2,565
Views
13
CrossRef citations to date
0
Altmetric
Article

Papain bioinspired gold nanoparticles augmented the anticancer potency of 5-FU against lung cancer

, , , , , , , & show all
Pages 109-128 | Received 15 Jan 2020, Accepted 18 Mar 2020, Published online: 07 May 2020

Abstract

Lung cancer is one of the most widely recognised types of cancer and the acquisition of resistance towards chemotherapeutic drugs worsens the situation. Therefore, a site-directed, multi-targeting drug delivery system that is compliant with patients is the need of the hour. 5-Fluorouracil (5-FU), an anti-cancer chemotherapy drug, was delivered to the lung cancer cells using papain-inspired gold nanoparticles (PpGNPs) as the drug delivery system. Papain is anti-cancer by nature; hence, it rendered this anti-cancer property to the PpGNPs as well. This bio-conjugation of 5-FU and PpGNPs worked synergistically and more efficiently in combating lung cancer. The synthesis, stability, and the size of the PpGNPs, and their bio-conjugation with 5-FU (5F-PpGNPs), were confirmed by different physical techniques: for example, UV-Vis spectroscopy, TEM, DLS, and estimation of the zeta potential. The drug-loading efficiency of 5-FU in 5F-PpGNPs was confirmed and validated by UV-Vis spectroscopy. The efficacy of 5F-PpGNPs (IC50 11.7 µg/mL) against human lung cancer A549 cell line was found to have improved significantly over that of pure 5-FU (IC50 25.6 µg/mL). However, the 5F-PpGNPs did not show any significant toxicity (even up to a fairly high concentration) towards normal mouse embryonic fibroblasts cell line (3T3-L1). The apoptotic effects, nuclear condensation, ROS generation, and the loss of mitochondrial membrane potential (ΔΨm) of 5F-PpGNPs were analysed. The results clearly showed that conjugation with papain-inspired gold nanoparticles (5F-PpGNPs) significantly augmented the potency of 5-FU by acting synergistically. Thus, the enhanced anti-proliferating effect of 5F-PpGNPs over that of the pure drug would be an important step that will help to overcome the resistance of chemotherapeutic drugs towards lung cancer cells and prove to be of great advantage during lung cancer treatment.

1. Introduction

Lung cancer is the most inexhaustible type of cancer with a 5-year survival rate of only 18% [Citation1]. Unfortunately, in most cases, the diagnosis of lung cancer takes place only in the advanced stages [Citation2]. The platinum-based regimen has been used as the standard first-line of chemotherapy against advanced lung cancer. However, this causes severe side effects and has a success rate of not more than 30%. Several chemotherapeutic agents such as camptothecin, taxanes, platinating agents, nucleoside and nucleotide analogues are in vogue against certain cancers for several years [Citation3]. However, these chemotherapeutic agents are not specific in their target, cause non-specific interactions with normal cells, and pose several severe side effects such as cardiotoxicity, cytotoxicity, neurotoxicity, nephrotoxicity and ototoxicity [Citation4]. 5-Fluorouracil (5-FU) is one of the most commonly used chemotherapeutic drugs because of its high efficacy. 5-FU inhibits thymidylate synthase by acting as an antimetabolite. After administration, 5-FU gets converted into a cytotoxic metabolite like fluorodeoxyuridine monophosphate (FdUMP) that structures a covalent ternary complex with thymidylate synthase; it hinders the enzyme activity by preventing its binding with the natural substrate [Citation5]. This leads to an increase in the concentration of dUMP and a decrease in the production of thymidine (dTMP), which is required for DNA synthesis during cell multiplication [Citation5–8].

There is a need for the development of highly effective therapeutic agents with minimal side-effects and maximum patient-compliance. The existing therapeutic agents can be improved and delivered to the target site ‘unaltered’, along with other natural agents, so that both can act in synergy. Nature has been a pool of medicinal agents since the beginning of life. Different therapeutic plants have been utilised in our daily life to treat illnesses globally for a considerable length of time [Citation9]. Papaya latex constitutes several proteolytic enzymes such as papain, chymopapain, glutamine cyclotransferase, chymopapain A, B, and C, peptidase A and B and lysozymes. Papain, the major constituent of papaya, is a plant cysteine-protease with cysteine, histidine and aspartic acid in its active site [Citation10]. It recognizes and cleaves the peptide bonds of basic amino acids, particularly arginine, lysine and the residues following phenylalanine [Citation11]. The cells of many cancers have a fibrin coating that protects them and helps in remaining undetected for a longer period. Papain has the ability to break this fibrin coat belonging to the cancer cells; therefore, it can be used as a potential anti-cancer agent. The α-1-antitrypsin and α-2-macroglobulin are antiproteinases that strongly and irreversibly interact with papain and induce the synthesis of other antiproteinases. This enhanced synthesis of other antiproteinases causes the deactivation of additional proteinases like cathepsins that contribute to the risk of metastasis [Citation12]. The cysteine proteinases from plant sources (e.g. bromelain, papain) can manipulate the comparative concentration of proteinases and anti-proteinases and consequently impact tumour metastasis [Citation13]. Additionally, papain is thought to communicate with a network of cytokines. It can readily bind with α-2-macroglobulins and form α-2-macroglobulin-proteinase complexes having higher affinities towards cytokines (e.g. transforming growth factor-β (TGF-β), interferon-γ, interleukin-1 (IL-1) and interleukin-6 (IL-6), and ultimately clear these cytokines from the system. TGF-β is known to promote immune suppression in the host cells and also promote tumour growth therein. Papain also interacts with adhesion molecules such as CD-44, CD-49, CD-54 and CD-58, which are the key components in tumour development and metastasis, and down regulate their activity [Citation14].

Several drug delivery systems are being extensively utilised in the biomedical industry. Among these, the inorganic nanoparticle-based delivery systems are gaining maximum success due to their novel physicochemical properties that are a consequence of their nanoscale dimensions [Citation15]. Currently, these nanomaterials can be synthesised and functionalised using various functional groups, thus enabling their bioconjugation with selective drugs, ligands, and antibodies [Citation16, Citation17]. This development has paved the way for a vast range of biological applications, targeted drug delivery, magnetic separation, gene delivery vehicles and diagnostic imaging. Biogenic inorganic nanoparticles claim to produce maximum success and exhibit greater patient-compliance [Citation18–20]. Metallic nanoparticles synthesised using enzymes have been used successfully in drug delivery systems to deliver cisplatin, doxorubicin, methotrexate, and several other drugs directly to the nucleus at their target site [Citation15, Citation21–23]. The disulfide bridges and thiol (–SH) moieties present in proteins or peptides are thought to be the catalytic sites for reduction [Citation24]. Interestingly, 5-FU, when used in isolation, is not considered a good therapeutic agent against lung cancer due to its low bioavailability and severe side effects. However, in the present study, 5-FU was bioconjugated with gold nanoparticles synthesised using papain and was delivered ‘unaltered’ into the lung cancer cells where it worked with papain synergistically and at much lower concentrations.

2. Materials and methods

2.1. Materials

Chemicals and reagents tetrachloroauric [III] acid (HAuCl4), 5-fluorouracil and papain were bought from Sigma-Aldrich. Unless stated otherwise, all chemicals and solvents were of analytical grade and were utilised as received.

2.2. In vitro synthesis of PpGNPs

Papain-capped GNPs were synthesised by the biological method using papain as a reducing as well as a capping agent. The synthesis of PpGNPs was performed at 6 °C for 48 H in a 3-mL reaction mixture containing 1 mM tetrachloroauric [III] acid (H[AuCl4]) and 2 mg/mL papain in 100 mM HEPES buffer at pH 6.0. The same reaction was also performed at different temperatures (such as 16 °C, 25 °C, 37 °C and 40 °C) and in buffer solutions of different pH (such as pH 5.0, 6.0, 7.0 and 8.0), but favourable results were obtained only at conditions described above. The completion of the reaction was confirmed by UV–vis spectroscopy when the colour changed from colourless to pink. After completion, the mixture was treated with 50% ethanol to remove the unbound papain and then utilised for further purposes.

2.3. Bioconjugation of PpGNPs with 5-FU

In vitro, the above-synthesised gold nanoparticles (PpGNPs) were bioconjugated with the anti-cancer drug, 5-fluorouracil (5-FU). The coupling agent 1-ethyl-3-(3-dimethyl) carbodiimide (EDC) was used to establish a covalent bond between the free amino groups (secondary amine) of 5-FU and the carboxylate groups of papain in the PpGNPs [Citation25]. For this purpose, 250 µg of 5-FU was incubated with 250 µg of PpGNPs in a 50 mM HEPES buffer solution at a temperature of 30 °C. The solution of 5 mM EDC was added in aliquots in the given reaction mixture within 3H.

2.4. Characterisation of PpGNPs and 5F-PpGNPs

The UV–vis spectroscopic analysis of PpGNPs and 5F-PpGNPs was performed in the range of wavelengths 220–800 nm, in a quartz cuvette with a path length of 1 cm. The study was conducted on a double-beam spectrophotometer (Shimadzu, model UV-1601 PC). The transmission electron microscope (TEM) was used to analyse the size and morphology of PpGNPs and 5F-PpGNPs. The FEI TecnaiTM G2 Spirit BioTWIN from FEI Company was operated at an accelerating voltage of 80 kV. For the preparation of TEM samples, 1 µL droplets of the PpGNPs and 5F-PpGNPs were deposited and dried overnight onto a carbon-coated copper disc. The prepared samples were then kept in a vacuum chamber (∼4 × 10−3 Torr) for 12 h. The Gatan digital micrograph was used to manually estimate the size range of the PpGNPs and 5F-PpGNPs from the obtained TEM images. The dynamic light scattering (DLS) method was used to analyse the mean particle size (MPS) of PpGNPs and 5F-PpGNPs by employing a particle size analyzer (Zetasizer Nano-ZS, Model ZEN3600, Malvern Instrument Ltd, Malvern, UK). The sample powder was diluted in deionised water to a concentration of 0.5% (w/v) and sonicated for 1 min before analysis. The sample was taken in a DTS0112-low volume disposable sizing cuvette of 1.5 mL. MPS was taken as the average of measurements performed in triplicate for a single sample. Zeta potential was also measured using a Zetasizer Nano-ZS, Model ZEN3600 (Malvern Instrument Ltd, Malvern, UK). To validate the binding of 5-FU with PpGNPs, Fourier transform infrared spectroscopy (FTIR) was used. Films of PpGNPs and 5F-PpGNPs were encrusted by putting a single drop of each on a Si (111) substrate and moderate heating was used to evaporate the water over the coated surface. The Shimadzu FTIR-8201 PC instrument, operated at a resolution of 4 cm−1, in the diffuse reflectance mode, was used to record the FTIR spectra of the film. In order to achieve good signal-to-noise ratios, 256 different scans were obtained from the analysed films, in the range of 400–4000 cm−1.

2.5. Drug-loading efficiency

The percentage loading of 5-FU on PpGNPs was calculated using EquationEquation (1), where the values of A and B were substituted. The characteristic absorbance of 5-FU was recorded at a wavelength of 265 nm. (1) Percent loading of  5 FU on PpGNPs = [ A B * 100 A ] (1) where A is the absorbance value of total 5-FU (50 µg) added along with PpGNPs (prior to bioconjugation) and B is the absorbance value of 5-FU after bioconjugation in the supernatant of 5F-PpGNPs [Citation26].

The efficiency of drug loading was calculated by determination of the UV–vis absorbance spectra at a wavelength of 265 nm after calibration with the absorbance obtained for the standard curve of 5-FU. The standard graph was drawn by increasing the concentration of 5-FU from 1.56 to 25 µg. (2) Percent bioconjugation = Amount of drug bioconjugated Total drug added * 100 (2)

Further, diameters (d) of the particles in the range of 35–100 nm were calculated from the peak position [Citation27] according to EquationEquation (3): (3) d = ln λ spr λ ο L 1 L 2 A = π r 2 (3)

For theoretical value   of d > 25   nm   ( λ ο = 512 ;   L 1 = 6.34 ; L 2 = 0.0216 ;   λ spr = 332 ) , fit parameters were determined accurately with only a 3% error against experimental values. Therefore, the given equation can be precisely used to determine particle size in the range of 35–110 nm. d = ln λ spr λ ο L 1 L 2 d = ln 532 512 6.53 0.0216 d = 51.77   n

Considering the authenticity of the above equation, it was successfully used to determine the number density of the particles (N) from the absorbance of the hydrosol. (4) N = A 450 × 10 14 d 2 0.295 + 1.36 e xp d 96.8 78.2 2 (4) A 450 - is the absorbance at 450 nm wavelength and d is the particle diameter in nanometre A 450 = 0.829 ; d = 51.77   nm N = A 450 × 10 14 d 2 0.295 + 1.36 e xp d 96.8 78.2 2 N = 0.829 × 10 14 ( 51.77 ) 2 0.295 + 1.36 e xp 51.77 96.8 78.2 2 N = 12 × 10 9

2.6. In vitro anticancer studies of PpGNPs and 5F-PpGNPs

2.6.1. Cell culture

The human lung cancer cell line (A549) and the mouse embryonic fibroblasts cell line (3T3-L1) were procured from National Centre for Cell Sciences (NCCS), Pune, India. The DMEM and the DMEM-F12 media were supplemented with 10% foetal bovine serum and 1% antimicrobial cocktail comprising of 10 mg of streptomycin, 10,000 units of penicillin, and 25 µg amphotericin, which were utilised to separately culture the cells in the monolayer at 37 °C in a CO2-humidified incubator. Stocks were maintained in 75 cm2 tissue culture flasks.

2.6.2. Determination of cell viability by MTT assay

The cytotoxicity of PpGNPs (50, 100, 150, 200, 250 and 300 µM), 5F-PpGNPs (concentration 5–30 µM) and 5-FU (5, 10, 15, 20, 25 and 30 µM) were assessed by MTT assays on human A549 and mouse 3T3-L1 cell lines [Citation28]. The absorbance of the given assay was recorded at 490 nm and the percentage (%) of live cells was determined in comparison with that of the untreated control. The Origin 6.0 Professional software was used to calculate IC50 values of the given samples.

2.6.3. Morphological analysis by phase-contrast microscopy

Eventually, morphological variations were observed after treating the A549 cells with different concentrations of papain (100–700 µM), PpGNPs (50–300 µM), 5F-PpGNPs (5–30 µM) and 5-FU (5–30 µM) for 48 h [Citation28]. An inverted phase-contrast microscope (Labomed, U.S.A) was used to observe the changes in the morphology of the treated cells in comparison with that of the untreated cells.

2.6.4. Detection of nuclear condensation by DAPI staining

DAPI (4′,6-diamidino-2-phenylindole), a fluorescent nuclear dye was used to study the apoptotic effects at different concentrations of PpGNPs (50, 100, 150, 200, 250 and 300 µM), 5F-PpGNPs (5–30 µM) and 5-FU (5, 10, 15, 20, 25 and 30 µM) on human A549 cell line [Citation15]. An inverted fluorescence magnifying instrument (Nikon ECLIPSE Ti-S, Japan) was employed to capture these stained cells.

2.6.5. Measurement of intracellular ROS level

The detection and quantitative estimation of the ROS production in the A549 cell line treated with PpGNPs (50, 100, 150, 200, 250 and 300 µM), 5F-PpGNPs and 5-FU (5, 10, 15, 20, 25 and 30 µM) were performed by the DCFH-DA method [Citation15]. Images were captured by a fluorescence microscope (Nikon ECLIPSE Ti-S, Japan). The fluorescence of treated cells was observed using a multi-wall smaller-scale plate reader (Synergy H1 Hybrid Multi-Mode Microplate Reader, BioTek, U.S.A.) at an excitation wavelength of 485 nm and an emission wavelength of 528 nm. The final results were expressed as the change in fluorescence intensity in comparison to the control samples.

2.6.6. Assessment of mitochondrial membrane potential (ΔΨm) by MitoTracker red

The Mito Tracker Red CMX Ros dye was used to assess the mitochondrial membrane potential in the A549 cell line treated with PpGNPs (50, 100, 150, 200, 250 and 300 µM), 5F-PpGNPs (5–30 µM), and 5-FU (5, 10, 15, 20, 25 and 30 µM) [Citation29]. The pictures were captured under an inverted fluorescence microscope (Nikon ECLIPSE Ti-S, Japan).

2.7. Statistical analysis

The result was portrayed as the mean ± S.E.M. of three independent studies performed in triplicate. One-way ANOVA utilising Dunnett’s multiple comparison test, and two-tailed, paired Student’s t-test (*p < 0.01, **p < 0.001, ***p < 0.0001 represent a significant difference compared with control) were employed for statistical analysis.

2.8. Study of synergistic effects of PpGNPs and 5-FU

Chou and Talalay’s method was used to determine the synergism and antagonism of PpGNPs and 5-FU in the given study [Citation30] using the ‘combination index’ (CIA is considered for drugs whose responses are linear to their doses). In the given study, PpGNPs and 5-FU were combined with doses (d1 – concentration of PpGNPs and d2 - concentration of 5-FU) to get a 50% effect. The combined effect was calculated by: CIA = d 1 ED 50   1   + d 2 ED 50 2 = P P + Q ED 50 1 + Q P + Q ED 50 2 ED 50 c here ,   d 1 = P P + Q ED 50 c d 2 = Q P + Q ED 50 c where ED50c is the combined effect of 50% inhibition; Q and P are concentrations of PpGNPs and 5-FU required to get a 50% effect, respectively. For the combination to be synergistic in action, CIA < 1; for antagonistic action, CIA > 1; and at CIA = 1, the combination becomes additive.

3. Results and discussions

The unique surface chemistry of the nanomaterials is a consequence of their nanometre sizes and different shapes. This property makes them highly valuable in different applications including medicine [Citation31], material science [Citation32], microelectronics [Citation33], energy storage [Citation34] and biomedical devices [Citation35].

3.1. Results

The present investigation deals with papain-inspired (a cysteine protease involved in antitumor/anti-angiogenesis) bioengineering of gold nanoparticles (PpGNPs). explains the synthesis and functionalisation of gold nanoparticles and their subsequent role in the delivery of 5-FU into the A549 lung cancer cell line. Proteases, such as bromelain and trypsin [Citation21–22], are the ideal candidates for the synthesis and stabilisation of nanomaterials.

Figure 1. Schematic representation of papain mediated synthesis of gold nanoparticles and their ioconjugation with 5-fluorouracil (5-FU) so that 5-FU could be delivered through caveolae-mediated endocytosis to the nucleus of lung cancer cells.

Figure 1. Schematic representation of papain mediated synthesis of gold nanoparticles and their ioconjugation with 5-fluorouracil (5-FU) so that 5-FU could be delivered through caveolae-mediated endocytosis to the nucleus of lung cancer cells.

3.1.1. Synthesis and characterisation of PpGNPs

Papain could not produce stable and mono-dispersed nano-emulsions at the optimum temperature of its activity. Therefore, an ideal set of parameters was standardised by performing reactions at different temperatures, pH and concentrations of papain. Astonishingly, it was found that a mono-dispersed and a stable emulsion was produced only at a temperature of 6 °C when HEPES buffer of pH 6.0 was used at 2 mg/mL concentration of papain. Nucleation of nanoparticles took place in 48 h at 6 °C, leading to the synthesis of a mono-dispersed and a stable nano-emulsion. Papain acted as a reducing as well as a capping agent. The synthesis of PpGNPs was confirmed by recording the characteristic surface plasmon resonance (SPR) spectra of GNPs at a wavelength of 530 nm (). Being a 23.6 kDa protease, papain contains 345 amino acids and three disulfide bridges; thus, making it an ideal template to encapsulate seeding nanoparticles. Different types of interactions including thiol bridges, electrostatic interactions, hydrogen, and dative bonds, and Van-der Waals forces [Citation36] are formed due to the presence of –OH, –OOC, –NH2, –CH3, –(–CH2)n and –SH groups. With the DLS technique, the hydrodynamic diameter of PpGNPs was found to be 49.53 nm (); this was further confirmed by EquationEquation (1), where size was found to be 51.77 nm and the number of particles was calculated to be 12 × 109. Topographical studies with absolute size determination were performed under TEM by manually using the Gatan digital micrograph. The micrograph confirmed the average sizes of PpGNPs to be ∼16 nm () and they had a uniform distribution with spherical shapes. The nano-emulsion of PpGNPs was found to be highly stable and anionic in nature with a zeta potential of −10.5 mV (). The stability of the nano-emulsion is described by zeta potential as well as the Hamaker constant [Citation37] by calculating electrostatic interaction and Van der Waal forces, respectively. Smaller values of the Hamaker constant with lower values of zeta potential can produce stable nanoemulsion [Citation37].

Figure 2. Characterisation of PpGNPs and 5F-PpGNPs under (A) UV–vis spectra, (B) dynamic light scattering (inset: 5F-PpGNPs), (C) transmission electron microscopy (inset: 5F-PpGNPs), (D) zeta potential (inset: 5F-PpGNPs).

Figure 2. Characterisation of PpGNPs and 5F-PpGNPs under (A) UV–vis spectra, (B) dynamic light scattering (inset: 5F-PpGNPs), (C) transmission electron microscopy (inset: 5F-PpGNPs), (D) zeta potential (inset: 5F-PpGNPs).

3.1.2. Bioconjugation of anti-cancer flutamide drug with PpGNPs

Further, the bioengineered PpGNPs were functionalised with the anti-cancer drug 5-FU by using EDC (a coupling agent) to establish physical bonding between the carboxylate groups of papain (present in surplus over the surface of PpGNPs) and the secondary amino groups of 5-FU, respectively. The functionality was confirmed by recording the SPR absorption spectra of 5F-PpGNPs (at 535 nm) in comparison to PpGNPs (at 530 nm) and 5-FU (at 265 nm) (). The changes in the broadening of the peaks, decrease in intensity, and shifting of the spectra towards a higher wavelength (red-shift, from 530 nm to 535 nm) were observed in UV–vis spectroscopy (). The attachment of any ligand on the surface of nanoparticles causes a change in the absorption intensity, with a minor change in the full width of the half maximum [Citation38]. Further, the hydrodynamic diameter of particles, observed under DLS increased to 57.53 nm after bioconjugation (, inset). A blurred image of 5F-PpGNPs under the TEM (, inset) with an increase in average sizes further confirmed the bioconjugation. The average size was found to be ∼18 nm (, inset) which is bigger than that of PpGNPs. After functionalisation, 5F-PpGNPs were found to be spherical, mono-dispersed, and stable. The zeta potential of the anionic nano-emulsion after functionalisation was different from the earlier one (−10.5 mV for 5F-PpGNPs); its exact value was −9.41 mV (, inset). Further, the bioconjugation of PpGNPs with 5-FU was confirmed by FTIR spectroscopy. The spectrum of PpGNPs () was compared with that of 5F-PpGNPs () and it was found that the broadband contour in both the spectra was in the range of 3600–3000 cm−1, which was corresponding to the –NH stretch of the peptide bonds of papain. Further, the peaks at 1737 cm−1 represent C = O anhydride stretching vibration.

Figure 3. FTIR spectra of (A) GNPs encapsulated papain, (B) papain after bioconjugation of PpGNPs with 5FU and (C) UV–vis spectra of pure 5-FU.

Figure 3. FTIR spectra of (A) GNPs encapsulated papain, (B) papain after bioconjugation of PpGNPs with 5FU and (C) UV–vis spectra of pure 5-FU.

3.1.3. Drug loading efficiency

After functionalisation, the percentage loading of 5-FU on PpGNPs was detected to be 76.2% by EquationEquation (1). The estimations of A and B were acquired as 2.95 and 0.70 separately and inserted in EquationEquation (1).

Another method based on UV–vis spectroscopy was employed to quantitatively estimate the bioconjugation of 5-FU with PpGNPs (). The absorbance spectra of the pure 5-FU were recorded at 265 nm [Citation39] using five different concentrations (1.56, 4, 8, 15 and 25 µg/mL). The amount of 5-FU conjugated to PpGNPs was found to be ∼79%, thus revealing an efficient binding of 5-FU with PpGNPs.

3.1.4. Study of synergistic effects of 5-FU and PpGNPs

The synergistic effect of 5-FU and PpGNPs was evaluated according to Chou and Talalay’s method of combination index (CIA). It was found that 156.8 µg/mL of 5-FU and 12.7 µg/mL of PpGNPs were used to achieve 50% inhibition of the A549 cancer cells. The value of the CIA against A549 cells was calculated to be 0.076. This proved that PpGNPs and 5-FU work in a synergistic manner.

3.1.5. In vitro anticancer study of PpGNPs, 5F-PpGNPs, and 5-FU

The in vitro anticancer activities of PpGNPs, 5F-PpGNPs and 5-FU were studied against lung cancer cell line (A549). The functionalised 5F-PpGNPs were found to be more effective against the A549 cells. The IC50 values of PpGNPs (), 5F-PpGNPs (), 5-FU () and papain () were found to be 255 µM, 11.7 µM, 25.6 µM and 548 µM, respectively. The anticancer effects were calculated in a dose-dependent manner. The enhanced activity of 5F-PpGNPs in comparison with 5-FU and PpGNPs is because of the synergistic effect of both papain and 5-FU. However, PpGNPs and 5F-PpGNPs were found to be non-toxic towards the 3T3-L1 cells up to a concentration of 50 µg/mL (data not shown).

Figure 4. The cytotoxicity (dose-dependent) study of (A) PpGNPs (B) pure 5-FU and 5F-PpGNPs and (C) pure papain on A549 cells. All the data were expressed in mean ± SD of three experiments.

Figure 4. The cytotoxicity (dose-dependent) study of (A) PpGNPs (B) pure 5-FU and 5F-PpGNPs and (C) pure papain on A549 cells. All the data were expressed in mean ± SD of three experiments.

3.1.6. Measurement of cytomorphological changes in the A549 cells

The morphological changes in the A549 cells were observed at the respective IC50 values of papain, PpGNPs, 5F-PpGNPs and 5-FU under phase-contrast microscopy (). The control (untreated) cells appeared to be uniformly spread and normal in surface with no distinctive or momentous changes in morphology even after 48 h of incubation (). However, noteworthy changes were seen in the cells treated with papain, PpGNPs, 5F-PpGNPs and 5-FU (). After 48 h of treatment, these treated cells shrunk in size became irregular and necrotic and got detached from the surface of the wells. Some cells were found with their plasma membranes unblemished, demonstrating that apoptosis had begun. It was observed that PpGNPs induced shrinking and apoptosis in the cells, yet this phenomenon occurred at considerably higher concentrations of PpGNPs (than those required for induction of apoptosis by 5F-PGNPs).

Figure 5. Image showing cytotoxic effect of (A) untreated control, (B) papain treated, (C) PpGNPs treated, (D) 5F-PpGNPs treated, and (E) 5-FU treated A549 cells at their respective IC50 concentrations at 20× magnifications.

Figure 5. Image showing cytotoxic effect of (A) untreated control, (B) papain treated, (C) PpGNPs treated, (D) 5F-PpGNPs treated, and (E) 5-FU treated A549 cells at their respective IC50 concentrations at 20× magnifications.

3.1.7. Analysis of PpGNPs, 5F-PpGNPs and 5-FU-mediated disruption of the mitochondrial membrane in A549 cells

The process of apoptosis is elicited with the activation of caspases, a group of cysteine-aspartate proteases. Moreover, the intracellular apoptosis is stimulated by the release of the mitochondrial cytochrome C (a characteristic feature of the mitochondrial apoptotic pathway) and the loss of mitochondrial membrane potential (ΔΨm). The mitochondrial membrane disruption (Δψm) was analysed at the corresponding IC50 values of the A549 cells treated with PpGNPs, 5F-PpGNPs and 5-FU in comparison with that of the untreated cells by the Mito Tracker Red CMXRos (). It was observed that the maximum intensity was observed for untreated cells followed by the cells treated with PpGNPs, 5-FU and 5F-PpGNPs in this order. This dye can distinguish between the changes in the Δψm in the cells; these changes were recorded under an inverted fluorescence microscope (Nikon ECLIPSE Ti-S, Japan) by the Image J software. In comparison with the untreated cells (), the A549 cells treated with PpGNPs (), 5F-PpGNPs () and 5-FU () showed a diminished fluorescence intensity. The minimum intensity was observed in the 5F-PpGNPs-treated cells; a gradual increase in intensity was observed for the 5-FU-treated and PpGNPs-treated A549 cells. However, untreated cells produced maximum intensity.

Figure 6. Mitochondrial depolarisation by disrupting mitochondrial membrane potential (ΔΨm) in A549 cells and images observed by staining with Mitotracker Red CMXROS. (A) Untreated control cells, (B) PpGNPs treated cells, (C) 5-FU treated cells, (D) 5F-PpGNPs treated cells. (E) Graph showing change in intensity of DCFDA stained control, PpGNPs, 5-FU and 5F-PpGNPs treated cells.

Figure 6. Mitochondrial depolarisation by disrupting mitochondrial membrane potential (ΔΨm) in A549 cells and images observed by staining with Mitotracker Red CMXROS. (A) Untreated control cells, (B) PpGNPs treated cells, (C) 5-FU treated cells, (D) 5F-PpGNPs treated cells. (E) Graph showing change in intensity of DCFDA stained control, PpGNPs, 5-FU and 5F-PpGNPs treated cells.

3.1.8. Estimation of production of reactive oxygen species

In general, ROS-mediated toxicity is one of the modes by which nanomaterials act as anticancer agents. The intracellular ROS generation in A549 cells treated with PpGNPs, 5F-PpGNPs, and 5-FU was estimated using the 5 (6)-carboxy-2′,7′-dichlorodihydrofluorescein diacetate (H2DCFDA) (Sigma-Aldrich) as an oxidation-sensitive fluorogenic marker of ROS in the viable cells (). When A549 cells were treated with 5F-PpGNPs, 5-FU and PpGNPs at their respective IC50 values, the generation of intracellular ROS increased gradually; maximum ROS generation was found with 5F-PpGNPs followed by 5-FU, PpGNPs and untreated cells (). The fluorescence was quantified by the ImageJ software. Moreover, micrographs of the untreated () PpGNPs- (), 5F-PpGNPs- () and 5-FU-treated () A549 cells were captured by the fluorescence microscope (Nikon ECLIPSE Ti-S, Japan). We could observe that the untreated cells showed minimum fluorescence, followed by PpGNPs-, 5-FU- and 5F-PpGNPs-treated cells (). The 5F-PpGNPs-treated cells were found to emanate a splendid fluorescence with a vandalised morphological structure because of the disturbing effect in the compactness of the plasma membrane brought about by the generation of the ROS. However, the untreated cells did not exhibit any fluorescence and maintained their natural morphology. The impact of cytotoxicity may be applied through the generation of oxidative pressure and apoptosis with a plausible association of overproduction of reactive oxygen species (ROS).

Figure 7. Images showing DCFDA staining under phase contrast microscope after 48 h of treatment on A549 cells at 20× magnification. (A) Control of DCFDA, (B) PpGNPs treated cells, (C) 5F-PpGNPs treated cells, (D) 5-FU treated cells. (E) Graph showing change in intensity of DCFDA stained control, PpGNPs, 5-FU and 5FPpGNPs treated cells.

Figure 7. Images showing DCFDA staining under phase contrast microscope after 48 h of treatment on A549 cells at 20× magnification. (A) Control of DCFDA, (B) PpGNPs treated cells, (C) 5F-PpGNPs treated cells, (D) 5-FU treated cells. (E) Graph showing change in intensity of DCFDA stained control, PpGNPs, 5-FU and 5FPpGNPs treated cells.

3.1.9. Analysis of changes in the nuclear morphology

The interaction of nanomaterials with the genetic material is inevitable and size-dependant. Moreover, 5-FU is an antimetabolite drug that incorporates fluoronucleotides instead of nucleotides. This interaction ultimately inhibits thymidylate synthase (TS), an enzyme involved in nucleic acid synthesis. Therefore, 5F-PpGNPs, 5-FU, and PpGNPs were found to interact with the nuclear material and cause substantial variations in the morphology and the nuclear structure of the A549 cells. The mode of cellular internalisation and, ultimately interaction with chromatin, were assessed by utilising a fluorescent dye (4′,6-diamidino-2-phenylindole) DAPI (). The A549 cells were treated with PpGNPs, 5F-PpGNPs and 5-FU at their corresponding IC50 values for 24 h at 37 °C and stained by the DAPI dye; the untreated cells were also stained to prepare them as controls (). The observed blue fluorescence intensities were measured under an inverted fluorescence magnifying instrument (Nikon ECLIPSE Ti-S, Japan). The apoptotic effects in comparison with the untreated cells () could be easily noticed in the PpGNPs (), 5F-PpGNPs- (), and 5-FU-treated () cells by observing the expanded cell membrane penetrability that produced condensed chromatin and a dark blue fluorescent consolidated nucleus. It was interesting to observe that PpGNPs also initiated apoptosis: this was further intensified by the bioconjugation of 5-FU with PpGNPs. The most critical and particular indication of the cytotoxic impact of stress is the condensation of the nucleus. The graphical representation of the observed intensities is shown in .

Figure 8. Images showing DAPI staining under phase contrast microscope after 48 h of treatment on A549 cells at 20× magnification. (A) Control of DAPI, (B) PpGNPs treated cells, (C) 5F-PpGNPs treated cells, (D) 5-FU treated cells. (E) Graph showing change in intensity of DAPI stained control, PpGNPs, 5-FU and 5FPpGNPs treated cells.

Figure 8. Images showing DAPI staining under phase contrast microscope after 48 h of treatment on A549 cells at 20× magnification. (A) Control of DAPI, (B) PpGNPs treated cells, (C) 5F-PpGNPs treated cells, (D) 5-FU treated cells. (E) Graph showing change in intensity of DAPI stained control, PpGNPs, 5-FU and 5FPpGNPs treated cells.

3.2. Discussion

In the given study, papain – a cysteine protease – was used to reduce the gold salt into gold nanoparticles. In this process, papain also acted as an encapsulating agent. The role of the protein in the nucleation of NPs is critical as the nucleation process depends upon the redox potential of the protein. The reducing properties of a protein increase with an increase in temperature and the overall charge of the protein. The regulation of overall charge and the ability to transfer electrons are the key factors during the nucleation of NPs. The reaction was carried out at 6 °C in the present study because, at its optimum temperature (37 °C), papain acted as a very strong reducing agent and resulted in the formation of NPs with a larger size and lesser stability. At 6 °C, the overall charge of papain decreased and this restricted its electron-transferring ability. Hence, the reaction progressed in a regulated and slow manner and also provided enough time for papain to reduce and encapsulate the NPs. During the complete synthesis of the NPs, the initially formed NPs behaved as a catalyst towards the completion of the reaction.

Cancer is one of the most deadly diseases worldwide and claims maximum deaths annually. Though radiotherapy, surgery and chemotherapy contribute a lot in fighting against cancer, their development is not sufficient to curb this most devastating disease of the world [Citation39]. Conventional chemotherapy has become voluntary torture to the patient, due to its non-selective action on normal cells. Target-specific drug delivery systems with trivial non-specific interactions must be evolved that may lead to a minimum or inconsequential side effects to the patients. Non-specific interactions cause huge collateral damages to patients [Citation40]. The evolution and progress of nanotechnology have extended a new archetype of potential in the development of nanomedicines that comprise of nanometre sizes and avail special treatment and behaviour from the immune system of the patient. These nanomedicines can be engineered to produce non-significant immunogenic responses, sustained release of the drug, insignificant non-specific interactions and site-specificity [Citation41]. Several drug delivery systems have been developed with huge success such as PEGylated PLGA, NP-encapsulated paclitaxel, etoposide [Citation42], and doxorubicin-conjugated bisphosphonate nanoparticles [Citation43] against osteosarcoma; methoxy-poly (ethylene glycol) aldehyde conjugated with doxorubicin and curcumin against HepG-2 cancer cells [Citation44]. In the same line of action, papain-encapsulated nanoparticles have been developed as effective drug delivery systems because they were found to be biocompatible, displayed anticancer property and had enhanced pharmacokinetics and pharmacodynamics of the bioconjugated 5-FU drug with vigorous intracellular delivery. In general, the gold nanoparticles, by the excellence of their metallic properties and size, hinder the propagation of the cancer cells by eliciting the NF-κB and Nrf-2 signalling pathways [Citation45], and they remain safe up to a concentration of 1012 particles/mL [Citation46]. 5-FU is an antimetabolite drug that inhibits the thymidylate synthase by interacting with its metabolite, 5-fluoro-2′-deoxyuridine-5′-monophosphate, which ultimately hinders normal RNA processing. However, for the treatment of advanced stages of cancer, its bioavailability decreases to 15% and it also causes rigorous side-effects on the gastrointestinal tract, haematological, neural, cardiac and dermatological processes [Citation47]. Also, it is catabolized (more than 80%) in the liver by the dihydropyrimidine dehydrogenase (DPD) enzyme into dihydrofluorouracil (DHFU) [Citation48]. The bioconjugation protects 5-FU from degradation and prevents non-specific interactions, thus leading to a minimum or no side effects. Nanoparticles, by virtue of their sizes, avoid interactions with the reticuloendothelial system and tend to improve the concentration of drugs in the cancer cells by passive and active targeting mechanism as well as by diminishing the drug efflux from the cancer cells. In passive targeting, the nanomedicines leak from the blood vessels supplying blood to the cancer cells and accumulate in the cells by the enhanced permeability and retention (EPR) effect [Citation49]. On the other hand, in active targeting, the internalisation of nanomedicines takes place via receptor-mediated endocytosis. The anionic caveolae-mediated endocytosis and the cationic clathrin-mediated endocytosis processes rely on the net charges, shapes, and the presence of conjugated ligands over the surface of these nanoparticles, resulting in increased cellular uptake and therefore increased drug accumulation in the cancer cells. This mechanism works with the interaction between the tumour ligands conjugated on the surface of the nanoparticles and cell-surface receptors or antigens on the cancer cell surfaces [Citation28].

The selection of capping agent/s was based on its/their abilities to bioconjugate with the drug for its safe and prompt delivery to the site. Various proteins such as transferrin (a serum glycoprotein), HSA, bromelain, trypsin [Citation21–22] have been used successfully as capping and functionalising agents. Similarly, papain was selected by virtue of its reducing and anticancer properties. Papain is a sulfhydryl protease from the latex of the plant Carica papaya and displays a powerful digestive action that is stronger than that of the pancreatic enzymes. Therefore, it has been used for a long time in herbal medicine in different countries [Citation50,Citation51]; yet, very limited information on its molecular targets and anticancer effects is available. Papain can inhibit the propagation and invasion of TFK-1 and CC cells by inhibiting the phosphorylated forms of NFκB, STAT-3, AKT and ERK. It also up regulates the phosphorylated AMPK in the SZ-1 cells. After adopting multiple targeting processes using different pathways to inhibit cancer cell proliferation, this combination was found to be highly active and site-directed.

4. Conclusion

In the search for a highly effective drug delivery system, papain-encapsulated and 5-FU-bioconjugated gold NPs were synthesised using papain as the reducing and capping agent because papain itself is anticancer by nature. Gold NPs, being non-toxic up to a very high concentration, were used for the bioconjugation with 5-FU (a known anticancer agent) because this bioconjugation protected 5-FU from the formation of non-specific interactions with degrading cellular materials. Therefore, GNPs, papain, and 5-FU worked synergistically and prohibited the proliferation of the cancer cells through different signalling pathways. This combination/bioconjugate was nano-sized in dimension and anionic in nature, which enabled its internalisation into the nucleus of the cancer cells via the caveolae-mediated endocytosis pathway. Therefore, the given combo/bioconjugate can be proposed to be a highly effective drug delivery system against lung cancer.

Acknowledgements

The authors acknowledge Minzu University of China, Beijing and Guangxi University of Chinese Medicine, Nanning, China for providing infrastructure to conduct this research.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work was supported by National Natural Science Foundation of China (Project Nos. 31460074 and 81573535), National Key Research and Development Program of China (No. 2017YFC1704000), The Self-Selected Project of Key Laboratory of Ethnic Medicine, Ministry of Education of China (No. KLEM-ZZ201805), GuangXi Key Laboratory of Zhuang and Yao Ethnic Medicine (No. [2014] 32), Collaborative Innovation Center of Zhuang and Yao Ethnic Medicine (No. [2013] 20), Guangxi Talent Highland for Zhuang and Yao Medicine and Combination of Medical Care and Elderly Care (NoTing Fa [2017] 44), Development Program of High-Level Talent Team under Qihuang Project of Guangxi University of Chinese Medicine (No. 2018005).

References

  • DeSantis CE, Lin CC, Mariotto AB, et al. Cancer treatment and survivorship statistics, 2014. CA Cancer J Clin. 2014;64(4):252–271.
  • Zarogoulidis P, Pataka A, Terzi E, et al. Intensive care unit and lung cancer: when should we intubate? J Thorac Dis. 2013;5:S407.
  • Rodzinski A, Guduru R, Liang P, et al. Targeted and controlled anticancer drug delivery and release with magnetoelectric nanoparticles. Sci Rep. 2016;6:20867.
  • Baldo BA, Pham NH. Adverse reactions to targeted and non-targeted chemotherapeutic drugs with emphasis on hypersensitivity responses and the invasive metastatic switch. Cancer Metastasis Rev. 2013;32(3–4):723–761.
  • Longley DB, Harkin DP, Johnston PG. 5-Fluorouracil: mechanisms of action and clinical strategies. Nat Rev Cancer. 2003;3(5):330–338.
  • Parker WB, Cheng YC. Metabolism and mechanism of action of 5-fluorouracil. Pharmacol Ther. 1990;48(3):381–395.
  • Pinedo HM, Peters GF. Fluorouracil: biochemistry and pharmacology. J Clin Oncol. 1988;6(10):1653–1664.
  • Chu E, Callender MA, Farrell MP, et al. Thymidylate synthase inhibitors as anticancer agents: from bench to bedside. Cancer Chemother Pharmacol. 2003;52:80–89.
  • Rajan S, Thirunalasundari T, Jeeva S. Anti-enteric bacterial activity and phytochemical analysis of the seed kernel extract of Mangifera indica Linnaeus against Shigella dysenteriae (Shiga, corrig.) Castellani and Chalmers. Asian Pac J Trop Med. 2011;4(4):294–300.
  • Hejazin RK, El-Qudah M. Effect of proteases on meltability and stretchability of Nabulsi cheese. Am J Agric Biol Sci. 2009;4:173–178.
  • Menard R, Khouri HE, Plouffe C, et al. A protein engineering study of the role of aspartate 158 in the catalytic mechanism of papain. Biochemistry. 1990;29(28):6706–6713.
  • Beuth J. Proteolytic enzyme therapy in evidence-based complementary oncology: fact or fiction? Integr Cancer Ther. 2008;7(4):311–316.
  • Verspaget HW. Proteases as prognostic markers in cancer: Proteolytic enzymes and their inhibitors influence the spread of cancer. BMJ. 1998;316(7134):790–791.
  • Gebauer F, Micheel B, Stauder G, et al. Proteolytic enzymes modulate the adhesion molecule CD44 on malignant cells in vitro. Int J Immunother. 1997;13:111–119.
  • Iram S, Zahera M, Wahid I, et al. Cisplatin bioconjugated enzymatic GNPs amplify the effect of cisplatin with acquiescence. Sci Rep. 2019;9:1–16.
  • Thambiraj S, Shruthi S, Vijayalakshmi R, et al. Evaluation of cytotoxic activity of docetaxel loaded gold nanoparticles for lung cancer drug delivery. Cancer Treat Res Commun. 2019;21:100157.
  • Graczyk A, Pawlowska R, Jedrzejczyk D, et al. Gold nanoparticles in conjunction with nucleic acids as a modern molecular system for cellular delivery. Molecules. 2020;25(1):204.
  • Wang Y, Han L, Liu F, et al. Targeted degradation of anaplastic lymphoma kinase by gold nanoparticle-based multi-headed proteolysis targeting chimeras. Colloids Surf B Biointerfaces. 2020;188:110795.
  • Qin M, Zhang J, Li M, et al. Proteomic analysis of intracellular protein corona of nanoparticles elucidates nano-trafficking network and nano-bio interactions. Theranostics. 2020;10(3):1213–1229.
  • Samanta K, Setua S, Kumari S, et al. Gemcitabine combination nano therapies for pancreatic cancer. Pharmaceutics. 2019;11(11):574.
  • Khan S, Rizvi SMD, Avaish M, et al. A novel process for size controlled biosynthesis of gold nanoparticles using bromelain. Mater Lett. 2015;159:373–376.
  • Khan S, Rizvi SM, Saeed M, et al. A novel approach for the synthesis of gold nanoparticles using trypsin. Adv Sci Lett. 2014;20(5):1061–1065.
  • Rahim M, Iram S, Khan MS, et al. Glycation-assisted synthesized gold nanoparticles inhibit growth of bone cancer cells. Colloids Surf B Biointerfaces. 2014;117:473–479.
  • Bechtel TJ, Weerapana E. From structure to redox: the diverse functional roles of disulfides and implications in disease. Proteomics. 2017;17(6):1600391.
  • Hermanson GT. Information on protein crosslinking, labeling and surface attachment. Bioconjugate technique. Academic Press; 1996.
  • Cojocaru IC, Ochiuz L, Spac A, et al. The validation of the UV spectrophotometric method for the assay of 5 fluorouracil. Farmacia. 2012;3:60.
  • Haiss W, Thanh NTK, Aveyard J, et al. Determination of size and concentration of gold nanoparticles from UV-vis spectra. Anal Chem. 2007;79(11):4215–4221.
  • Iram S, Zahera M, Khan S, et al. Gold nanoconjugates reinforce the potency of conjugated cisplatin and doxorubicin. Colloids Surf B Biointerfaces. 2017;160:254–264.
  • Kholmukhamedov A, Schwartz JM, Lemasters JJ. MitoTracker probes and mitochondrial membrane potential. Shock. 2013;39(6):543.
  • Chou T-C. Drug combination studies and their synergy quantification using the Chou-Talalay method. Cancer Res. 2010;70(2):440–446.
  • Sun T, Zhang YS, Pang B. Engineered nanoparticles for drug delivery in cancer therapy. Angew Chem Int Ed. 2014;53:12320–12364.
  • Lu A, Salabas EL, Schüth F. Magnetic nanoparticles: synthesis, protection, functionalization, and application. Angew Chem Int Ed Engl. 2007;46(8):1222–1244.
  • Grisolia J, Viallet B, Amiens C, et al. 99% random telegraph signal-like noise in gold nanoparticle micro-stripes. Nanotechnology. 2009;20(35):355303.
  • Bullis K. Nanoparticle networks promise cheaper batteries for storing renewable energy. MIT Technol Rev; 2015. Available from http://www.technologyreview.com/news/526811/nanoparticle
  • Bajaj A, Miranda OR, Kim I-B, et al. Detection and differentiation of normal, cancerous, and metastatic cells using nanoparticle-polymer sensor arrays. Proc Natl Acad Sci USA. 2009;106(27):10912–10916.
  • Treuel L, Brandholt S, Maffre P, et al. Impact of protein modification on the protein corona on nanoparticles and nanoparticle-cell interactions. ACS Nano. 2014;8(1):503–513.
  • Leite FL, Bueno CC, Da Róz AL, et al. Theoretical models for surface forces and adhesion and their measurement using atomic force microscopy. Int J Mol Sci. 2012;13(10):12773–12856.
  • Khan S, Haseeb M, Baig MH, et al. Improved efficiency and stability of secnidazole – an ideal delivery system. Saudi J Biol Sci. 2015;22(1):42–49.
  • Famta P, Mishra V, Khatik GL. Formulation and evaluation of 5-fluorouracil and methotrexate gold nanoparticles [doctoral dissertation]. Phagwara (India): Lovely Professional University; 2017.
  • Peer D, Karp JM, Hong S, et al. Nanocarriers as an emerging platform for cancer therapy. Nat Nanotechnol. 2007;2(12):751–760.
  • Veiseh O, Gunn JW, Zhang M. Design and fabrication of magnetic nanoparticles for targeted drug delivery and imaging. Adv Drug Deliv Rev. 2010;62(3):284–304.
  • Wang B, Yu X-C, Xu S-F, et al. Paclitaxel and etoposide co-loaded polymeric nanoparticles for the effective combination therapy against human osteosarcoma. J. Nanobiotechnology. 2015;13(1):22.
  • Rudnick-Glick S, Corem-Salkmon E, Grinberg I, et al. Doxorubicin-conjugated bisphosphonate nanoparticles for the therapy of osteosarcoma. JSM Nanotechnol Nanomed. 2014;1022:2.
  • Zhang Y, Yang C, Wang W, et al. Co-delivery of doxorubicin and curcumin by pH-sensitive prodrug nanoparticle for combination therapy of cancer. Sci Rep. 2016;6:21225.
  • Nel A, Xia T, Mädler L, et al. Toxic potential of materials at the nanolevel. Science. 2006;311(5761):622–627.
  • Dykman LA, Khlebtsov NG. Gold nanoparticles in biology and medicine: recent advances and prospects. Acta Nat. 2011;3(2):34–55.
  • Cho K, Wang XU, Nie S, et al. Therapeutic nanoparticles for drug delivery in cancer. Clin Cancer Res. 2008;14(5):1310–1316.
  • Hiremath CG, Kariduraganavar MY, Hiremath MB. Synergistic delivery of 5-fluorouracil and curcumin using human serum albumin-coated iron oxide nanoparticles by folic acid targeting. Prog Biomater. 2018;7(4):297–306.
  • Danhier F, Feron O, Préat V. To exploit the tumor microenvironment: passive and active tumor targeting of nanocarriers for anti-cancer drug delivery. J Control Release. 2010;148(2):135–146.
  • da Silva CR, Oliveira MBN, Motta ES, et al. Genotoxic and cytotoxic safety evaluation of papain (Carica papaya L.) using in vitro assays. Biomed Res. Int. 2010;2010:1–8.
  • Müller A, Barat S, Chen X, et al. Comparative study of antitumor effects of bromelain and papain in human cholangiocarcinoma cell lines. Int J Oncol. 2016;48(5):2025–2034.