin
polymerization: modication of the A segment, Bioorg. Med. Chem. Lett. 14
(2004) 53175322.
[146] L. Jiang, Y. Wang, G. Liu, H. Liu, F. Zhu, H. Ji, B. Li, C-Phycocyanin exerts anti-
cancer effects via the MAPK signaling pathway in MDA-MB-231 cells, Cancer Cell
Int. 18 (2018) 12, https://doi.org/10.1186/s12935-018-0511-5.
[147] M. Wieczorek, J. Tcherkezian, C. Bernier, A.E. Prota, S. Chaaban, Y. Rolland,
C. Godbout, M.A. Hancock, J.C. Arezzo, O. Ocal, C. Rocha, N. Olieric, A. Hall,
H. Ding, A. Bramoull´
e, M.G. Annis, G. Zogopoulos, P.G. Harran, T.M. Wilkie, R.
A. Brekken, P.M. Siegel, M.O. Steinmetz, G.C. Shore, G.J. Brouhard, A. Roulston,
The synthetic diazonamide DZ-2384 has distinct effects on microtubule curvature
and dynamics without neurotoxicity, Sci. Transl. Med. 8 (2016) 365.
[148] J.M. Conlon, M. Mechkarska, M. Prajeep, K. Arafat, M. Zaric, M.L. Lukic,
S. Attoub, Transformation of the naturally occurring frog skin peptide, alyteserin-
2a into a potent, non-toxic anti-cancer agent, Amino Acids 44 (2013) 715723.
[149] M.A. Ghareeb, M.A. Tammam, A. El-Demerdash, A.G. Atanasov, Insights about
clinically approved and Preclinically investigated marine natural products, Curr.
Res. Biotechnol. 2 (2020) 88102, https://doi.org/10.1016/j.crbiot.2020.09.001.
[150] E.A. Perez, D.W. Hillman, P.A. Fishkin, J.E. Krook, W.W. Tan, P.A. Kuriakose, S.
R. Alberts, S.R. Dakhil, Phase II trial of dolastatin-10 in patients with advanced
breast cancer, Investig. N. Drugs 23 (2005) 257261, https://doi.org/10.1007/
s10637-005-6735-y.
[151] P. Kerbrat, V. Dieras, N. Pavlidis, A. Ravaud, J. Wanders, P. Fumoleau, O. EORTC
Early Clinical Studies Group/New Drug Development, Phase II study of LU
103793 (dolastatin analogue) in patients with metastatic breast cancer, Eur. J.
Cancer 39 (2003) 317320, https://doi.org/10.1016/S0959-8049(02)00531-2.
[152] J. WATANABE, M. MINAMI, M. KOBAYASHI, Antitumor activity of TZT-1027
(Soblidotin), Anticancer Res. 26 (2006) 19731981.
[153] A.M. Mayer, K.B. Glaser, C. Cuevas, R.S. Jacobs, W. Kem, R.D. Little, J.
M. McIntosh, D.J. Newman, B.C. Potts, D.E. Shuster, The odyssey of marine
pharmaceuticals: a current pipeline perspective, Trends Pharmacol. Sci. 31 (2010)
255265.
[154] M.J. Mu˜
noz-Alonso, L. Gonz´
alez-Santiago, N. Zarich, T. Martínez, E. Alvarez, J.
M. Rojas, A. Mu˜
noz, Plitidepsin has a dual effect inhibiting cell cycle and inducing
apoptosis via Rac1/c-Jun NH2-terminal kinase activation in human melanoma
cells, J. Pharmacol. Exp. Ther. 324 (2008) 10931101.
[155] A. Zask, J. Kaplan, S. Musto, F. Loganzo, Hybrids of the hemiasterlin analogue
taltobulin and the dolastatins are potent antimicrotubule agents, J. Am. Chem.
Soc. 127 (2005) 1766717671.
[156] K. Senthilkumar, V. Jayachandran, K. Se-Kwon, Marine derived bioactive
compounds for breast and prostate cancer treatment: a review, Curr. Bioact.
Compd. 10 (2014) 6274, https://doi.org/10.2174/
1573407210666140327212945.
[157] G.-M. Suarez-Jimenez, A. Burgos-Hernandez, J.-M. Ezquerra-Brauer, Bioactive
peptides and depsipeptides with anticancer potential: sources from marine
animals, Mar. Drugs 10 (2012) 963986, https://doi.org/10.3390/md10050963.
[158] L. Zheng, X. Lin, N. Wu, M. Liu, Y. Zheng, J. Sheng, X. Ji, M. Sun, Targeting
cellular apoptotic pathway with peptides from marine organisms, Biochim.
Biophys. Acta (BBA)-Rev. Cancer 1836 (2013) 4248.
[159] R. Beesoo, V. Neergheen-Bhujun, R. Bhagooli, T. Bahorun, Apoptosis inducing
lead compounds isolated from marine organisms of potential relevance in cancer
treatment, Mutat. Res./Fundam. Mol. Mech. Mutagen. 768 (2014) 8497, https://
doi.org/10.1016/j.mrfmmm.2014.03.005.
[160] L. Zheng, Y. Xu, X. Lin, Z. Yuan, M. Liu, S. Cao, F. Zhang, R.J. Linhardt, Recent
progress of marine polypeptides as anticancer agents. Recent patents on anti-
cancer drug discovery 13, 445454, 2018.
[161] A. Sorolla, E. Wang, E. Golden, C. Duffy, S.T. Henriques, A.D. Redfern,
P. Blancafort, Precision medicine by designer interference peptides: applications
in oncology and molecular therapeutics, Oncogene 39 (2020) 11671184,
https://doi.org/10.1038/s41388-019-1056-3.
[162] A. Sorolla, M.A. Sorolla, E. Wang, V. Ce˜
na, Peptides, proteins and
nanotechnology: a promising synergy for breast cancer targeting and treatment,
Expert Opin. Drug Deliv. 17 (2020) 15971613, https://doi.org/10.1080/
17425247.2020.1814733.
[163] M.T. Hamann, P.J. Scheuer, Kahalalide F: a bioactive depsipeptide from the
sacoglossan mollusk Elysia rufescens and the green alga Bryopsis sp, J. Am. Chem.
Soc. 115 (1993) 58255826.
[164] J.D. Serrill, X. Wan, A.M. Hau, H.S. Jang, D.J. Coleman, A.K. Indra, A.W. Alani, K.
L. McPhail, J.E. Ishmael, Coibamide A, a natural lariat depsipeptide, inhibits
VEGFA/VEGFR2 expression and suppresses tumor growth in glioblastoma
xenografts, Investig. N. Drugs 34 (2016) 2440.
[165] K. Schneider, S. Keller, F.E. Wolter, L. R¨
oglin, W. Beil, O. Seitz, G. Nicholson,
C. Bruntner, J. Riedlinger, H.P. Fiedler, R.D. Süssmuth, Proximicins A, B, and
Cantitumor furan analogues of netropsin from the marine actinomycete
Verrucosispora induce upregulation of p53 and the cyclin kinase inhibitor p21,
Angew. Chem. Int. Ed. 47 (2008) 32583261.
S. Ahmed et al.
... 7 Bioactive peptides have been reported to show various biological activities, such as anti-inflammatory, anti-hypertensive, anti-thrombotic, immunomodulatory, anticancer and anti-microbial. 8,9 In addition, those bioactive peptides have greater application prospects as functional food ingredients and medical disease treatment due to the low toxicity and mild side effects. 10 Current research has demonstrated that many sources of bioactive peptides have antibreast cancer activity. ...
... 10,11 Furthermore, bioactive peptides suppressed the breast cancer cells proliferation by causing apoptosis, autophagy suppression, and cell cycle arrest. 9 Hung et al. found that the peptides obtained from tuna cooking juice induced the MCF-7 cells apoptosis by enhancing the expression levels of B-cell lymphoma (Bcl-2) and caspase 9, and decreasing the expression levels of Bax and cleaved-caspase 3. 12 The bioactive peptides that were extracted from the skin mucus of Anabas testudineus also increased the level of p53 expression and caused cell cycle arrest at the G0/G1 phase in breast cancer cells. 13 Two dogfish peptides that have been isolated have shown anti-breast cancer activity by altering the cytoskeleton and causing early autophagy suppression. ...
Article
Full-text available
BACKGROUND Porcupine quills, a by‐product of porcupine pork, are rich in keratin, which is an excellent source of bioactive peptides. The objective of this study was to investigate the underlying mechanism of anti‐proliferation effect of porcupine quills keratin peptides (PQKPs) on MCF‐7 cells. RESULTS Results showed that PQKPs induced MCF‐7 cells apoptosis by significantly decreasing the secretion level of anti‐apoptosis protein Bcl‐2 and increasing the secretion levels of pro‐apoptosis proteins Bax, cytochrome c, caspase 9, caspase 3 and PARP. PQKPs also arrested the cell cycle at G0/G1 phase via remarkably reducing the protein levels of CDK4 and enhancing the protein levels of p53 and p21. High‐performance liquid chromatography–tandem mass spectrometry (HPLC–MS/MS) analysis identified nine peptides with molecular weights less than 1000 Da in PQKPs. Molecular docking results showed that TPGPPT and KGPAC identified from PQKPs could bind with p53 mutant and Bcl‐2 protein by conventional hydrogen bonds, carbon hydrogen bonds and van der Waals force. Furthermore, the anti‐proliferation impact of synthesized peptides (TPGPPT and KGPAC) was shown in MCF‐7 cells. CONCLUSION These findings indicated that PQKPs suppressed the proliferation of MCF‐7 breast cancer cells by triggering apoptosis and G0/G1 cell cycle arrest. Moreover, the outcome of this study will bring fresh insights into the production and application of animal byproducts. © 2023 Society of Chemical Industry.
... Over half of the drugs the FDA authorized in the 1980s and 1990s originated from aquatic life. Sea-derived medications are a substantial and essential source of anticancer therapies (Ahmed et al. 2021). Multiple pharmaceuticals originating from marine sources have received clearance as anticancer agents, starting with the first permission for cytarabine in 1969. ...
Article
Full-text available
Breast cancer is the most common cancer among women worldwide, and marine creatures are the most abundant reservoir of anticancer medicines. Tachyplesin peptides have shown antibacterial capabilities, but their potential to inhibit cancer growth and trigger cancer cell death has not been investigated. A synthetic tachyplesin nucleotide sequence was generated and inserted into the pcDNA3.1( +) Mammalian Expression Vector. PCR analysis and enzyme digesting procedures were used to evaluate the vectors' accuracy. The transfection efficiency of MCF-7 and MCF10-A cells was 57% and 65%, respectively. The proliferation of MCF-7 cancer cells was markedly suppressed. Administration of plasmid DNA (pDNA) combined with tachyplesin to mice with tumors did not cause any discernible morbidity or mortality throughout treatment. The final body weight curves revealed a significant reduction in weight among mice treated with pDNA/tachyplesin and tachyplesin at a dose of 100 µg/ml (18.4 ± 0.24 gr, P < 0.05; 11.4 ± 0.24 gr P < 0.01) compared to the control group treated with PBS (22 ± 0.31 gr). Animals treated with pDNA/tachyplesin and tachyplesin exhibited a higher percentage of CD4 + Foxp3 + Tregs, CD8 + Foxp3 + Tregs, and CD4 + and CD8 + T cell populations expressing CTLA-4 in their lymph nodes and spleen compared to the PBS group. The groups that received pDNA/tachyplesin exhibited a substantial upregulation in the expression levels of caspase-3, caspase-8, BAX, PI3K, STAT3, and JAK genes. The results offer new possibilities for treating cancer by targeting malignancies using pDNA/tachyplesin and activating the mTOR and NFκB signaling pathways.
... Several marine-derived compounds have been reported to exhibit anticancer properties in PCa models [41,42]. While there is a potentially wide range of mechanisms through which their anticancer effects may be mediated, in pre-clinical studies, these peptides have been shown to sensitize cancer cells to anticancer agents [43,44]. While some challenges remain before marine-based peptides may have widespread clinical applications, some anticancer peptides have reached the clinical testing stage [45]. ...
Article
Full-text available
Recent research highlights the key role of iron dyshomeostasis in the pathogenesis of prostate cancer (PCa). PCa cells are heavily dependent on bioavailable iron, which frequently results in the reprogramming of iron uptake and storage pathways. Although advanced-stage PCa is currently incurable, bioactive peptides capable of modulating key iron-regulatory genes may constitute a means of exploiting a metabolic adaptation necessary for tumor growth. Recent annual increases in PCa incidence have been reported, highlighting the urgent need for novel treatments. We examined the ability of LNCaP, PC3, VCaP, and VCaP-EnzR cells to form colonies in the presence of androgen receptor inhibitors (ARI) and a series of iron-gene modulating oligopeptides (FT-001-FT-008). The viability of colonies following treatment was determined with clonogenic assays, and the expression levels of FTH1 (ferritin heavy chain 1) and TFRC (transferrin receptor) were determined with quantitative polymerase chain reaction (PCR). Peptides and ARIs combined significantly reduced PCa cell growth across all phenotypes, of which two peptides were the most effective. Colony growth suppression generally correlated with the magnitude of concurrent increases in FTH1 and decreases in TFRC expression for all cells. The results of this study provide preliminary insight into a novel approach at targeting iron dysmetabolism and sensitizing PCa cells to established cancer treatments.
Article
Full-text available
Short peptides especially marine ones are unique in their amino acids constituents that build their structure/function/specificity. Their physical and chemical properties make them diverse in their bioeffectiveness. In this study, different bioeffective peptides from marine bio-resources were investigated to map distinctive properties in their structure that could affect their function and specificity. The investigated short peptides were Pardaxins (1-5), Arenicin (1-2), Clavanin-(A-E), Aurelin, Dicynthaurin, Halocidin subunit (A-B), Tachyplesin (II-III), and Polyphemusin (I-II). Their different sequences were analyzed comparatively through alignment, their clustering (phylogeny tree), amino acids-%, and some other physical and chemical parameters. For each of them, cyclic structure based on disulfide bonds if present, and intrinsic disorder, were inspected. First alignment revealed significant diversion among their sequences, while removing unshaded amino acids from the alignment made them closer in their phylogenetic relationships and more evident to design hybrids or consensus of sequences and functions. Polyphemusin I and II had the highest charges (net) with values of +7, yet Pardaxins 1, 3, and 5 had 0 charges. All of them had higher numbers of hydrophilic amino acids than hydrophobic ones except for Aurelin, Tachyplesin II and III, and Polyphemusin I and II. Seven from twenty studied peptides had cyclic structures via disulfide bond between their cysteine residues. One amino acid was not found in all sequences; methionine. All sequences displayed other missed amino acids than methionine. Dicynthaurin is a naturally ordered peptide (via PONDR score), but Halocidin subunit B was totally disordered. This study paves the way for designing hybrids or consensus of these peptides in sequences and functions depending on their distinctive features.
Article
Medium‐chain fatty acids (MCFAs) have 6–12 carbon atoms and are instantly absorbed into the bloodstream before traveling to the portal vein and the liver, where they are immediately used for energy and may have antitumor effects. Its role in breast cancer is poorly understood. To investigate the apoptosis‐inducing effect of MCFAs in breast cancer cells, cell viability assay, colony formation assay, cell migration assay, cell invasion assay, nuclear morphology, cell cycle assay, intracellular reactive oxygen species (ROS), matrix metalloproteinase (MMP), apoptosis, RT‐qPCR analysis, and Western blot analysis were performed. In the present study, MCFA treatments reduced proliferative capability, increased ROS level, increased the depletion of MMP, induced G0/G1 and S phase cell cycle arrest, and late apoptosis of breast cancer cells in an effective concentration. Besides, MCFA treatment contributed to the upregulation of proapoptotic protein (BAK) and caspase‐3, and the downregulation of antiapoptotic protein (Bcl‐2). Mechanistically, phosphorylation levels of EGFR, Akt, and mTOR were significantly reduced in breast cancer cells treated with MCFAs. However, no significant changes in apoptosis and signaling‐related proteins were observed in lauric acid‐treated ER‐positive cancer cells. Our findings suggested that MCFAs suppressed breast cancer cell proliferation by modulating the PI3K/Akt/mTOR signaling pathway. MCFAs may be a promising therapeutic drug for treating breast cancer.
Article
Full-text available
In seeking alternative cancer treatments, antimicrobial peptides (AMPs), sourced from various life forms, emerge as promising contenders. These endogenous peptides, also known as host defense peptides (HDPs), play crucial roles in immune defenses against infections and exhibit potential in combating cancers. With their diverse defensive functions, plant-derived AMPs, such as thionins and defensins, offer a rich repertoire of antimicrobial properties. Insects, amphibians, and animals contribute unique AMPs like cecropins, temporins, and cathelicidins, showcasing broad-spectrum activities against bacteria, fungi, and viruses. Understanding these natural peptides holds significant potential for developing effective and targeted therapies against cancer and infectious diseases. Antimicrobial peptides (AMPs) exhibit diverse structural characteristics, including α-helical, β-sheet, extended, and loop peptides. Environmental conditions influence their structure, connecting to changes in cell membrane hydrophobicity. AMPs’ actions involve direct killing and immune regulation, with additional activities like membrane depolarization. In this review, we focus on antimicrobial peptides that act as anticancer agents and AMPs that exhibit mechanisms akin to antimicrobial activity. Buforin AMPs, particularly Buforin I and II, derived from histone H2A, demonstrate antibacterial and anticancer potential. Buforin IIb and its analogs show promise, with selectivity for cancer cells. Despite the challenges, AMPs offer a unique approach to combat microbial resistance and potential cancer treatment. In various cancer types, including HeLa, breast, lung, ovarian, prostate, and liver cancers, buforins demonstrate inhibitory effects and apoptosis induction. To address limitations like stability and bioavailability, researchers explore buforin-containing bioconjugates, covalently linked with nanoparticles or liposomes. Bioconjugation enhances specificity-controlled release and combats drug resistance, presenting a promising avenue for targeted cancer treatment. Clinical translation awaits further evaluation through in vivo studies and future clinical trials.
Article
Marine populations serve as a repository for novel bioactive metabolites with a wide range of chemical configurations. The influence of marine species is highlighted in this review, with a focus on marine plants, algae, bacteria, actinomycetes, fungi, sponges, and soft corals. Disease ailment patterns are shifting, and new diseases are arising as a result of changing settings. The massive increase in the world’s population has put a strain on the available drug resources. As a result, drug companies are constantly on the search for fresh resources to help them develop effective and safe drugs to meet the growing demands of the global population. The marine environment contains an abundance of various materials for developing novel medications to treat important diseases such as cancer. Cancer is still one of the deadliest diseases on the planet. New medications with novel modes of action are desperately needed, thus much research has been done on new anticancer treatments derived from natural sources, including plants, microorganisms, and marine organisms. The anti-cancer benefits of marine natural products in in vitro and in vivo research were initially discussed, as well as their activity in tumor prevention and associated compoundinduced apoptosis and cytotoxicities.
Article
Full-text available
Fermentation technology is a biorefining tool that has been used in various industrial processes to recover valuable nutrients from different side streams. One promising application of this technique is in the reclamation of nutritional components from seafood side streams. Seafood processing generates significant amounts of waste, including heads, shells, and other side streams. These side streams contain high quantities of valued nutritional components that can be extracted using fermentation technology. The fermentation technology engages the application of microorganisms to convert the side stream into valuable products like biofuels, enzymes, and animal feed. Natural polymers such as chitin and chitosan have various purposes in the food, medicinal, and agricultural industry. Another example is the fish protein hydrolysates (FPH) from seafood side streams. FPHs are protein-rich powders which could be used in animal nutrition and nutraceutical industry. The resulting hydrolysate is further filtered and dried resulting in a FPH powder. Fermentation technology holds great possibility in the recovery of valuable nutrients from seafood side streams. The process can help reduce waste and generate new value-added products from what would otherwise be considered a waste product. With further research and development, fermentation technology can become a key tool in the biorefining industry.
Article
Full-text available
Marine natural products have as of now been acknowledged as the most important source of bioactive substances and drug leads. Marine flora and fauna, such as algae, bacteria, sponges, fungi, seaweeds, cor-als, diatoms, ascidian etc. are important resources from oceans, accounting for more than 90% of the total oceanic biomass. They are taxonomically different with huge productive and are pharmacologically active novel chemical signatures and bid a tremendous opportunity for discovery of new anti-cancer molecules. The water bodies a rich source of potent molecules which improve existence suitability and serve as chemical shield against microbes and little or huge creatures. These molecules have exhibited a range of biological properties antioxidant, antibacterial, antitumour etc. In spite of huge resources enriched with exciting chemicals, the marine floras and faunas are largely unexplored for their anticancer properties. In recent past, numerous marine anticancer compounds have been isolated, characterized, identified and are under trials for human use. In this write up we have tried to compile about marine-derived compounds anticancer biological activities of diverse flora and fauna and their underlying mechanisms and the generous raise in these compounds examined for malignant growth treatment in the course of the most recent quite a long while.
Article
Full-text available
Cancer is one of the leading causes of death in the world, and antineoplastic drug research continues to be a major field in medicine development. The marine milieu has thousands of biological species that are a valuable source of novel functional proteins and peptides, which have been used in the treatment of many diseases, including cancer. In contrast with proteins and polypeptides, small peptides (with a molecular weight of less than 1000 Da) have overwhelming advantages, such as preferential and fast absorption, which can decrease the burden on human gastrointestinal function. Besides, these peptides are only connected by a few peptide bonds, and their small molecular weight makes it easy to modify and synthesize them. Specifically, small peptides can deliver nutrients and drugs to cells and tissues in the body. These characteristics make them stand out in relation to targeted drug therapy. Nowadays, the anticancer mechanisms of the small marine peptides are still largely not well understood; however, several marine peptides have been applied in preclinical treatment. This paper highlights the anticancer linear and cyclic small peptides in marine resources and presents a review of peptides and the derivatives and their mechanisms.
Article
Full-text available
Breast cancer is a malignant tumor that occurs in the epithelial tissue of the breast, mostly in female. At present, drug resistance has emerged in the treatment of breast cancer. Therefore, the discovery of new drugs for breast cancer is particularly important. Some peptides have been found to have anti-cancer effects. This article reviews the recent discoveries of anti-breast cancer peptides, hoping to provide some help for the development of breast cancer treatment.
Article
Full-text available
The marine environment is a rich source of biologically active molecules for the treatment of human diseases, especially cancer. The adaptation to unique environmental conditions led marine organisms to evolve different pathways than their terrestrial counterparts, thus producing unique chemicals with a broad diversity and complexity. So far, more than 36,000 compounds have been isolated from marine micro-and macro-organisms including but not limited to fungi, bacteria, microalgae, macroalgae, sponges, corals, mollusks and tunicates, with hundreds of new marine natural products (MNPs) being discovered every year. Marine-based pharmaceuticals have started to impact modern pharmacology and different anti-cancer drugs derived from marine compounds have been approved for clinical use, such as: cytarabine, vidarabine, nelarabine (prodrug of ara-G), fludarabine phosphate (pro-drug of ara-A), trabectedin, eribulin mesylate, brentuximab vedotin, polatuzumab vedotin, enfortumab vedotin, belantamab mafodotin, plitidepsin, and lurbinectedin.
Article
Full-text available
Since the early research efforts focusing on bioactive marine substances such as spongouridine, from the marine sponge Cryptotethya crypta, marine natural products (MNPs) have arisen as a robust and sustainable supplier for bioactive drug leads. Marine natural products present definite, unprecedented structural diversifications and varieties of interesting biomedical potentialities with novel mechanisms of action. Until today, eight clinically approved marine natural products-based drugs by two worldwide medical organizations (including U.S. FDA, European Medicines Agency (EMEA)), have been developed for the treatment of different forms of carcinoma, pain, Alzheimer’s disease and other current medical challenges. Recent clinical trial analysis disclosed that the current clinical pipeline contains more than twenty listed drug candidates in different clinical trials in phase III, II, or I. Herein, we present recent insights centered to clinically approved and preclinically investigated marine bioactive compounds, as well as sampling techniques, extraction & identification tools, classification of MNPs, some reported biological activities, and challenges faced during MNPs development.
Article
Full-text available
Bioactive peptide with anti-proliferative activity is a novel prospect to develop cancer drugs with fewer or no side effects, as well as cost benefits. The current review provides a development strategy for anti-proliferative peptides derived from fish and its discards through fish protein hydrolysate. The study indicates that enzymatic hydrolysates act as a superior anti-proliferative agent through superior enzyme selection. In general, ion-exchange chromatography, RP-HPLC, ultra, and gel filtration methods are used to purify bioactive peptides. Characterization of the purified peptide can be done with the help of cytotoxicity assay, molecular mass, amino acid composition, and sequence. Most studies on anti-proliferative bioactive peptides have been carried out via in vitro analysis. However, very few studies have provided confirmatory results through in vivo analysis.
Article
Full-text available
Cancer is one of the most extreme medical conditions in both developing and developed countries around the world, causing millions of deaths each year. Chemotherapy and/or radiotherapy are key for treatment approaches, but both have numerous adverse health effects. Furthermore, the resistance of cancerous cells to anticancer medication leads to treatment failure. The rising burden of cancer overall requires novel efficacious treatment modalities. Natural medications offer feasible alternative options against malignancy in contrast to western medication. Furanocoumarins' defensive and restorative impacts have been observed in leukemia, glioma, breast, lung, renal, liver, colon, cervical, ovarian, and prostate malignancies. Experimental findings have shown that furanocoumarins activate multiple signaling pathways, leading to apoptosis, autophagy, antioxidant, antimetastatic, and cell cycle arrest in malignant cells. Additionally, furanocoumarins have been shown to have chemo preventive and chemotherapeutic synergistic potential when used in combination with other anticancer drugs. Here, we address different pathways which are activated by furanocoumarins and their therapeutic efficacy in various tumors. Ideally, this review will trigger interest in furanocoumarins and their potential efficacy and safety as a cancer lessening agents.
Article
Full-text available
Anti-cancer peptides (ACPs) are a series of short peptides composed of 10–60 amino acids that can inhibit tumour cell proliferation or migration, or suppress the formation of tumour blood vessels, and are less likely to cause drug resistance. The aforementioned merits make ACPs the most promising anti-cancer candidate. However, ACPs may be degraded by proteases, or result in cytotoxicity in many cases. To overcome these drawbacks, a plethora of research has focused on reconstruction or modification of ACPs to improve their anti-cancer activity, while reducing their cytotoxicity. The modification of ACPs mainly includes main chain reconstruction and side chain modification. After summarizing the classification and mechanism of action of ACPs, this paper focuses on recent development and progress about their reconstruction and modification. The information collected here may provide some ideas for further research on ACPs, in particular their modification.
Article
Marine peptides are one of the richest sources of structurally diverse bioactive compounds and a considerable attention has been drawn towards their production and bioactivity. However, there is a paucity in consolidation of emerging trends encompassing both production techniques and biological application. Herein, we intend to review the recent advancements on different production, purification and identification technologies used for marine peptides along with presenting their potential health benefits. Bibliometric analysis revealed a growing number of scientific publications on marine peptides (268 documents per year) with both Asia (37.2 %) and Europe (33.1 %) being the major contributors. Extraction and purification by ultrafiltration and enzymatic hydrolysis, followed by identification by chromatographic techniques coupled with an appropriate detector could yield a high content of peptides with improved bioactivity. Moreover, the multifunctional health benefits exerted by marine peptides including anti-microbial, antioxidant, anti-hypertension, anti-diabetes and anti-cancer along with their structure-activity relationship were presented. The future perspective on marine peptide research should focus on finding improved separation and purification technologies with enhanced selectivity and resolution for obtaining more novel peptides with high yield and low cost. In addition, by employing encapsulation strategies such as nanoemulsion and nanoliposome, oral bioavailability and bioactivity of peptides can be greatly enhanced. Also, the potential health benefits that are demonstrated by in vitro and in vivo models should be validated by conducting human clinical trials for a technology transfer from bench to bedside.
Article
Introduction The use of nanoparticles for breast cancer targeting and treatment has become a reality. They have shown to be safer and to possess interesting peculiarities such as the unspecific accumulation to the tumor site and the possibility to activate a tight control of drug release as compared to free drugs. However, there are still many areas of improvement which can certainly be addressed with the use of peptide-based elements. Areas covered The article reviews different preclinical strategies employing peptides and proteins in combination with nanoparticles for breast cancer targeting and treatment as well as peptide and protein-targeted encapsulated drugs, and it lists the current clinical status of therapies using peptides and proteins for breast cancer. Expert opinion The conjugation of protein and peptides can improve tumor homing of nanoparticles, increase cellular penetration and attack specific drivers and vulnerabilities of the breast cancer cell to promote tumor cytotoxicity while reducing secondary effects in healthy tissues. Examples are the use of antibodies, arginylglycylaspartic acid (RGD) peptides, membrane disruptive peptides, interference peptides, and peptide vaccines. Although their implementation in the clinic has been relatively slow up to now, we anticipate great progress in the field which will translate into more efficacious and selective nanotherapies for breast cancer.