• No results found

A newly isolated strain of Halomonas sp. (HA1) exerts anticancer potential via induction of apoptosis and G(2)/M arrest in hepatocellular carcinoma (HepG2) cell line

N/A
N/A
Protected

Academic year: 2022

Share "A newly isolated strain of Halomonas sp. (HA1) exerts anticancer potential via induction of apoptosis and G(2)/M arrest in hepatocellular carcinoma (HepG2) cell line"

Copied!
15
0
0

Loading.... (view fulltext now)

Full text

(1)

www.nature.com/scientificreports

A newly isolated strain of Halomonas sp. (HA1) exerts

anticancer potential via induction of apoptosis and G 2 /M arrest

in hepatocellular carcinoma (HepG2) cell line

islam M. el‑Garawani

1*

, Sabha M. el‑Sabbagh

2

, nasser H. Abbas

3

, Hany S. Ahmed

2

, omaima A. eissa

2

, Doaa M. Abo‑Atya

4

, Shaden A. M. Khalifa

5

& Hesham R. el‑Seedi

4,6,7*

Marine bacterial strains are of great interest for their ability to produce secondary metabolites with anticancer potentials. Isolation, identification, characterization and anticancer activities of isolated bacteria from El-Hamra Lake, Wadi El-Natrun (Egypt) were the objectives of this study. The isolated bacteria were identified as a moderately halophilic alkaliphilic strain. Ethyl acetate extraction was performed and identified by liquid chromatography-mass spectrophotometry (LC–MS–MS) and nuclear magnetic resonance analysis (NMR). Cytotoxicity of the extract was assessed on the HepG2 cell line and normal human peripheral lymphocytes (HPBL) in vitro. Halomonas sp. HA1 extract analyses revealed anticancer potential. Many compounds have been identified including cyclo-(Leu- Leu), cyclo-(Pro-Phe), C17-sphinganine, hexanedioic acid, bis (2-ethylhexyl) ester, surfactin C14 and C15. The extract exhibited an IC

50

of 68 ± 1.8 μg/mL and caused marked morphological changes in treated HepG2 cells. For mechanistic anticancer evaluation, 20 and 40 µg/mL of bacterial extract were examined. The up-regulation of apoptosis-related genes’ expression, P53, CASP-3, and BAX/

BCL-2 at mRNA and protein levels proved the involvement of P53-dependant mitochondrial apoptotic pathway. The anti-proliferative properties were confirmed by significant G

2

/M cell cycle arrest and PCNA down-regulation in the treated cells. Low cytotoxicity was observed in HPBL compared to HepG2 cells. In conclusion, results suggest that the apoptotic and anti-proliferative effects of Halomonas sp. HA1 extract on HepG2 cells can provide it as a candidate for future pharmaceutical industries.

Abbreviations

NMR Nuclear magnetic resonance

LC–MS–MS Liquid chromatography-mass spectrophotometry

MeOH Methanol

DMSO Dimethyl sulphoxide

open

1

Department of Zoology, Faculty of Science, Menoufia University, Menoufia 32511, Egypt.

2

Department of

Botany and Microbiology, Faculty of Science, Menoufia University, Menoufia 32511, Egypt.

3

Department of

Molecular BiologyGenetic Engineering and Biotechnology Research Institute, University of Sadat City, Sadat

City 32958, Egypt.

4

Department of Chemistry, Faculty of Science, Menoufia University, Menoufia 32511,

Egypt.

5

Department of Molecular Biosciences, The Wenner-Gren Institute, Stockholm University, 10691 Stockholm,

Sweden.

6

Pharmacognosy Group, Department of Pharmaceutical Biosciences, Uppsala University, Biomedical

Centre, 75 123 Uppsala, Sweden.

7

International Research Center for Food Nutrition and Safety, Jiangsu University,

Zhenjiang 212013, China.

*

email: dr.garawani@yahoo.com; hesham.el-seedi@ilk.uu.se

(2)

www.nature.com/scientificreports/

HPBL Normal human peripheral lymphocytes BCL-2 B-cell lymphoma-2 proteins

BAX BCL-2-associated X protein

Hepatocellular carcinoma (HCC) is the fifth type of cancer in patients with chronic liver disease and cirrhosis and the second cause of death in cancer patients

1

. There are nearly 800,000 deaths annually due to HCC

2

, and the HCC’s prevalence has gradually increased over the past decade

3

. Cancer treatment can involve surgical inter- vention, radiation, chemotherapy, targeted therapy and/or other therapies

4

. Despite the extreme efforts to treat HCC effectively, side effects are involved, such as immunodeficiency, cell damage, and neurological, renal, and cardiac toxicity

5–7

. Currently, the pharmaceutical industry and research are focused on developing and applying new strategies to lower the side effects and enhance the economic value

8,9

. Therefore, finding and developing natural products-derived anticancer drugs with good efficacy and low toxicity

10,11

from different sources such as essential oils

12,13

, certain types of mushrooms

14

or even plant crude extracts

15,16

are very crucial.

Microbial metabolites are among the most important natural cancer chemotherapeutics

17

. With actinomycin discovery, they began to appear around 1940, and since then, most anticancer compounds have been derived from natural sources

18

. More than 60 percent of the existing compounds with antineoplastic activities are initially either identified from or inspired by natural compounds and their derivatives

19

. Bacteria are the greatest produc- ers of bioactive products and thus of immense importance for the drugs discovery

20

. Marine pharmacology is a new discipline that explores the potential pharmaceutical products that originated in the marine world. In the last two decades, extensive screening of marine compounds and their antivirals

21

, antibacterial

22

, antifungal

23

, antiparasitic

24

, antitumor

25

, and anti-inflammatory

26

activities have been reported. Systematic studies of various soda lakes have shown that microorganisms, which thrive in these areas, are adapted to the extreme pH and salt conditions, many of which are alkaliphilic and halophilic or extremely halotolerant, and many represent separate alkaliphilic lines within the recognized taxa. El-Hamra Lake, Wadi El-Natrun (Egypt) has a high water salinity (300 g/L) and pH 10.0. In both saline and alkaline conditions, haloalkaliphilic bacteria can be present and the preliminary results prompted the identification of the metabolites produced by alkaliphilic bacteria

27

. Marine bacteria have genes such as PKS (polyketides) and NRPS (natural products) that are characteristic of secondary metabolite production. Halomonas strains from a marine origin may generate siderophores (low molecular weight Fe

3+

chelating molecules) and loihichelins

28

. The progression of gastric adenocarcinoma cell lines (HM02), hepatocellular carcinoma (HepG2) and breast cancer (MCF-7) has been inhibited by apoptosis initiation and cell cycle arrest when treated with marine-derived Halomonas sp. (GWS-BW-H8hM strain)

29,30

. However, to the best of our knowledge, the novel strain has not been studied yet. Consequently, the characteriza- tion of bioactive compounds and evaluation of the possible anticancer potential of the bacterial extract against the HepG2 cell line were warranted.

Results

Identification of bacterial strain. Morphological investigations showed two isolates of gram-negative short motile rods. Full-length 16S rRNA (about 1,500 bp) was sequenced and deposited under GenBank acces- sion numbers (KT223026) for H. HA1 isolate. The isolated H. HA1 16S rRNA sequence showed 99% similarity with the 16S rRNA gene sequence of Halomonas denitrificans BC7. However, the isolated strain is closer to Halomonas nitroreducens evolutionary group (Fig. 1).

LC–MS-MS and NMR fraction analyses of H. HA1 extract. The molecular networking of metabo- lome mass profile for the species Halomonas sp.HA1 was screened and exhibited in 67 parent ions (nodes) (Fig. 2).

The number of nodes was considered an indication of the unique peaks. Compounds with related molecular

weight and shared class grouped together to form a cluster. Four clusters are noticed with some chemically

Figure 1. The phylogenetic tree based on 16 s rRNA sequences constructed by the neighbor-joining method,

showing the position of strain HA1 and representatives of some related taxa.

(3)

www.nature.com/scientificreports/

related identified compounds. From the 67 nodes only 15 parent ions within the molecular network matched 15 known standards from the molecular networking database but three of them in the yellow nodes [M+H]

+

(m/z; 200.128, 228.164 and 393.265) have wrong precursor parent ions, so their identification can’t be considered (Fig. 2 and Table 1).

Six identified dipeptides (Ile-Pro, cyclo-(L-pro-L-val), cyclo-(Pro-Phe), Pro-Tyr, cyclo-(Leu-Leu) and cyclo- [L-(4-hydroxy-Pro)-L-leu]) with [M + H]

+

(211.06, 197.064, 245.042, 261.038, 227.115 and 227.058) respectively Figure 2. Molecular network of 67 parent ions produced from Halomonas sp. HA1. The fuchsia nodes indicate to the whole molecular weights that have unique detected peaks in the molecular network. The green Nodes represent parent ions that are identified from the molecular networking database as they have been already isolated before. The yellow nodes indicate some identified compounds but with wrong precursor parent ions so they should not be recognized.

Table 1. The parent and the fragments’ masses of the identified compounds compared with that of the standards from the molecular networking database.

Compound name Precursor mass (m/z) Database fragments Identified fragments

Cyclo-(L-pro-L-val) 197.064 179.12, 169.14, 152.11, 141.14, 124.12 and 70.07 –

Ile-Pro 211.06 – 183.128, 138.125, 114.102, 98.065, 86.105 and 70.064

Cyclo-[L-(4-hydroxy-Pro)-L-leu] 227.058 199.15, 181.14, 153.15, 136.12 and 86.08 –

Cyclo-(Leu-Leu) 227.115 – –

Cyclo (Pro-Phe) 245.042 217.69, 120.03 and 70.79 154.100, 120.10 and 70.064

Palmitamide 256.21 – –

Pro-Tyr 261.038 243.06, 233.09, 215.04, 135.92, 119.94 and 85.94 187.183, 120.085 and 86.056

9-Octadecenamide 282.205 265.27, 247.30, 149.10 and 135.13 –

C17-Sphinganine 288.218 – –

Hexanedioic acid, bis(2-ethylhexyl) ester 371.135 – –

Surfactin C14 1,022.48 909.46, 891.41, 86e3.38, 685.37, 667.44, 582.31, 554.33

and 441.23 -

Surfactin C15 1,036.52 1,018.47, 923.58, 905.42, 877.39, 808.44, 792.44,

695.38, 685.37, 667.35, 596.41, 554.27 and 441.24 923.574, 685.500, 596.445, 483.382, 441.298, 326.238, 272.167 and 86.088

(4)

www.nature.com/scientificreports/

are clustered together (Fig. 2 and Table 1). Palmitamide and 9-octadecenamide with [M + H]

+

(256.21 and 282.205) (Fig. 2 and Table 1), are two identified fatty acid amides produced as endogenous substances. One sphingolipid (aminodiol) that is C17-sphinganine with m/z 288.218 [M+H]

+

(Fig. 2 and Table 1), was isolated from different fungus species

31

. C17-sphingolipid identified as mycotoxin (C17-SAMT) analog exerted potent toxicity in different assays

32,33

. The most potent identified compounds were two biosurfactants (Surfactin C14 and surfactin C15) with m/z 1,022.48 and 1,036.52 [M+H]

+

(Fig. S1).

The parent ion masses fragmentation data for Surfactin C15 and Surfactin C14 from the molecular network- ing database were compared (Figs. S1,S2).

Surfactin C15 parent ion fragmentations are recognized from MS/MS chromatogram and the fragment with [M+H]

+

m/z 685.500 represents the base peak ion (Fig. S3).

Surfactin C15 fragmentation mechanism in the positive ion mode starts by loss of fragment ion [M+]

+

m/z 113.13 and C

8

H

17•

then continuous cleavage of different peptide bonds and loss of different amino acid fragments to reach the last fragment with [M+H]

+

(m/z 85.09) and molecular formula C

5

H

11

N

2·

(Fig. S4).

Anticancer activities. Cytotoxicity on HepG2 cells. The bacterial extract has been examined for their an- ticancer activity against HepG2 cell line using MTT assay. Results, as shown in Fig. 3, reflected a strong cytotoxic potential of the extract with 68 ± 1.8 μg/mL maximal inhibitory concentration (IC

50

).

Assessment of morphological changes in HepG2. Results of the inverted phase-contrast examination demonstrated the remarkable anticancer potentials as evidenced by cellular shrinkage and irregular cell shapes following bacterial extract treatments in a dose-dependent manner when compared with control groups (Fig. 4).

Furthermore, AO/EB staining revealed cytoplasmic vacuolation, membrane blebbing and irregular nuclear morphology as hallmarks of apoptosis and cell death. The records of abnormal morphology were significant (P

< 0.05) with H. HA1 extract incubations compared to control HepG2 cells (Fig. 5).

flow cytometric analysis of cell cycle distribution. Cell cycle analysis revealed significant (P < 0.05) G

2

/M cell cycle arrest in H. HA1 crude extract-treated HepG2 cells (Fig. 6A). Results showed remarkable accu- mulation of G

2

/M populations in treated cells with 18.42 and 24.18% for 20 and 40 µg/mL respectively when compared to 13.45% in untreated cells. However, cisplatin induced significant (P < 0.05) apoptosis particularly in S phase when compared with untreated cells (Fig. 6B).

immunocytochemical assessment of apoptosis and proliferation‑related proteins. Changes in protein expression of P53, CASP-3, and BAX/BCL-2 ratio were examined in treated cells and controls. Results revealed that H. HA1 extract induced apoptosis as confirmed by P53 and CASP-3 significant (P < 0.05) up-reg- ulation of protein expression. The evaluated immunocytochemical reactivities showed an increase of ~ 5 and 8 folds, in P53 protein level, for 20 and 40 µg/mL respectively, relative to the control. The level of CASP-3 (procas- pase-3) protein was increased by ~ 11 and 6 folds, for 20 and 40 µg/mL respectively, when compared to untreated cells. Moreover, BAX/BCL-2 ratio exhibited a significant (P < 0.05) concentration-dependent elevation when compared to the control confirming that the induction of apoptosis was via the intrinsic mitochondrial pathway (Fig. 7). However, PCNA, a nuclear proliferating marker, showed a significant (P < 0.05) down-regulation in H.

Figure 3. The cytotoxic effect of the H. HA1 extract on HepG2 using MTT assay after 24 h. Incubation with

serial concentrations of the extract showed a potent anticancer effect with an IC

50

: 68 ± 1.8 µg/mL. Data were

represented as (Mean ± SD) of three different experiments (n = 3). H. HA1: Halomonas. HA1.

(5)

www.nature.com/scientificreports/

HA1 (7.3 and 15.7% for 20 and 40 µg/mL, respectively) and cisplatin-treated cells (13.7%) when compared with the untreated group (91.7%).

Altered mRNA expression of apoptosis markers. The apoptotic effect of H. HA1 crude extract on HepG2 was confirmed by apoptosis-related genes expression. The mRNA expression was quantified by qRT- Figure 4. Representative photomicrographs of control and treated HepG2 cells after 24 h. Cellular

morphological alterations and detached cells were observed by phase contrast light inverted microscope (Olympus IMT-2, Japan). H. HA1: Halomonas. HA1; Cisplatin (3 µg/mL).

Figure 5. Representative photomicrographs of cellular alterations in untreated and treated HepG2 cells after 24 h of different incubations. Observable decrease in viability of cells and increased morphological alterations towards apoptosis induction such as membrane blebbing (white arrows) were noticed after acridine orange/

ethidium bromide (AO/EB) double fluorescent labeling (A). Data represent the means of three independent experiments; bars, standard deviation and a: significant (P < 0.05) with respect to the control (B). H. HA1:

Halomonas. HA1; Cisplatin (3 µg/mL).

(6)

www.nature.com/scientificreports/

PCR for P53, CASP-3, BAX and BCL-2 genes in the control and treated cells. Results of P53 and CASP-3 showed significant up-regulation among H. HA1 treated cells (~ 6 and 2.5 folds, for 20 and 40 µg/mL respectively). BAX/

BCL-2 ratio exhibited a significant (P < 0.05) elevation with H. HA1 crude extract treatment (~ 0.95 and 1.2 folds, for 20 and 40 µg/mL, respectively) compared to the untreated control. These results suggest that H. HA1 extract induced apoptosis in HepG2 cells via the intrinsic mitochondrial pathway. The lower concentration of H.

HA1 (20 µg/mL) was more potent in apoptosis induction than the higher one (Fig. 8).

Effect of the extract on normal human peripheral lymphocytes. Assessment of cytotoxicity on nor- mal HPBL. HPBLs were treated with H. HA1 crude extract for 24 h. Trypan blue exclusion method (Fig. 9) and AO/EB dual fluorescent staining (Fig. 10) were performed. Results revealed low cytotoxicity and morphological changes among extract-treated groups compared to higher toxicity of mitomycin C-treated cells (Figs. 9, 10B).

Assessment of DNA single-strand breaks (comet assay). For DNA single-strand breaks assessment in normal cells, HPBLs were treated with H. HA1 crude extract for 24 h. The comet assay was carried out (Fig. 11A). Results revealed significant (P < 0.05) genotoxic effect with high concentration (40 µg/mL) of H. HA1 extract as well as mitomycin C-treated group when compared to control cells (Fig. 11B).

Discussion

Currently, natural product extracts are considered to be the most promising source of new drugs for cancer

34

. Several studies have indicated the advantages of marine flora and fauna extracts in combating cancer and a num- ber of other diseases

35–37

. Bacteria are the greatest producers of bioactive natural products and are of immense importance for drug discovery

20

. In this study, the newly identified strain of bacteria Halomonas sp. HA1 was investigated for its anticancer potential against HepG2 cells. In agreement with some earlier studies on some microbial extracts, the results clearly indicated that the crude extract of Halomonas sp. HA1 showed a significant anticancer effect on HepG2 cells associated with decreased viability and morphological alterations of the treated cells leading eventually to cell death

29,30,38

. To investigate the pro-apoptotic pathway of H. HA1 bacterial extract, the changes in apoptosis regulatory genes upon applying the extract on HepG2 cells were evaluated.

Apoptosis is a process involving alterations in the expression of an array of genes. BCL-2 family plays a crucial role in apoptosis regulation

39

. However, one of the early apoptosis hallmark changes is a reduction in mitochondrial membrane potential. The potential decrease of the mitochondrial membrane found as a result of mitochondrial membrane depolarization indicated the mitochondrial damage. Another factor in the process Figure 6. Effect of different concentrations of H. HA1 crude extract incubation for 24 h on cell cycle

distribution of treated and control HepG2 cells. The cell cycle phases were analyzed according to DNA content as evaluated after propidium iodide (PI) labeling (B). Data represent means of three different experiments; bars, standard deviation; and a, significant increase (P < 0.05) relative to untreated cells (A). H. HA1: Halomonas.

HA1; Cisplatin (3 µg/mL).

(7)

www.nature.com/scientificreports/

Figure 7. Effects of crude extract incubation for 24 h on the positively stained cells in control and treated HepG2 cells. Increased percentage of P53, CASP-3 positive immuno-reactive cells and elevated BAX/BCL-2 ratio as well as decreased levels of PCNA were noticed in treated cells. Data represent means of five different fields; bars, standard deviation; and a, significant (P < 0.05) relative to untreated cells. H. HA1: Halomonas. HA1;

Cisplatin (3 µg/mL).

Figure 8. Effect of H. HA1 crude extract on relative mRNA alterations of P53, CASP-3 and BAX/BCL-2. Data

(means of three experiments) were standardized to the level of GAPDH mRNA and expressed as fold-induction

relative to the control mRNA levels. Bars, standard deviation and a: significant (P < 0.05) compared to untreated

cells. H. HA1: Halomonas. HA1; Cisplatin (3 µg/mL).

(8)

www.nature.com/scientificreports/

of mitochondrial apoptosis pathway after mitochondrial membrane destruction is the release of pro-apoptotic protein factors from the mitochondria

40

.

In the present study, BAX/BCL-2, CASP-3 (as a procaspase-3) and P53 were assessed at the transcriptional and protein levels. The ratio of BAX/BCL-2 mRNA expression recorded a significant increase as well as CASP-3 and P53 up-regulation in treated HepG2 cells. It is well-known that BAX is up-regulated by P53 protein while Figure 9. Effect on HPBL after incubation with various concentrations of H. HA1 extract, for 24 h. Cells were stained with trypan blue (TB) for cytotoxicity assessment. Data were represented as Mean ± SD of three independent experiments (n = 3). Bars, standard deviation and a: significant (P < 0.05) compared to the untreated cells. H. HA1: Halomonas. HA1; Mitomycin C (0.5 µg/mL).

Figure 10. Effect on HPBL after incubation with various concentrations of HL extract, for 24 h. Cells were

labeled with AO/EB fluorescent staining for cell death (A). Data were represented as Mean ± SD of three

independent experiments (B). Bars, standard deviation and a: significant (P < 0.05) compared to the untreated

cells (n = 3). H. HA1: Halomonas. HA1; Mitomycin C (0.5 µg/mL).

(9)

www.nature.com/scientificreports/

BCL-2 protein expression is down-regulated by P53

41

. P53 is a nuclear transcriptional factor that usually is activated in apoptosis by regulating numerous down-stream effectors

42

. Over-expression of BAX (pro-apoptotic protein)

43

and inhibition of BCL-2 protein expression can speed up cell apoptosis. Apoptosis can be induced by P53-mitochondrial localization in a transcription-independent manner

44,45

or by induction of endoplasmic reticulum stress to prevent P53-dependent apoptosis via the glycogen synthase kinase-3β pathway

46

. The expres- sion profile of H. HA1 extract indicated the P53-dependant mitochondrial apoptotic pathway. These results are in agreement with Ruiz-Ruiz et al.

47

, suggesting the anti-proliferative effect of the novel halophilic Halomonas sp. extract. The results of CASP-3 expression profile exhibted caspase-dependent apoptosis. However, due to the elevated toxicity and stress of the higher concentration of the extract, treated cells may tend to undergo apoptosis via caspase-independent pathway

48

owned to the processes of autophagy, endoplasmic reticulum stress or mitotic catastrophe

49

. Further, the different biological and technical factors, such as transcriptional or post-translational alterations as well as a long half-life of some proteins, can affect the association between mRNA expression and the level of proteins

50

. The treatments may cause internal cellular dysfunctions which lead to the degradation of mRNA molecules faster than the protein ones

51

. Although the cleaved caspase-3 has not been assessed in this study, the morphological apoptotic features such as membrane blebbing, loss of membrane integrity, nuclear fragmentation and elevated BAX/BCL-2 ratio suggested the occurrence of apoptosis

52

. Moreover, the loss of membrane integrity, confirmed by AO/EB labeling, is a late event in apoptosis. Caspase-3 is one of the death substrates that serve as an effector in apoptosis inducing phosphatidylserine (PS) externalization, shrinkage of cell, membrane blebbing, and DNA fragmentation by cleaving

53,54

. In the absence of caspase-3, the cells mor- phological features were impaired

55

and cytochrome c release was also delayed

56

. Due to apoptosis induction, the decreased value of CASP-3 protein in the higher concentration of the extract may be due to the increase of the expected cleavage process occurred in the procaspase-3 protein. However, the CASP-3 cleavage assessment is needed in the future studies. Collectively, the explanation of the different transcriptional level of CASP-3 in comparison to its protein level as well as the decreased levels of up-regulated P53 and CASP-3 in the higher concentration-treated cells could be proved.

Moreover, the analysis of cell cycle distribution revealed accumulations of sub-G

0

(apoptotic cells) and signifi- cant G

2

/M arrest in addition to the down-expression of PCNA (proliferating-cell nuclear antigen) protein. The expression of PCNA protein regulates the cell cycle and promotes apoptosis in fetal and neonatal mouse ovaries

57

. Further, the arrest of G

2

/M phase may enable damaged cells to undergo apoptosis

58

. However, cisplatin-induced arrest in sub-G

0

and S phases (DNA synthesis) eventually led to DNA damage and apoptosis

59

. The down- regulation of PCNA, the anti-proliferative effects and the cell cycle arrest are key players of the apoptotic cascade.

The results suggested the anti-proliferative and apoptotic effects of Halomonas sp. HA1 extract can be attrib- uted to the secondary metabolites identified by LC–MS-MS and NMR analysis of Halomonas sp. HA1 extract.

However, sphinganine is the backbone precursor of sphingolipids

60

. Sphingolipids are produced in most eukar- yotic cells but rarely in prokaryotes, which possess only glycerol-based phospholipids in their membrane

61

. Figure 11. Effect of H. HA1 extract on DNA single-strand breaks (alkaline comet assay) in HepG2 cells stained with ethidium bromide (A). Data were represented as Mean ± SD of three independent experiments (B). Bars, standard deviation and a: significant (P < 0.05) compared to the untreated cells (n = 3). H. HA1: Halomonas.

HA1; Mitomycin C (0.5 µg/mL).

(10)

www.nature.com/scientificreports/

However, a few isolated bacterial species produce sphingolipids

62–64

. A previous study reported that, both cyclic dipeptides cyclo-(prolyltyrosyl) and cyclo-(prolylphenylalanyl) isolated from Bacillus spp. showed potential anticancer activity

65

. Another study showed that a CDP mix composed of cyclo-(L-Pro-L-Tyr), cyclo-(L-Pro-L- Val), and cyclo-(L-Pro-L-Phe), isolated from the P. aeruginosa PAO1 strain, exhibited anticancer activity in HeLa and Caco-2 cell lines

66

. In another study cyclo-(-Pro-Tyr) exhibited anticancer and anti-proliferative activity against HepG2 cell lines

67

. A previous study has proven the anticancer activity of 1, 2-Benzenedicarboxylic acid, bis (2-ethylhexyl) ester on PC3, MCF, HCT-116, A549, and MIAPACA cell lines

68

. Surfactin biosynthesis occurs in many natural microorganisms i.e. Bacillus pumilus, B. mojavensis, B. circulans, B. natto, and B. amylolique- faciens

69

; B. safensis F4 and B. subtilis strains

70,71

. It exhibits powerful therapeutic activities i.e. antimicrobial, antimycoplasma, antiviral, anti-inflammatory, antibacterial and antitumor

10,70,72–75

. Additionally, it is widely applicable environmentally in biopesticides, food processing, pharmaceuticals, cosmetics and oil recovery

76–78

.

In fact, surfactin mediated a forceful anticancer activity on various cancer cell lines. It showed anti-prolif- erative potentials on human breast cancer cells (MCF-7)

79

, cervical cancer (HeLa) and hepatoma (Bel-7402)

80

. An in vivo anticancer activities of surfactin on mice ascites tumors were also reported

81

. It also inhibited the growth of LoVo colon cancer cells, with IC

50

of 26 µM at 48 h

82

. The anticancer activity of surfactin is attributed to its amphiphilic nature. Lipopeptides including surfactin exert cytotoxicity on cancer cells with different IC

50

depending on the length of fatty acid chains

80

. Therefore, surfactin C15 should exhibit higher anticancer activity than surfactin C14 and it is supposed to be responsible for the anticancer activity of the extract.

In conclusion, we reported for the first time that, Halomonas sp. HA1 extract has a potent anticancer activity on HepG2 cells via anti-proliferative and apoptotic potentials. Surfactin C14 and Surfactin C15, biosurfactants, were identified in Halomonas sp. HA1 extract by LC–MS-MS and NMR analyses. However, further in vivo studies are needed to investigate the metabolism and the impact of biotransformation on the efficacy of H. HA1 extract’s active constituents as anticancer agents.

Methods

isolation of haloalkaliphilic bacteria. Water samples were collected from El-Hamra Lake in Wadi El- Natrun, Egypt where the pH was 10.0 and water salinity was 300 g/L. Halophile growth medium (HGM) was prepared according to Dorn et al.

83

, and pH 9 was adjusted using NaHCO

3

. The media were supplemented with 3 M NaCl and 2.0 mL 10% (w/v) yeast extract per liter. The water sample was added to HGM media with 1:10 (v/v) and incubated at 30 °C for 48 h. Two more successive culturing were carried out using 50 µL of diluted culture spread on an agar plate of the same media. Single colonies were plated on new agar plates with the same media for biochemical characterization.

Phylogenetic analysis of the 16S ribosomal RNA gene. Genomic DNA was extracted from the pure culture using GeneJET Genomic DNA Purification Kit (Thermo Fisher Scientific, US). PCR amplification of the 16S rRNA gene carried out using Bact 27f. (5′-AGA GTT TGATC (A/C)-TGG CTC AG-3′), Bact 1492r (5′-TACGG(C/T)-ACC TTG TTA CGA CTT-3′), and Bact 1098r (5′-AAG GGT TGC GCT CGT TGC G-3′)

84

. Using a 3,100 Genetic Analyzer (Applied Biosystems, US), the amplified products were sequenced according to the manufacturer’s protocols. The obtained sequence of 16S rRNA gene was compared using the BLASTN program against the nucleotide sequences collection (nr/nt) database, available at the National Center for Biotechnol- ogy Information website (https ://blast .ncbi.nlm.nih.gov/Blast .cgi). Phylogenetic tree relative to high scoring BLASTP hits was performed using the (MEGA 7.0.26) software.

The complete sequence of 1, 2 CDG of Halomonas was isolated using universal primer UDOG: (5′-ATG ACT GTT AAA ATT TAT GAC ACC CCT GAA G-3′) and reversal primer RDOG: (5′-TTA TGG ACG CGC TTG CAG CTC-3′). Primers were designed according to sequence alignment of 1, 2 DOG NCBI database sequences. The PCR conditions were 94 °C for 30 s, 60 °C 30 s and 72 °C for 1 min. Using a 3100 Genetic Analyzer (Applied Biosystems, US), PCR products were sequenced following the manufacturer’s protocols. The obtained sequence of 16S rRNA gene was compared using the BLASTN program against the nucleotide sequences collection database, available at the NCBI website (https ://blast .ncbi.nlm.nih.gov/Blast .cgi). The phylogenetic tree relative to high scoring BLASTP hits was performed using the (MEGA 7.0.26) software.

Preparation of crude extract. The bacterial isolate was cultured in (HGM) medium and incubated in shaker incubator 100 rpm at 30 °C for 72 h. The culture medium was centrifuged at 6,500 rpm for 20 min. The supernatant was collected and filtered through a 0.45 µm sterile membrane. Ethyl acetate was added to culture filtrate (1:1) and stirred at 130 rpm for 12 h. the ethyl acetate phase on the upper part of the culture was removed and followed by vacuum evaporation to obtain the dry extracts by vacuum rotary evaporator 40 °C

85

. Crude extracts produced were stored at − 20 °C for further investigations.

Characterization of bacterial extract using NMR and LC–MS-MS. Generally, 1H and 13C NMR spectra were recorded on a Bruker DRX 600 spectrometer at 600.1 and at 150.9 MHz, respectively, at 25 °C using CD3OD as the solvent. All chemical shifts are expressed relative to TMS

86

.

For the LC–MS analysis the crude extract was dissolved in 50% acetonitrile (ACN), 0.1% formic acid (FA).

The samples were injected by syringe through a PicoTip emittor at 0.3 µL/min connected to a Q-Tof Micro (Waters, Milford, Massachusetts, US) with the voltage set at 1.4 kV. The analysis was carried out in positive ion mode and linear gradient from 10% (v/v) H

2

O to 99% (v/v) ACN in 0.1% (v/v) FA at a flow rate of 0.3 µL/min over 75 min. The analysis was followed by LC/MS–MS via dissolved 0.1 mg/mL in LC–MS solvent: 60% MeCN in 0.1% FA with a linear gradient ranging from 10–60% (v/v) MeCN in 0.1% (v/v) FA at a flow rate of 0.3 mL/

min over 75 min. The capillary temperature was set at 220 °C and the spray voltage at 4 kV

87

.

(11)

www.nature.com/scientificreports/

Biological investigations. The effect of the extract was evaluated against hepatocellular carcinoma (HepG2) cell line as well as the normal human peripheral lymphocytes (HPBL) in vitro. The study was approved and followed the Institutional Ethical Committee guidelines at Faculty of Science, Menoufia University, Egypt (MUFS-F-GE-4-20).

Anticancer activities. Maintenance of HepG2 cell line. Hepatocellular carcinoma (HepG2) cell line was acquired from VACSERA, Giza, Egypt. The concentration of cells per milliliter was determined using a hemocy- tometer and calculated using the following equation:

The cell line was maintained and cultured in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10% fetal calf serum, 100 U/mL penicillin, and 100 μg/mL streptomycin. Cells were incubated in a humidi- fied 5% CO

2

atmosphere at 37 °C in T25 culture flasks at a density of 2 × 10

4

cells/cm

2

. The medium in flasks was changed every 48 h. The confluency of cells was confirmed by an inverted microscope until reaching 75%.

Cells were harvested after trypsinization (0.025% trypsin and 0.02% EDTA) then washed twice with phosphate- buffered saline (PBS). All experiments were done in triplicates. All reagents and media were purchased from Lonza supplier, Egypt.

Cytotoxicity on HepG2 cell line. To evaluate the maximal half inhibitory concentration (IC

50

), cytotoxic- ity of H. HA1 bacterial extract was carried out by microculture tetrazolium (MTT) assay method, 3-(4, 5-dimeth- ylthiazolyl-2)-2, 5-diphenyltetrazolium bromide, against HepG2 cell line. Cells were seeded into 96-well plates at a plating density of 1 × 10

4

cells/well and incubated to allow the attachment of cells prior to the addition of treat- ments. The bacterial extract was dissolved in dimethyl sulfoxide (DMSO) and diluted in serum-free medium.

After reaching the confluency of 75%, various concentrations (0–1,000 μg/mL) of the extract were added to the cultures then incubated for 24 h. Cisplatin (3 μg/mL) was used as a positive control

46

and DMSO (8 μL/ mL) as negative solvent control. After 24 h, 100 μL of MTT in PBS were added to each well and incubated at 37 °C for 4 h. The formazan crystals were formed then solubilized in 100μL of acidified isopropanol and then measured the absorbance at 630 nm by using enzyme-linked immunosorbent assay (ELISA) microplate reader (Bio-RAD microplate reader, Japan). The percentages of cell inhibition were determined using the following formula:

This assay was carried out in triplicate. The IC

50

value was determined from % cell inhibition and concentra- tion curve

88

. For further anticancer mechanistic investigations, HepG2 cells were incubated with 20 and 40 μg/

mL of bacterial extract for 24 h.

Morphological changes in HepG2. Phase contrast inverted microscopy. HepG2 cells were examined under the inverted Olympus BX41 microscope (Tokyo, Japan) at × 400 magnification to observe morphological alterations among treated groups.

Acridine orange/ethidium bromide (AO/EB) double fluorescent labeling. In order to investigate the altered cell structure towards apoptosis or necrosis, AO/EB staining was carried out

89

. Briefly, 4 μL of treated and control cell suspensions were transferred to glass slides then they were stained with 1μL AO/ EB staining solution (100 μg/mL AO and 100 μg/mL EB). Cells were examined immediately by a fluorescent microscope (Olympus BX 41, Japan) at 400 × magnification. One hundred cells were counted per each field of randomly selected five fields.

flow cytometric analysis of cell cycle distribution. Cell cycle distribution and DNA content of treated cells and controls were evaluated by flow cytometry after propidium iodide (PI) labeling. After various treatments period, cells were trypsinized and washed with PBS twice and then fixed with cold ethanol (70%) for 2 h. The alcohol was entirely removed and cells were collected and washed with PBS at 1,200 rpm for 5 min.

Cells were kept in PBS containing 50 μg/mL of RNase A for 2 h at 37 °C, mixed with 25 μg/mL of PI stain

90,91

and analyzed on a FACS flow cytometer (Becton Dickinson, US) according to the manufacturer’s instructions.

The analyses of cell cycle phases were performed using BD FACS Diva software. All reagents used were Sigma- Aldrich, Germany.

immunocytochemical staining. The immunocytochemical reactions were performed using the avi- din–biotin complex immunoperoxidase technique

92

. Control and treated cells were harvested and smeared on positive-charged slides

13

. Slides were then processed for P53 (tumor suppressor protein), BCL-2 (anti-apop- totic protein), BAX (BCL-2 associated X protein, apoptosis regulator) and caspase-3 (pro-apoptotic protein) immune-reaction. Moreover, PCNA (proliferating-cell nuclear antigen) were also used as proliferation markers.

Across five randomly selected fields, two hundred HepG2 cells were studied. As the immunocytochemical reac- tivity, the mean percentage of positive cells was calculated (positive cells/total number of counted cells) × 100.

Cells were counted under a light microscope (Olympus BX41, Japan) at × 400 magnification. All antibodies used were Invitrogen, CA, US.

Cells/mL = 10

4

× Average count per square × (Dilution factor)

% Cell inhibition = (1 − OD (absorbance) test/OD Control) × 100

(12)

www.nature.com/scientificreports/

Quantitative real-time polymerase chain reaction (qPCR). The qPCR was carried out in treated and control cells to assess BCL-2, BAX, P53, and CASP-3 mRNA expressions. Briefly, total RNA was extracted using RNeasy Plus Minikit (Qiagen, Hilden, Germany) following the manufacturer’s protocol. The quality of RNA was assessed by agarose gel and quantified by nanodrop (Quawell Q5000 UV–Vis Spectrophotometer, UK).

For cDNA synthesis, total RNA was reverse transcribed following the manufacturer’s instructions RevertAid H Minus Reverse Transcriptase (Thermo Fisher Scientific, US). Real-time PCR was performed using Power SYBR Master Mix (Thermo Fisher Scientific, US) on an Applied Biosystems 7500 system (Foster City, US) following the standard program of 40 cycles with 58 °C for primer annealing. Samples of cDNA were run in triplicate.

All data were then normalized to the endogenous control, GAPDH, a housekeeping gene. Fold change in gene expression was calculated using the comparative threshold cycle (

ΔΔ

C

T

) method. Primer sequences and acces- sion numbers of genes used in this study are provided in Table 2.

Cytotoxicity of the bacterial extract on normal human peripheral blood lymphocytes (HPBL). Cells’ culture and isolation. The extract toxicity was done on HPBL isolated from three male vol- unteers (healthy and non-smoker). Written informed consent was obtained from the three involved volunteers.

Peripheral venous blood samples were collected using sterile syringes and then transferred into sterile tubes (KEMICO vacutainer, Egypt) containing K

2

-EDTA. Samples were processed for culturing in RPMI-1640 me- dium supplemented with 15% fetal calf serum, phytohemagglutinin (2%) and 1% (100 U/mL penicillin and 100 μg/mL streptomycin) at 37 °C and humidified 5% CO

2

atmosphere. After 48 h of culture setup, various treatments were applied and the cultures extended for 24 h. For HPBL isolation after the treatment period, cultures were incubated with five folds of erythrocyte lysing buffer (0.015 M NH

4

C1, 1 mM NaHCO

3

, 0.l mM EDTA) for 15 min. Then, centrifugation was done for 5 min at 1,500 rpm in a cooling centrifuge (Sigma 3 K 30, Germany). The incubation was repeated until a white pellet of lymphocytes appeared

93

. All reagents were from Lonza, Switzerland.

Trypan blue exclusion assay. Using the 0.4 percent trypan blue (Sigma-Aldrich, Germany) solution, the HPBL cells were stained and examined immediately under a light microscope at × 200 magnification (Olympus BX41, Japan). Randomly five fields were selected and 200 cells/field were scored. Finally, cytotoxicity was calcu- lated as follows:

Acridine orange/ethidium bromide (AO/EB) dual fluorescent staining. The AO/EB staining was performed to investigate the viability of cells. Briefly, 4 μL of treated and control cells were stained with 1 μL stain solution AO/EB (100 μg/mL AO and 100 μg/mL EB) on glass slides and examined immediately by a fluorescent microscope (Olympus BX 41, Japan) at × 400 magnification. Randomly five fields were observed and 200 cells were counted from each. Two types of cells were observed, based on the emitted fluorescence: viable cells were green-colored cells with intact structures and late apoptotic or dead cells showed an orange-to-red color

89

. DNA single-strand breaks (comet assay). The  alkaline single-cell gel electrophoresis (comet assay) method

94

was carried out to evaluate the genotoxicity on normal HPBL. Briefly, cells were suspended in low melting point agarose gel (Sigma-Aldrich, Germany) on a microscopic glass slide between two layers of ultra- pure normal melting agarose (Sigma-Aldrich, Germany). Then, slides were immersed in lysis buffer (2.5 M NaCl, 100 mM EDTA and 10 mM Tris, pH 10.0) with freshly added 1% Triton X-100 (Sigma-Aldrich, Germany) and 10% DMSO for 1 h at 4 °C. Subsequently, slides were kept in the freshly prepared alkaline buffer (300 mM NaOH and 1 mM EDTA, pH > 13) for 20 min at 4 °C. Then, slides were electrophoresed in electric current of 25 V and 300 mA for 10 min. The slides were then immersed for 3 min in neutralizing buffer (0.4 M Tris–HCl, pH 7.5). They were stained with ethidium bromide (Sigma-Aldrich, Germany). Visualization of cells was carried out by a fluorescence microscope (Olympus BX 41, Japan), and representative images were taken. About 100 randomly selected cells were examined per one field of the total examined five fields. The results were divided as normal nuclei with no migrated tails and damaged with a migrated tail or with no distinct nucleus.

Cytotoxicity = 1 − (No. of unstained viable cells/No. of stained and unstained cells) × 100.

Table 2. Primer sequences of genes analyzed by real-time PCR. GAPDH Glyceraldehyde-3-phosphate dehydrogenase; BCL-2 B-cell lymphoma 2; BAX BCL-2-associated X protein; CASP-3 Caspase-3.

Name Accession number Sense (5′–3′) Antisense (5′–3′)

GAPDH NM_001289745.2 GGA TTT GGT CGT ATT GGG GGA AGA TGG TGA TGG GAT T BCL-2 NM_001114735.1 TAC AGG CTG GCT CAG GAC TAT CGC AAC ATT TTG TAG CAC TCTG BAX NM_001291431.1 CCC GAG AGG TCT TTT TCC GAG CCA GCC CAT GAT GGT TCT GAT CASP-3 NM_001354777.1 GGC GCT CTG GTT TTC GTT AAT CAG TTC TGT ACC ACG GCA GG P53 NM_001126112.2 TTC CCT GGA TTG GCC AGA CT GCA GGC CAA CTT GTT CAG TG

(13)

www.nature.com/scientificreports/

Statistical analysis. Data are reported as the mean ± standard deviation (SD). Data for multiple variable comparisons were analyzed by one-way analysis of variance. Duncan’s test was used as a post-hoc test using the statistical package SPSS version 17.0 software to compare the significance of differences between groups (SPSS, Inc., Chicago, IL, US) and significance was considered at a probability level of P < 0.05.

Data availability

All data of this study are introduced in this published article.

Received: 9 March 2020; Accepted: 28 July 2020

References

1. Zhu, R. X., Seto, W. K., Lai, C. L. & Yuen, M. F. Epidemiology of hepatocellular carcinoma in the Asia-Pacific region. Gut Liver 10, 332–339 (2016).

2. Akinyemiju, T. et al. The burden of primary liver cancer and underlying etiologies from 1990 to 2015 at the global, regional, and national level: results from the Global Burden of Disease Study 2015. JAMA Oncol. 3, 1683–1691 (2017).

3. Yang, J. D. & Roberts, L. R. Hepatocellular carcinoma: A global view. Nat. Rev. Gastroenterol. Hepatol. 7, 448–458 (2010).

4. Kreidieh, M., Zeidan, Y. H. & Shamseddine, A. The Combination of Stereotactic Body Radiation Therapy and Immunotherapy in Primary Liver Tumors. J. Oncol. 2019, (2019).

5. Florea, A.-M.M. & Büsselberg, D. Cisplatin as an anti-tumor drug: Cellular mechanisms of activity, drug resistance and induced side effects. Cancers (Basel). 3, 1351–1371 (2011).

6. Hanif, R. et al. Effects of nonsteroidal anti-inflammatory drugs on proliferation and on induction of apoptosis in colon cancer cells by a prostaglandin-independent pathway. Biochem. Pharmacol. 52, 237–245 (1996).

7. Kamesaki, H. Mechanisms involved in chemotherapy-induced apoptosis and their implications in cancer chemotherapy. Int. J.

Hematol. 68, 29–43 (1998).

8. Tohamy, A. A., El-Garawani, I. M., Ibrahim, S. R. & Abdel Moneim, A. E. The apoptotic properties of Salvia aegyptiaca and Trigo- nella foenum-graecum extracts on Ehrlich ascites carcinoma cells: The effectiveness of combined treatment. Res. J. Pharm. Biol.

Chem. Sci. 29:710–722 (2016).

9. Prada-Gracia, D., Huerta-Yépez, S. & Moreno-Vargas, L. M. Application of computational methods for anticancer drug discovery, design, and optimization. Bol. Méd. Hosp. Infant. México 73, 411–423 (2016).

10. Zhai, S. Q., Li, J. & Shi, Q. C. Composition characterization, antioxidant capacities and anti-proliferative effects of the polysac- charides isolated from Trametes lactinea (Berk.) Pat. Int. J. Biol. Macromol. 115, 114–123 (2018).

11. El-Garawani, I. et al. In vitro antigenotoxic, antihelminthic and antioxidant potentials based on the extracted metabolites from lichen, candelariella vitellina. Pharmaceutics https ://doi.org/10.3390/pharm aceut ics12 05047 7 (2020).

12. El-Garawani, I., El Nabi, S. H., Nafie, E. & Almeldin, S. Foeniculum Vulgare and Pelargonium Graveolens essential oil mixture triggers the cell cycle arrest and apoptosis in MCF-7 cells. Anticancer. Agents Med. Chem. https ://doi.org/10.2174/15733 99815

66619 03261 15116 (2019).

13. El-Garawani, I. M., El-Nabi, S. H., Dawoud, G. T., Esmail, S. M. & Abdel Moneim, A. E. Triggering of apoptosis and cell cycle arrest by fennel and clove oils in Caco-2 cells: the role of combination. Toxicol. Mech. Methods 29, 710–722 (2019).

14. Elkhateeb, W. A. et al. Ganoderma applanatum secondary metabolites induced apoptosis through different pathways: In vivo and in vitro anticancer studies. Biomed. Pharmacother. 101, 264–277 (2018).

15. El-Nabi, S. H., Dawoud, G., El-Garawani, I. & El-Shafey, S. HPLC analysis of phenolic acids, antioxidant activity and in vitro effectiveness of green and roasted Caffea arabica bean extracts: a comparative study. Anti-Cancer Agents Med. Chem. 18, 1281–1288 (2018).

16. Ahmed, A., Ali, M. & Sherif, N. Anticancer activity of Morus nigra on human breast cancer cell line ( MCF-7): the role of fresh and dry fruit extracts. J. Biosci. Appl. Res 2, 352–361 (2016).

17. El-Garawani, I. M. et al. Candelariella vitellina extract triggers in vitro and in vivo cell death through induction of apoptosis: A novel anticancer agent. Food Chem. Toxicol. 127, 110–119 (2019).

18. Demain, A. L. & Sanchez, S. Microbial drug discovery: 80 Years of progress. J. Antibiot. (Tokyo) 62, 5–16 (2009).

19. Charu, G., Dhan, P., Garg, A. P. & Sneh, G. Nutraceuticals from microbes. In Phytochemicals of Nutraceutical Importance 79–102 (CABI, Wallingford, 2014).

20. Berdy, J. Bioactive microbial metabolites. J. Antibiot. (Tokyo) 58, 1–26 (2005).

21. Dang, V. T., Benkendorff, K., Green, T. & Speck, P. Marine snails and slugs: a great place to look for antiviral drugs. J. Virol. 89, 8114–8118 (2015).

22. Desbois, A. P., Mearns-Spragg, A. & Smith, V. J. A fatty acid from the diatom Phaeodactylum tricornutum is antibacterial against diverse bacteria including multi-resistant Staphylococcus aureus (MRSA). Mar. Biotechnol. 11, 45–52 (2009).

23. Plaza, A. et al. Celebesides A-C and theopapuamides B-D, depsipeptides from an Indonesian sponge that inhibit HIV-1 entry. J.

Org. Chem. 74, 504–512 (2009).

24. Wei, X., Nieves, K. & Rodríguez, A. D. Neopetrosiamine A, biologically active bis-piperidine alkaloid from the Caribbean sea sponge Neopetrosia proxima. Bioorgan. Med. Chem. Lett. 20, 5905–5908 (2010).

25. Nuijen, B. et al. Pharmaceutical development of anticancer agents derived from marine sources. Anticancer. Drugs 11, 793–811 (2000).

26. Asolkar, R. N. et al. Arenamides A-C, cytotoxic NFkappaB inhibitors from the marine actinomycete Salinispora arenicola. J. Nat.

Prod. 72, 396–402 (2009).

27. Eltem, R. & Ucar, F. The determination of antimicrobial activity spectrum of 23 Bacillus strains isolated from Denizli-Acigol (Bitter Lake) which is soda lake (Na2SO4). J. KUKEM 21, 57–64 (1998).

28. Giddings, L.-A. & Newman, D. J. Bioactive compounds from marine extremophiles. In Bioactive Compounds from Marine Extre-

mophiles 1–124 (Springer, Cham, 2015).

29. Bitzer, J. et al. New aminophenoxazinones from a marine Halomonas sp.: fermentation, structure elucidation, and biological activ- ity. J. Antibiot. (Tokyo) 59, 86–92 (2006).

30. Sagar, S. et al. Induction of apoptosis in cancer cell lines by the Red Sea brine pool bacterial extracts. BMC Complement. Altern.

Med. 13, 344 (2013).

31. Mi, J. N., Wang, J. R. & Jiang, Z. H. Quantitative profiling of sphingolipids in wild Cordyceps and its mycelia by using UHPLC-MS.

Sci. Rep. 6, 1–11 (2016).

32. Marrouchi, R. et al. Toxic C17-sphinganine analogue mycotoxin, contaminating tunisian mussels, causes flaccid paralysis in rodents. Mar. Drugs 11, 4724–4740 (2013).

33. Kharrat, R. et al. The marine phycotoxin gymnodimine targets muscular and neuronal nicotinic acetylcholine receptor subtypes

with high affinity. J. Neurochem. 107, 952–963 (2008).

(14)

www.nature.com/scientificreports/

34. El-Gendy, M. M. A., Shaaban, M., El-Bondkly, A. M. & Shaaban, K. A. Bioactive benzopyrone derivatives from new recombinant fusant of marine streptomyces. Appl. Biochem. Biotechnol. 150, 85–96 (2008).

35. Cooper, E. L. & Albert, R. Tunicates: A vertebrate ancestral source of antitumor compounds. In Handbook of Anticancer Drugs

from Marine Origin 383–395 (Springer, Cham, 2015).

36. Lee, J.-C. et al. Marine algal natural products with anti-oxidative, anti-inflammatory, and anti-cancer properties. Cancer Cell Int.

13, 55 (2013).

37. Sithranga Boopathy, N. & Kathiresan, K. Anticancer drugs from marine flora: an overview. J. Oncol.

https ://doi.

org/10.1155/2010/21418 6 (2010).

38. Ganguly, R. K., Midya, S. & Chakraborty, S. K. Antioxidant and anticancer roles of a novel strain of Bacillus anthracis isolated from vermicompost prepared from paper mill sludge. Biomed Res. Int. https ://doi.org/10.1155/2018/10736 87 (2018).

39. Burlacu, A. Regulation of apoptosis by Bcl-2 family proteins. J. Cell. Mol. Med. 7, 249–257 (2003).

40. Zhu, L., Wang, P., Qin, Q. L., Zhang, H. & Wu, Y. J. Protective effect of polypeptides from larva of housefly (Musca domestica) on hydrogen peroxide-induced oxidative damage in HepG2 cells. Food Chem. Toxicol. 60, 385–390 (2013).

41. Kumar, S., Sharma, V. K., Yadav, S. & Dey, S. Antiproliferative and apoptotic effects of black turtle bean extracts on human breast cancer cell line through extrinsic and intrinsic pathway. Chem. Cent. J. 11, 1–10 (2017).

42. Tan, J., Wang, B. & Zhu, L. Regulation of survivin and Bcl-2 in HepG2 cell apoptosis induced by quercetin. Chem. Biodivers. 6, 1101–1110 (2009).

43. Abdel Moneim, A. E. Indigofera oblongifolia prevents lead acetate-induced hepatotoxicity, oxidative stress, fibrosis and apoptosis in rats. PLoS ONE 11, e158965 (2016).

44. Marchenko, N. D., Zaika, A. & Moll, U. M. Death signal-induced localization of p53 protein to mitochondria a potential role in apoptotic signaling. J. Biol. Chem. 275, 16202–16212 (2000).

45. Mihara, M. et al. p53 has a direct apoptogenic role at the mitochondria. Mol. Cell 11, 577–590 (2003).

46. Qu, L. et al. Endoplasmic reticulum stress induces p53 cytoplasmic localization and prevents p53-dependent apoptosis by a pathway involving glycogen synthase kinase-3β. Genes Dev. 18, 261–277 (2004).

47. Ruiz-Ruiz, C. et al. An exopolysaccharide produced by the novel halophilic bacterium Halomonas stenophila strain B100 selectively induces apoptosis in human T leukaemia cells. Appl. Microbiol. Biotechnol. 89, 345–355 (2011).

48. Liu, T., Brouha, B. & Grossman, D. Rapid induction of mitochondrial events and caspase-independent apoptosis in Survivin- targeted melanoma cells. Oncogene 23, 39–48 (2004).

49. Bröker, L. E., Kruyt, F. A. E. & Giaccone, G. Cell death independent of caspases: a review. Clin. Cancer Res. 11, 3155–3162 (2005).

50. Maier, T., Güell, M. & Serrano, L. Correlation of mRNA and protein in complex biological samples. FEBS Lett. https ://doi.

org/10.1016/j.febsl et.2009.10.036 (2009).

51. Wu, G., Nie, L. & Zhang, W. Integrative analyses of posttranscriptional regulation in the yeast Saccharomyces cerevisiae using transcriptomic and proteomic data. Curr. Microbiol. https ://doi.org/10.1007/s0028 4-008-9145-5 (2008).

52. Elmore, S. Apoptosis: a review of programmed cell death. Toxicol. Pathol. https ://doi.org/10.1080/01926 23070 13203 37 (2007).

53. Shalini, S., Dorstyn, L., Dawar, S. & Kumar, S. Old, new and emerging functions of caspases. Cell Death Differ. https ://doi.

org/10.1038/cdd.2014.216 (2015).

54. Galluzzi, L. et al. Molecular mechanisms of cell death: Recommendations of the Nomenclature Committee on Cell Death 2018.

Cell Death Differ. https ://doi.org/10.1038/s4141 8-017-0012-4 (2018).

55. Jänicke, R. U., Sprengart, M. L., Wati, M. R. & Porter, A. G. Caspase-3 is required for DNA fragmentation and morphological changes associated with apoptosis. J. Biol. Chem. https ://doi.org/10.1074/jbc.273.16.9357 (1998).

56. Blanc, C. et al. Caspase-3 is essential for procaspase-9 processing and cisplatin-induced apoptosis of MCF-7 breast cancer cells.

Cancer Res. 60, 4386–4390 (2000).

57. Xu, B. et al. Proliferating cell nuclear antigen (PCNA) regulates primordial follicle assembly by promoting apoptosis of oocytes in fetal and neonatal mouse ovaries. PLoS ONE 6, e16046 (2011).

58. DiPaola, R. S. To arrest or not to G(2)-M Cell-cycle arrest. Clin. Cancer Res. 8, 3311–3314 (2002).

59. Bennukul, K., Numkliang, S. & Leardkamolkarn, V. Melatonin attenuates cisplatin-induced HepG2 cell death via the regulation of mTOR and ERCC1 expressions. World J. Hepatol. 6, 230 (2014).

60. Brandwagt, B. F. et al. A longevity assurance gene homolog of tomato mediates resistance to Altemaria alternata f sp lycopersici toxins and fumonisin B1. Proc. Natl. Acad. Sci. USA 97, 4961–4966 (2000).

61. Van Meer, G. & De Kroon, A. I. P. M. Lipid map of the mammalian cell. J. Cell Sci. 124, 5–8 (2011).

62. Olsen, I. & Jantzen, E. Sphingolipids in bacteria and fungi. Anaerobe 7, 103–112 (2001).

63. Batrakov, S. G., Mosezhnyi, A. E., Ruzhitsky, A. O., Sheichenko, V. I. & Nikitin, D. I. The polar-lipid composition of the sphingolipid- producing bacterium Flectobacillus major. Biochim. Biophys. Acta 1484, 225–240 (2000).

64. An, D. et al. Sphingolipids from a symbiotic microbe regulate homeostasis of host intestinal natural killer T cells. Cell 156, 123–133 (2014).

65. Hong, S. et al. Inhibitory effect against Akt by cyclic dipeptides isolated from Bacillus sp. J. Microbiol. Biotechnol. 18, 682–685 (2008).

66. Vázquez-Rivera, D. et al. Cytotoxicity of cyclodipeptides from pseudomonas aeruginosa PAO1 leads to apoptosis in human cancer cell lines. Biomed Res. Int. 20, 15. https ://doi.org/10.1155/2015/19760 8 (2015).

67. Karanam, G., Arumugam, M. K. & Sirpu Natesh, N. Anticancer effect of marine sponge-associated Bacillus pumilus AMK1 derived dipeptide Cyclo (-Pro-Tyr) in human liver cancer cell line through apoptosis and G2/M phase arrest. Int. J. Pept. Res. Ther. 26, 445–457 (2020).

68. Save, S. A., Lokhande, R. S. & Chowdhary, A. S. Determination of 1, 2-benzenedicarboxylic acid, bis (2-ethylhexyl) ester from the twigs of Thevetia peruviana as a colwell biomarker. JIPBS 2, 349–362 (2015).

69. Wu, Y. S. et al. Anticancer activities of surfactin potential application of nanotechnology assisted surfactin delivery. Front. Phar-

macol. 8, 761 (2017).

70. Abdelli, F. et al. Antibacterial, anti-adherent and cytotoxic activities of surfactin(s) from a lipolytic strain Bacillus safensis F4.

Biodegradation 30, 287–300 (2019).

71. Alvionita, M. & Hertadi, R. Bioconversion of glycerol to biosurfactant by halophilic bacteria Halomonas elongata BK-AG18.

Indones. J. Chem. 19, 48–57 (2019).

72. Kim, S. D. et al. A comparison of the anti-inflammatory activity of surfactin A, B, C, and D from Bacillus subtilis. J. Microbiol.

Biotechnol. 16, 1656–1659 (2006).

73. Meena, K. R., Sharma, A. & Kanwar, S. S. Antitumoral and antimicrobial activity of surfactin extracted from Bacillus subtilis KLP2015. Int. J. Pept. Res. Ther. 26, 423–433 (2020).

74. Vollenbroich, D., Pauli, G., Özel, M. & Vater, J. Antimycoplasma properties and application in cell culture of surfactin, a lipopeptide antibiotic from Bacillus subtilis. Appl. Environ. Microbiol. 63, 44–49 (1997).

75. Kracht, M. et al. Antiviral and hemolytic activities of surfactin isoforms and their methyl ester derivatives. J. Antibiot. 52, 613–619 (1999).

76. Sen, R. Surfactin: Biosynthesis, Genetics and Potential Applications 316–323 (Springer, New York, 2010).

77. Jacques, P. Surfactin and Other Lipopeptides from Bacillus spp 57–91 (Springer, Berlin, 2011).

(15)

www.nature.com/scientificreports/

78. Heerklotz, H. & Seelig, J. Detergent-like action of the antibiotic peptide surfactin on lipid membranes. Biophys. J. 81, 1547–1554 (2001).

79. Lee, J. H. et al. The production of surfactin during the fermentation of cheonggukjang by potential probiotic Bacillus subtilis CSY191 and the resultant growth suppression of MCF-7 human breast cancer cells. Food Chem. 131, 1347–1354 (2012).

80. Liu, X. et al. Effect of themicrobial lipopeptide on tumor cell lines: apoptosis induced by disturbing the fatty acid composition of cell membrane. Protein Cell 1, 584–594 (2010).

81. Kameda, Y. & Kanatomo, S. Abstracts papers, The 88th Annual Meeting of Pharmaceutical Society of Japan‏. In The 88th Annual

Meeting of Pharmaceutical Society of Japan (p. 431) (1968).

82. Kim, S. Y. et al. Surfactin from Bacillus subtilis displays anti-proliferative effect via apoptosis induction, cell cycle arrest and survival signaling suppression. FEBS Lett. 581, 865–871 (2007).

83. Dorn, E., Hellwig, M., Reineke, W. & Knackmuss, H. J. Isolation and characterization of a 3-chlorobenzoate degrading pseu- domonad. Arch. Microbiol. 99, 61–70 (1974).

84. Chang, Y.-J. et al. Phylogenetic analysis of aerobic freshwater and marine enrichment cultures efficient in hydrocarbon degrada- tion: effect of profiling method. J. Microbiol. Methods 40, 19–31 (2000).

85. Thomas, A. T. et al. In vitro anticancer activity of microbial isolates from diverse habitats. Braz. J. Pharm. Sci. 47, 279–287 (2011).

86. Miao, L. et al. The anti-inflammatory potential of Portulaca oleracea L. (purslane) extract by partial suppression on NF-κB and MAPK activation. Food Chem. 290, 239–245 (2019).

87. El-Seedi, H. R. et al. The traditional medical uses and cytotoxic activities of sixty-one Egyptian plants: discovery of an active cardiac glycoside from Urginea maritima. J. Ethnopharmacol. 145, 746–757 (2013).

88. Thakur, A. N. et al. Antiangiogenic, antimicrobial, and cytotoxic potential of sponge-associated bacteria. Mar. Biotechnol. 7, 245–252 (2005).

89. Liu, K., Liu, P. C., Li, R. & Wu, X. Dual AO/EB staining to detect apoptosis in osteosarcoma cells compared with flow cytometry.

Med. Sci. Monit. Basic Res. 21, 15–20 (2015).

90. Dassonneville, L. et al. Cytotoxicity and cell cycle effects of the plant alkaloids cryptolepine and neocryptolepine: relation to drug- induced apoptosis. Eur. J. Pharmacol. 409, 9–18 (2000).

91. Dengler, R. et al. Immunocytochemical and flow cytometric detection of proteinase 3 (myeloblastin) in normal and leukaemic myeloid cells. Br. J. Haematol. 89, 250–257 (1995).

92. Hsu, S.-M., Raine, L. & Fanger, H. X. Use of avidin-biotin-peroxidase complex (ABC) in immunoperoxidase techniques: a com- parison between ABC and unlabeled antibody (PAP) procedures. J. Histochem. Cytochem. 29, 577–580 (1981).

93. El-Garawani, I. M. Ameliorative effect of Cymbopogon citratus extract on cisplatin-induced genotoxicity in human leukocytes. J.

Biosci. Appl. Res. 1, 304–310 (2015).

94. Singh, N. P., McCoy, M. T., Tice, R. R. & Schneider, E. L. A simple technique for quantitation of low levels of DNA damage in individual cells. Exp. Cell Res. 175, 184–191 (1988).

Acknowledgements

We are very grateful to the Swedish Research links grant VR (Stockholm, Sweden), Grant number 2016-05908 for generous financial support. Dr. S.A.M. Khalifa thanks the Department of Molecular Biosciences, Wenner- Grens Institute, Stockholm University, Sweden.

Author contributions

H.S., S.S. and I.EL-G. planned and designed the experiments; I.El-G. and N.A. performed analyzed and wrote the biological and molecular experiments; H.S. and D.M. performed, analyzed and wrote the phytochemical experiments; H.A. performed the microbiological experiments; H.A. and I.El-G. wrote the manuscript; I.El-G., S.K. and O.E. revised the manuscript. All authors read and approved the final manuscript.

competing interests

The authors declare no competing interests.

Additional information

Supplementary information is available for this paper at https ://doi.org/10.1038/s4159 8-020-70945 -8.

Correspondence and requests for materials should be addressed to I.M.E.-G. or H.R.E.-S.

Reprints and permissions information is available at www.nature.com/reprints.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creat iveco mmons .org/licen ses/by/4.0/.

© The Author(s) 2020

References

Related documents

In order to study whether the WNT5A-induced migration of OSCC cells, observed in paper I, was due to EMT and alterations in cell-adhesion proteins, we examined the expression of

Patients with metastatic renal cell carcinoma (MRCC) were eligible for this study, they had either a single lesion with a diameter of ≥20mm or multiple lesions with a diameter

Paper I Records of 87 patients subjected to NSS performed between 1980 and 1999 were reviewed, survival rate was determined with reference to grade stage and multiplicity of renal

Furthermore, IL-6 and IL-8 are well-known to affect the cancer stem cell propagation [76, 147, 179] and induced secretion of these cytokines could partially be responsible for

To study how amplification of CCND1 and PAK1 affect the postmenopausal breast cancer patient’s prognosis and response to adjuvant treatment and to analyse the importance of

Andra studier har visat att när cyklin D1 är överuttryckt kan det hjälpa till att aktivera östrogenreceptorn även när östrogen inte är bundet till receptorn vilket kan leda

• Cyclin E overexpression in MDA-MB-468 breast cancer cells induces an elevated S-phase fraction and hampers cell motility and invasiveness. • High levels of cyclin E protein

Expression of p63, Cox-2, EGFR, β-catenin, PP2A and p53 isoforms was mapped in tumours and normal tumour adjacent tissue from patients with SCCHN using western blot or