Identification and biochemical characterisation of Acanthamoeba castellanii cysteine protease 3

DOI: https://doi.org/10.21203/rs.3.rs-58845/v1

Abstract

Background: Acanthamoeba spp. are free-living amoeba that are ubiquitously distributed in the environment. This study examines pathogenic Acanthamoeba cysteine proteases (AcCPs) belonging to the cathepsin L-family and explores the mechanism of AcCP3 interaction with host cells.

Methods: Six AcCP genes were amplified by PCR. Quantitative real-time polymerase chain reaction was used to analyse the relative mRNA expression of AcCPs during the encystation process, and between pre- and post-reactivated trophozoites. To further verify the role of AcCP3 in these processes, AcCP3 recombinant proteins were expressed in Escherichia coli, and the hydrolytic activity of AcCP3 was determined. The influence of the AcCP3 on the hydrolytic activity of trophozoites and the toxicity of trophozoites to human corneal epithelial cells (HCECs) was examined by inhibiting AcCP3 expression using siRNA. Further, the levels of p-Raf and p-Erk were examined in HCECs following coculture with AcCP3 gene knockdown trophozoites by western blotting.

Results: During encystment, five out of six AcCPs exhibited decreased expression, and only AcCP6 was substantially up-regulated at the mRNA level, indicating that most AcCPs were not directly correlated to encystation. Further, six AcCPs exhibited increased expression level following trophozoite reactivation with HEp2 cells, particularly AcCP3, indicating that these AcCPs might be virulent factors. After refolding the AcCP3 recombinant protein, the 27 kDa mature protein from the 34 kDa pro-protein hydrolysed host haemoglobin, collagen and albumin, and showed high activity in an acidic environment. After AcCP3 knockdown, the hydrolysis activity of trophozoite crude protein against gelatin was decreased, suggesting that these trophozoite had decreased toxicity. Compared with untreated trophozoites or negative control siRNA-treated trophozoites, AcCP3-knockdown trophozoites were less able to penetrate and damage monolayers of HCECs. Western blot analysis showed that after inhibiting the expression of trophozoite AcCP3, the activation levels of the Ras/Raf/Erk/p53 signalling pathways in HCECs were decreased.

Conclusions: AcCP6 was correlated to encystation. Further, AcCP3 is a virulent factor in trophozoites, and participated in the activation of the Ras/Raf/Erk/p53 signalling pathways of host cells.

Background

Acanthamoeba castellanii are free-living amoeba that can cause Acanthamoeba keratitis in humans after infection, as well as granulomatous amoebic encephalitis and cutaneous acanthamoebiasis in immunocompromised individuals [1, 2]. In recent years, serious Acanthamoeba infections have been associated with an increase in the number of contact lens wearers and immunocompromised patients. This parasite has a life cycle with two phases, the trophozoite and the cyst, which differ morphologically. When trophozoites are exposed to starvation conditions or a change in osmotic pressure, the trophozites transform into cysts to better withstand harsh environmental conditions. This parasite readily encysts in response to nutrition source deprivation, osmotic shock or a combination of both [3]. Thus, the growth and encystation of Acanthamoeba trophozites, which are regulated by a number of biological factors, can lead to persistent infections and influence the pathogenicity of this parasite.

The first stage of Acanthamoeba infection is the interaction between mannose binding protein (MBP) on the surface of the parasite and the mannose receptor on the surface of host cells [4]. Next, a variety of proteolytic enzymes are secreted into the host, hydrolysing the host’s tissue and resulting in tissue oedema, inflammation and necrosis [5]. Among these enzymes, serine proteases [6] and cysteine proteases (CPs) [7] are the main proteolytic proteins. Studies have shown that many serine proteases with different molecular weights are involved in the degradation of host cells and extracellular matrix [8] during pathogenic Acanthamoeba infection, and are inhibited by proteasome inhibitors in a concentration dependent manner [6].

Many protozoa utilise CPs for nutrient uptake, host infection, tissue invasion and environmental adaptation (encystation). It was well established that pathogenicity-related CPs, including cysteine protease 1 (CP1), CP2 and CP5 are expressed during the pathogenic process of Entamoeba histolytica [9]. In Giardia lamblia, CPs and UDP-Nacetylglucosamine pyrophosphorylase are required during encystation. Acanthamoeba express many CPs with different molecular weights, including 38.5, 43, 50, 59, 70, 100 and 130 kDa proteases [1013]. Following sequencing and analysis, it was found that both the 990 bp A. castellanii CP gene (AcCP) [11] and the 1359 bp A. culbertsoni CP gene (AcCP2) [14] belong to the L cathepsin CP family. AcCP2 contains an Ex3Rx3Wx2N motif in the proregion and a proline/threonine-rich C-terminus. The amino acid sequence of AcCP contains a catalytic site with five residues, as well as ERFNIN and GNFD motifs. Moreover, recombinant AcCP protein was capable of hydrolysing host proteins, including haemoglobin, albumin, IgG, IgA and adhesion proteins, suggesting that AcCP may be an important pathogenic protease in Acanthamoeba [11].

The pathogenic process of Acanthamoeba mainly depends on the hydrolysis activity of the pathogenic proteases secreted by trophozites, which activate a series of cell signalling pathways in host cells. Several studies have revealed that various downstream molecules are involved, including G-protein-coupled receptors, beta adrenalin receptors [15], Toll-like receptor-4 (TLR4), TLR4-Myeloid differentiation primary response gene 88 (MyD88), Nuclear Factor-κB (NF-κB), Extracellular signal-regulated kinase (Erk) [16], PI3K [17]and cytosolic phospholipase (A2αcPLA2α) [18]. Due to the complex pathogenic mechanisms of Acanthamoeba infection, cytopathogenic effects in host cells are not solely mediated by these signalling molecules. Instead, changes in the expression levels of other signalling molecules in the host may also be involved. The Ras/Raf/Erk signalling pathway is well known as a key intracellular signal transduction pathway that regulates cell differentiation, proliferation and apoptosis, and which is itself regulated by various growth factors. When cells are stimulated, Ras protein kinase expression is up-regulated, thereby activating downstream Erk1/2 protein and inducing apoptosis. Further, the activation of Ras/Raf/ERK signalling can lead to the activation of the tumour suppressor p53. Increased phosphorylation of p53 can up-regulate apoptotic proteins and inhibit bcl-2 proteins, eventually inducing apoptosis [19]. However, whether the Ras/Raf/Erk/p53 signalling pathways of host cells are involved in the pathogenesis of Acanthamoeba has not yet been verified. In addition, it has not been reported whether the CPs secreted by trophozoites are involved in the activation of the Ras/Raf/Erk/p53 signalling pathways of host cells.

Here, we amplified the CP genes of A. castellanii and determined their roles in the pathogenesis and encystation of trophozoites. We confirmed that AcCP3 is specifically required for the pathogenicity and virulence of trophozoites. By potentially activating the Ras/Raf/Erk/p53 signalling pathway of host cells, AcCP3 is important in the pathogenesis of Acanthamoeba and may be involved in the pathophysiology of infection.

Methods

Amoeba and cell culture

Acanthamoeba castellanii (strain ATCC30011) was obtained from the American Type Culture Collection (ATCC). Trophozoites were cultured axenically in peptone-yeast-glucose (PYG) medium. Cultures were incubated at 26 °C and trophozoites were harvested in the late log phase after subculture for 72 h. HEp-2 cells purchased from ATCC were cultured in EMEM (Gibco) supplemented with 10% fetal bovine serum (FBS) (Thermo), 100 U/ml penicillin and 100 µg/ml streptomycin. Human corneal epithelial cells (HCECs) purchased from Bioleaf (Shanghai, China) were grown in DMEM (Gibco) with 10% FBS. Both cells were cultured in a 37 °C incubator with 5% CO2.

Encystation Assays

Encystation assays were performed as described previously [20] with slight modifications. Briefly, trophozoites from post-logarithmic growth phase cultures were treated with phosphate buffered saline (PBS; containing 137 mM NaCl, 2.7 mM KCl, 10 mM Na2HPO4, 1.8 mM KH2PO4) with 50 mM MgCl2 and 10% glucose in culture plates at 30 °C (5 × 105 trophozoites per ml of medium) for 72 h. The treated materials were collected at 24 h, 48 h and 72 h for RNA extraction.

Cloning of AcCP genes and expression level analysis

The CP genes of Acanthamoeba were cloned from the cDNA of trophozoites using the primers listed in Table S1 by PCR. PCR was performed in a 9902 Veriti 96-well Thermal Cycler (Applied Biosystems) (94 °C for 3 min; 35 cycles of 94 °C for 15 s, 55 °C for 30 s and 72 °C for 1 min; followed by 72 °C for 7 min). The amplified PCR products were purified and ligated into a pMD19-T vector (Takara), and the nucleotide sequences were obtained by automated sequencing.

A.castellanii mRNA was extracted from trophozoites and cysts using RNeasy® Plus Mini Kit (Qiagen, Hilden, Germany), and cDNA was synthesised using a PrimeScript® 1st strand cDNA synthesis kit (Takara). Quantitative real-time polymerase chain reaction (qRT-PCR) was performed to analyse the expression level of AcCPs. The primers for AcCPs genes and the GAPDH internal reference are listed in Table S2.

Expression, purification and refolding of recombinant AcCP3 protein

The correct plasmids containing AcCP3 were amplified with primers containing HindIII and BamHI restriction sites, and the PCR products were ligated into the pQE-30 expression vector (Qiagen). The sequences of all constructs were confirmed on both strands and analysed with Vector NTI software (Invitrogen, Waltham, USA). Plasmids were transformed into M15 (pREP4) cells (Qiagen) for protein expression. The selected clones were cultured in Luria-Bertani broth containing 100 µg/ml ampicillin and rAcCP3 expression was induced using 1 mM isopropyl-β-D-thiogalactoside for 3 h at 37 °C. The recombinant rAcCP3 protein was purified using a QIA Express kit in accordance with the manufacturer’s instructions. The purity and mass of protein was determined by sodium dodecyl sulphate polyacrylamide gel electrophoresis (SDS-PAGE). Since the N-terminal peptide sequence in recombinant CPs can inhibit the hydrolytic activity of the protease, it was necessary for the recombinant AcCP3 to be refolded to obtain the mature peptide with hydrolytic activity [21]. Refolding of the purified recombinant protein was performed as described previously [11, 21]. In brief, purified rAcCP3 (2 mg) was slowly added to 100 ml refolding buffer, containing 100 mM Tris-HCl (pH 8.0), 1 mM ethylenediaminetetracetic acid (EDTA), 250 mM L-arginine, 5 mM reduced glutathione (GST), and 1 mM oxidised glutathione. The protein was gently stirred at 4 °C overnight, and then dialysed against 10 mM Tris-HCl (pH 7.5). The obtained rAcCP3 was further processed as described previously [22]. In brief, sodium acetate buffers with pH 4.0 to 7.0 were used to analyse the optimal pH condition required for obtaining the fully matured rAcCP3 enzyme. The concentration of matured rAcCP3 enzyme was measured using a protein assay (Bio-Rad).

Analysis of biological characteristics of AcCP3

To assess the role of AcCP3 in Acanthamoeba pathogenesis, the reactivation of the physiological properties of trophozoites was performed using HEp-2 cell monolayers as described previously [23]. In brief, HEp-2 cells were cultured in 75 cm2 tissue culture flasks (Corning) at 37 °C under sterile conditions until the monolayer covered the bottom of the flask completely, at which point the supernatant was removed. Trophozoites (106) suspended in 25 mL physiological 0.9% NaCl were inoculated onto the monolayer, three times consecutively. Cocultures of amoebae and HEp-2 cells were incubated at 26 °C until the monolayer was completely lysed. Reactivated trophozoites were collected and total mRNA was extracted. qRT-PCR was performed to analyse the expression level of AcCP3. In addition, MBP was used as an example cytotoxicity protein in Acanthamoeba [24]. 18S rDNA was used as internal reference [25]. The primers for AcCP3, MBP and 18S rDNA are listed in Table S2.

To determine the effect of AcCP3 on host proteins, haemoglobin (from human blood), collagen (from human placenta) and albumin (from bovine serum) were purchased from Sigma. Each protein (2 mg/ml) was incubated with matured rAcCP3 (100 nM) in 50 mM sodium acetate (pH 4.0 or pH 7.0) with 1 mM GST for 3 h at 37 °C. The reactions were terminated by adding reducing sample buffer and the degradation activity of rAcCP3 was analysed by SDS-PAGE.

The protease activities of trophozoite crude proteins were analysed using Novex™ 10% Zymogram Plus (Gelatin) Protein Gels (Thermo Fisher scientific) in accordance with the manufacturer’s instructions. 1.5 µg crude protein extract was added to each lane. Various inhibitors for different proteases were used. Crude protein extracts treated and untreated with inhibitors (1 h before electrophoresis) were analysed with zymography. The final concentrations of inhibitors were as follows [26]: for serine proteases, 5 mmol/L phenylmethylsulfonyl fluoride (PMSF); for metalloproteases, 2 mmol/L EDTA; for CPs, 40, 60, 80 and 100 mmol/L N-ethylmaleimide (NEM).

Ac CP3 gene silencing

siRNA targeting the catalytic domain of AcCP3 was synthesised by Ribobio (Guangzhou, China) and based on the cDNA sequence. The sequence of the forward strand was 5’-AGUACAUCAUCAACAACAA-3’. Trophozoites were plated at a density of 5 × 104 cells in 48-well plates and cultured overnight, then transfected with siRNA (15 µg/ml) for 12 h using SuperFectin™ In Vitro siRNA Transfection Reagent (Pufei). As a control, a negative siRNA provided by Ribobio was also applied to cultured trophozoites. Untreated trophozoites and transfection reagent-treated trophozoites were also processed. After the transfection, the differentially treated trophozoites were harvested to determine the efficacy of the knockdowns by examining the expression level of AcCP3 with qRT-PCR, and then for the cytopathic tests.

Effect of AcCP3-knockdown in trophozoites on Acanthamoeba-mediated cytotoxicity

To determine the effects of reduced AcCP3 expression on Acanthamoeba-mediated HCECs death, cytotoxicity assays were performed as previously described [27]. Confluent HCEC monolayers in 12-well culture plates (Corning) were incubated with differentially treated trophozoites (ratio 1:2) at 37 °C in a 5% CO2 atmosphere for 24 h. Four different experimental groups were included. Group 1, normal cultured confluent HCECs; Group 2, HCECs co-cultured with normal trophozoites; Group 3, HCECs co-cultured with negative control siRNA-transfected trophozoites; Group 4, HCECs co-cultured with AcCP3-knockdown trophozoites. The cytopathic effects for the different groups of trophozoites were observed using light microscopy (Olympus). The trophozoites were detached with cold 2 mM EDTA-PBS buffer, chilled on ice for 20 min and then harvested. In addition, HCECs were cultured in fresh DMEM medium with CCK 8 reagent for 2 h, and the cytotoxicity was determined by measuring dehydrogenase release (Cell Counting Kit-8, DOJINDO). The absorbance of each well was measured at 450 nm using a Model 680 Microplate Reader (Bio-Rad). In order to investigate the signalling pathways in HCECs activated by Acanthamoeba, four groups of differentially treated HCECs were harvested after co-culture with amoeba for 24 h. qRT-PCR was performed to analyse the expression level of the Ras gene (NM_004985.4). The forward strand was 5’- AGGAAGCAAGTAGTAATTGATGGA-3’, the reverse strand was 5’- GCCTGTTTTGTGTCTACTGTTCT-3’. Human GAPDH used as an internal reference, with forward strand 5’- TCACCACCATGGAGAAGGC-3’, reverse strand 5’- GCTAAGCAGTTGGTGGTGCA-3’.

Western Blotting Assays

The activation states of Raf, Erk and p53 in differentially treated HCECs were determined using western blotting assays as previously described [28]. Briefly, four groups of differentially treated HCECs were harvested for western blotting after coculture with amoeba for 24 h. The cells were lysed in 100 µL lysis buffer (20 µL Phosphatase inhibitors, 20 µL Protease inhibitor cocktail, 100 µL PBS, 5 µL NP-40, 855 µL ddH2O). The cell lysate (20 µg) was resolved by SDS-PAGE and transferred to PVDF membranes (Roche). We used the following primary antibodies purchased from Cell Signalling: rabbit anti-p-Raf (Ser338), rabbit anti-Raf, rabbit anti-p-Erk (Thr202/Tyr204), rabbit anti-Erk, mouse anti-p-p53 (Ser15), and mouse anti-p53. HRP-labelled goat anti-mouse and goat anti-rabbit secondary antibodies (Abcam), and rabbit polyclonal to β-actin (Abcam), were detected with Tanon™ High-sig ECL Western Blotting Substrate (Tanon), observed with an ECL detection system (Tanon), and the scanned images were quantified using Image-Pro Plus 4.5.1 software (Media Cybernetics).

Statistical analysis

The results of qRT-PCR were calculated using the 2−△△Ct method. Statistical analyses were performed using GraphPad Prism software (San Diego). Significance was calculated by one-way analysis of variance followed by a Tukey test or Student’s t-test. Data are expressed as mean ± SD and at least three independent experiments were performed for each experiment.

Results

Identification of pathogenic AcCPs

The sequences encoding full-length CP genes were amplified using PCR procedures. Five new open reading frames (ORF) corresponding to CP genes were identified in A.castellanii. The sequences of AcCP3, AcCP10, AcCP9, AcCP5 and AcCP8 with the accession numbers LC472809, LC472813, LC472812, LC472811, LC472810, respectively, (Table S3) have been submitted to the GenBank database. AcCP8 does not present the typical characteristics of CP. The sequences of AcCP7 (XM004358251) and AcCP6 (XM004341651) from ATCC30011 have also been confirmed in the Neff strain.

The amino acid residues Q, C, H, N and W that act as the catalytic site of cathepsin L-family CPs exist in AcCPs. The modified ERFNIN and GNFD motifs that are conserved in cathepsin L-like CP family members have also been observed in AcCPs [29]. We found that while AcCP10 had no C active site, the amino acid sequences of the other CPs contained the conserved C site, which is mainly involved in the refolding of the protease. The nucleotide sequence homology of ATCC30011 and Neff 7 CPs was as high as 99%.

Expression pattern of AcCPs during encystation

qRT-PCR was used to analyse the expression levels of the AcCPs during the encystation process. As shown in Fig. 1, the mRNA levels of AcCP3, AcCP10, AcCP8, AcCP9 and AcCP7 were decreased at 24 h, 48 h and 72 h (with the exception of AcCP7 at 72 h), indicating that these AcCPs may not be involved in encystation. In contrast, AcCP6 was gradually but substantially up-regulated during encystation, suggesting that while most cathepsin L-family CPs were not involved in encystation, AcCP6 may play a uniquely important role in the formation of cysts. 

The relationship between AcCPs and protozoan virulence

In order to investigate the role of AcCPs in Acanthamoeba virulence, trophozoites that were axenically grown in vitro for several years were reactivated with HEp2 cells. The results of quantitative PCR analysis of AcCP and MBP mRNA expression in pre- and post-reactivated trophozoites are shown in Fig. 2. The MBP expression level was up-regulated in post-reactivated trophozoites. In addition, the relative mRNA expression levels of six of the CPs were up-regulated following trophozoite reactivation, while the expression level of AcCP3 was significantly increased in trophozoites that invaded HEp2 cells (Fig. 2). These data indicate that AcCPs may play an important role in Acanthamoeba pathogenesis by mediating host cell damage during infection. Further, AcCP3 in particular may act as a potential pathogenic factor during human cell invasion. 

Refolding of r Ac CP3 and degradation of host proteins

The AcCP3 gene contains a 993-bp ORF that encodes a 329-amino acid protein. A portion of the prodomain and the entire mature domain of AcCP3 was amplified and expressed using a pQE-30 expression vector. AcCP3 was expressed in M15 competent cells as an inclusion body protein with an apparent molecular mass of 34 kDa (Fig. 3a lane 1). The recombinant protein was purified by Ni-NTA affinity chromatography and then refolded. The refolded sample was further processed under reducing acidic conditions to allow the refolded protein to be processed into a fully active enzyme. The size of fully activated mature rAcCP3 coincided with the size predicted for the mature protease (27 kDa) (Fig. 3a lane 2). However, despite our extensive efforts to produce enzymatically active recombinant enzymes, we failed to successfully perform zymography with 0.1% gelatin as a substrate (data not shown). However, we were able to carry out an assay of host protein degradation using rAcCP3. SDS-PAGE showed that rAcCP3 possesses proteolytic activity against human haemoglobin, collagen and bovine serum albumin. All protein substrates used in this assay were hydrolysed at acidic pH, but no hydrolysis was observed at neutral pH (Fig. 3b, c, d).

Low proteolytic activity of AcCP3 knockdown trophozoites

In order to investigate the biological role of AcCP3 in Acanthamoeba trophozoites, trophozoites were transfected with AcCP3-siRNA and qRT-PCR revealed that the mRNA level was significantly reduced in AcCP3 gene knockdown trophozoites (Fig. 4a). There was no significant difference in the growth and proliferation rate of trophozoites between the knockdown and wild-type strains (data not shown). Further, there were no differences observed between superfectin-treated trophozoites (superfectin) and negative control siRNA-transfected trophozoites (Ne) compared to untreated trophozoites cultured with PYG. Subsequently, the hydrolytic activity of AcCP3 knockdown trophozoites was compared to that of wild-type trophozoites and protease inhibitor-treated trophozoites (Fig. 4b). 1.5 µg of differentially treated crude protein extracts were added to lane 1 to lane 8, and lane 10. The crude protein extracts in lanes 1 and 4, corresponding to wild-type trophozoites, showed high hydrolytic activity. Lane 2 showed that the serine proteinase inhibitor PMSF could inhibit this hydrolytic activity. Lane 3 showed that the metalloprotease inhibitor EDTA had no effect on this hydrolytic activity. Lanes 5 to 8 showed that 40, 60, 80 or 100 mmol/L NEM could inhibit this hydrolytic activity, and the hydrolytic activity decreased with increasing NEM concentration. The ability of PMSF and NEM to inhibit the hydrolytic activity of crude protein extracts from trophozoites confirmed previous results [26]. In lane 9, 1.5 µg crude protein extract from trophozoites transfected with negative control siRNA showed that the negative control siRNA has little effect on the hydrolytic activity. Lane 10 treated was the same as lane 5. In lane 11 is shown the effect AcCP3 siRNA on the hydrolytic activity of trophozoites. Zymography revealed that the proteolytic activity in trophozoites following AcCP3 knockdown was decreased, suggesting that the virulence of these trophozoites was also decreased. 

Effect of AcCP3 knockdown on protozoa-mediated cytotoxicity

Cytotoxicity assays were performed with four differently treated HCECs as described in the methods section. When HCECs were cultured alone, the cells grew into a confluent monolayer and were tightly connected to one another, while a small number of cells lifted off of the dish because of overgrowth (shown as Fig. 5a). The white arrow in Fig. 5b, c and d indicates the cellular voids formed when HCEC monolayer cells were destroyed by trophozoite penetration. Figure 5b shows confluent HCECs co-cultured with wild-type trophozoites; in this case, most adherent HCECs became suspended and the tight connections between the cells were disrupted, resulting in the formation of cellular voids. Figure 5c shows confluent HCEC monolayers were also destroyed by negative control siRNA-transfected trophozoites, and the cytotoxicity did not differ significantly from that of untreated trophozoites. As shown in Fig. 5d, trophozoites transfected with AcCP3 siRNA exhibited a decrease in Acanthamoeba-mediated HCEC cytotoxicity, and while the HCECs exhibited a few cellular voids, most of the HCECs remained attached. To test the cytopathic activity, a monolayer of HCECs was co-cultured with differently treated trophozoites. After 24 h, the trophozoites were removed. HCECs were sequentially cultured in fresh DMEM medium with CCK 8 reagent for 2 h and then analysed by photometry to assess the dehydrogenase released from cells. As shown in Fig. 5e, HCECs treated with AcCP3 knockdown trophozoites exhibited the highest dehydrogenase release compared to HCECs in treatment groups 2 and 3. These results indicate that the cytotoxicity induced by Acanthamoeba in HCECs was decreased in AcCP3 knockdown trophozoites, further suggesting that AcCP3 is a pathogenic factor for Acanthamoeba

Inhibiting the expression of AcCP3 reduces target cell Ras/Raf/Erk signalling

Four groups of HCECs treated as described in the methods section were harvested after co-cultured with trophozoites, qRT-PCR was used to analyse the mRNA expression levels of Ras genes. The data shows (Fig. 6a) that the expression level of K-Ras was significantly increased in HCECs co-cultured with wide-type trophozoites (group 2). On the contrary, the expression level of K-Ras was significantly decreased in HCECs cocultured with AcCP3 knockdown trophozoites (Group 4). This indicates that AcCP3 siRNA significantly decreased the mRNA expression level of K-Ras in Acanthamoeba-treated HCECs. In contrast, the mRNA expression levels of H-Ras and N-Ras showed no differences between groups 2 and 4 (data not shown). Previous studies have demonstrated that Ras recruits Raf kinase into a complex, which mediates Raf activation. Raf then phosphorylates MEK1 and MEK2, which in turn activate Erk1/2 by the tandem phosphorylation of threonine and tyrosine residues [30]. Thus, we further examined the levels of p-Raf and p-Erk in HCECs exposed to AcCP3 gene knockdown trophozoites by western blotting. That data shows that (Fig. 6, group 2) untreated trophozoites could significantly increase the phosphorylation levels of Raf, Erk, and p53 proteins in HCECs. Compared to untreated cultured confluent HCECs (group 1), HCECs co-cultured with untreated trophozoites (group 2) or HCECs co-cultured with negative control siRNA-transfected trophozoites (group 3), the levels of p-Raf and p-Erk in HCECs were markedly decreased after co-culture for 24 h (P < 0.01). Previous studies have suggested that phosphorylation of p53 might be an important mechanism in Ras/Raf/Erk-induced apoptosis. Thus, we also examined the levels of p-p53 in HCECs cocultured with AcCP3 knockdown trophozoites. The data shows that the level of p-p53 was also markedly decreased after treated for 24 h (P < 0.01). These results demonstrated that inhibiting the expression of AcCP3 protein could reduce the Acanthamoeba-mediated phosphorylation levels of Raf, Erk and p53 proteins in HCECs. This suggests that AcCP3 might participate in the activation of Raf, Erk and p53 proteins in HCECs.

Discussion

CP family proteins are essential to the life cycle of several protozoa by influencing several diverse processes, such as nutrient intake, protein degradation of parasites, immunomodulators, host cell invasion and encystment/excystment, in addition to well-established roles in protein processing and catabolism. They are also an important virulence factor for parasites [31]. It has recently been reported that A. castellanii expresses CPs with proteolytic effects [11]. However, little research has been done on the pathogenic and encystment mechanisms of Acanthamoeba CPs. Long-term sterile passage culture of Acanthamoeba trophozoites in vitro may reduce the toxicity of the trophozoites [32]. However, a potential method to reactivate attenuated trophozoite virulence was identified by growing the trophozoites on HEp-2 cell monolayers three times using the Acanthamoeba 1BU strain (ATCC no. Pra-105) [23]. Using this method, we found that the expression levels of six AcCPs were higher following reactivation, especially AcCP3, suggesting that AcCPs may play a pathogenic role in Acanthamoeba. In vitro, AcCP3 could effectively hydrolyse host haemoglobin, collagen and albumin after refolding, and it had strong hydrolytic activity in an acidic environment but almost no hydrolytic activity in a neutral environment. This characteristic is similar to that of vivapain from Plasmodium vivax [33] and CsCF-6 from Clonorchis sinensis [21].

The pathogenic process of Acanthamoeba in the host cell is regulated by various signalling molecules. G protein-coupled receptor and β adrenalin receptor inhibitors block the activity of trophozoite proteases, thereby affecting the growth, encystation, vitality, and pathogenicity of the parasite [15]. The HCECs used in our experiments primarily express the epithelial markers E-cadherin, ZO-1, and β-catenin. It has been reported that Toll-like receptor 4 (TLR4) is an important pathogenic target of Acanthamoeba in human corneal epithelial cells (HCECs). After attaching to the surface of host cells, Acanthamoeba activates TLR4, which in turn affects intracellular MyD88, NF-κB and Erk [34]. It has been found that thrombinase and trypsin can activate protease-activated receptor (including PAR1 and PAR2) in HCECs, and promote the secretion of pro-inflammatory factors such as IL-6, IL-8 and TNF-α, leading to ocular inflammation [35]. Furthermore, studies have shown that Acanthamoeba-induced host cell death is related to PI3K signalling [17]. Acanthamoeba-induced host cell apoptosis is associated with mitochondrial overexpression of pro-apoptotic proteins [36]. It was found that cytosolic phosolipase A2 alpha (cPLA2-α) is involved in Acanthamoeba-induced apoptosis of HCECs [18]. To determine whether Acanthamoeba AcCP3 interacted with HCECs via the TLR4 receptor, it will be necessary to determine whether CPs can bind to the TLR4 receptor on the surface of HCECs via pathogen-associated molecular patterns. In this study, zymography revealed that the proteolytic activity of trophozoites was decreased following AcCP3 gene knockdown, and trophozoites after AcCP3 gene silencing exhibited a decrease in Acanthamoeba-mediated HCEC cytotoxicity. The mechanism of AcCP3 in Acanthamoeba-mediated HCEC cytotoxicity must be confirmed with regards to its role in cell signalling pathways.

The Ras/Raf/Erk signalling pathway is the main signalling pathway regulating the occurrence and development of tumours, and is important in regulating apoptosis [19]. This pathway activates endogenous apoptotic pathways, such as the release of mitochondrial cytochrome c [37], as well as the activation of caspase-9 [38], caspase-8 [39], and p53 [40]. The ability of the amoeba to activate the Ras/Raf/p53 signalling pathways of host cells during infection has not been verified. Ras signalling molecules include H-Ras, K-Ras, and N-Ras [41]. The role of Ras in HCECs has been confirmed, and it has been reported that Ras can regulate the inflammatory response in the cornea of mice [42]. In addition, the Ras/Erk signalling pathway can be activated by hepatocyte growth factor [43] in HCECs. In this study, we found that Acanthamoeba infection mainly affected the expression level of the K-Ras gene in HCECs, and western blot analysis showed that trophozoites could also increase the phosphorylation levels of Raf, Erk1/2, and p53 in HCECs; taken together, this indicates that Acanthamoeba could activate the Ras/Raf/ERK/p53 signalling pathways of host cells. Moreover, studies have found that serine and cysteine protease inhibitors can down-regulate Ras pathway-induced apoptosis. Protease inhibitors can inhibit the activation of Ras, which may be related to the involvement of proteases in tumour cell invasion and metastasis [44]. In summary, proteases are involved in the activation of Ras protein and downstream signalling pathways. In this study, the phosphorylation of Ras/Raf/ERK/p53 signalling pathway components in HCECs was decreased when the expression of AcCP3 was inhibited in trophozoites. This suggests that AcCP3 is involved in the activation of the Ras/Raf/ERK/p53 signalling pathway during the trophozoite pathogenic process.

It is not well understood how CP expression and function changes during encystation in Acanthamoeba. We analysed the transcription levels of six genes and found that Acanthamoeba CP AcCP6 was the most highly expressed CP gene during the encystation process. This indicates that while most cathepsin L-family CPs are not involved in encystation, AcCP6 may participate in encystment. Of these CP genes, AcCP6 emerges as the most highly expressed and exhibits developmental regulation, with expression increasing dramatically during encystation and cyst stages. In general, the process of encystation involves the coordinated secretion of cyst wall materials to the periphery of a cell. It was suggested that trophozoites produce abundant cyst wall proteins, including AcCP6, which are packaged into encystation-specific materials in response to environmental cues. Indeed, cysteine endopeptidases in Giardia were localised to the encystation-specific vesicles during encystation. When fixed cysts were subjected to fluorescence in situ hybridization with AcCP3 and AcCP5 DNA probes, AcCP3 and AcCP5 were visualized in both submembranous and nucleus of cysts (Fig. S1). Our data indicated that some CPs could also be regulated by acidification of encystation-specific packages; it has been suggested that encystation vesicles fuse with peripheral vacuoles prior to formation of the cyst wall and that the activity of encystation-specific CP (AcCP6) toward protein substrates is greatly reduced in an acidic compartment.

Conclusions

These results suggested that various Acanthamoeba CPs may be required during pathogenic and cyst formation processes. We propose that various CPs play important roles in the regulation of pathogenic and encystment processes, especially AcCP3 and Ac AcCP6. Further study of the pathogenic mechanisms of Acanthamoeba trophozoites will provide a platform for the development of new anti-Acanthamoeba drugs. Further work will be required to clarify the role of the other CPs in Acanthamoeba. The present study preliminarily confirmed that Acanthamoeba trophozoites activate the Ras/Raf/Erk/p53 signalling pathways of HCECs and contribute to host cell death. The phosphorylation of the Ras/Raf/Erk/p53 signalling pathway decreased after inhibiting the expression of AcCP3, indicating that AcCP3 may be an important pathogenic factor involved in the pathogenesis of trophozoites.

Abbreviations

AcCPs: Acanthamoeba cysteine proteases; HCECs: human corneal epithelial cells; CP: cysteine proteases; TLR4: Toll-like receptor-4; MyD88: TLR4-Myeloid differentiation primary response gene. NF-κB: Nuclear Factor-κB; Erk: Extracellular signal-regulated kinase, PBS: phosphate buffered saline; qRT-PCR: quantitative real-time polymerase chain reaction; SDS-PAGE: sodium dodecyl sulphate polyacrylamide gel electrophoresis; EDTA: ethylenediaminetetracetic acid. GST: glutathione; PMSF: phenylmethylsulfonyl fluoride. ORF: open reading frames.

Declarations

Availability of data and materials

All data generated or analyzed during this study are included in the article.

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Funding

This research was funded by the National Natural Science Foundation of China. (NSFC, 81572020)

Authors’ contributions

XC and MF designed the study. ZW, DW, MF conducted the work. XC acquired funding. ZW and DW wrote the first draft of the manuscript. XC, MF and HT revised the manuscript. All authors read and approved the final manuscript.

Acknowledgements

Not applicable.

References

  1. Wu D, Feng M, Wang ZX, Qiao K, Tachibana H, Cheng XJ. Molecular and biochemical characterization of key enzymes in the cysteine and serine metabolic pathways of Acanthamoeba castellanii. Parasit Vectors. 2018; 11: 604.
  2. Khan NA. Acanthamoeba: biology and increasing importance in human health. FEMS Microbiol Rev. 2006; 30: 564-595.
  3. Anwar A, Khan NA, Siddiqui R. Combating Acanthamoeba cysts: what are the options? Parasit Vectors. 2018; 11: 26.
  4. Garate M, Cao Z, Bateman E, Panjwani N. Cloning and characterization of a novel mannose-binding protein of Acanthamoeba. J Biol Chem. 2004; 279: 29849-29856.
  5. Ng SL, Nordin A, Abd Ghafar N, Suboh Y, Ab Rahim N, Chua KH. Acanthamoeba-mediated cytopathic effect correlates with MBP and AhLBP mRNA expression. Parasit Vectors. 2017; 10: 625.
  6. Kim WT, Kong HH, Ha YR, Hong YC, Jeong HJ, Yu HS, et al. Comparison of specific activity and cytopathic effects of purified 33 kDa serine proteinase from Acanthamoeba strains with different degree of virulence. Korean J Parasitol. 2006; 44: 321-330.
  7. Moon EK, Hong Y, Chung DI, Kong HH. Cysteine protease involving in autophagosomal degradation of mitochondria during encystation of Acanthamoeba. Mol Biochem Parasitol. 2012; 185: 121-126.
  8. Khan NA, Jarroll EL, Panjwani N, Cao Z, Paget TA. Proteases as markers for differentiation of pathogenic and nonpathogenic species of Acanthamoeba. J Clin Microbiol. 2000; 38: 2858-2861.
  9. Betanzos A, Bañuelos C, Orozco E. Host invasion by pathogenic amoebae: Epithelial disruption by parasite proteins. Genes (Basel). 2019; 10: 618.
  10. Ramírez-Rico G, Martínez-Castillo M, de la Garza M, Shibayama M, Serrano-Luna J. Acanthamoeba castellanii proteases are capable of degrading iron-binding proteins as a possible mechanism of pathogenicity. J Eukaryot Microbiol. 2015; 62: 614-622.
  11. Hong Y, Kang JM, Joo SY, Song SM, Lê HG, Thái TL, et al. Molecular and biochemical properties of a cysteine protease of Acanthamoeba castellanii. Korean J Parasitol. 2018; 56: 409-418.
  12. Hong YC, Hwang MY, Yun HC, Yu HS, Kong HH, Yong TS, et al. Isolation and characterization of a cDNA encoding a mammalian cathepsin L-like cysteine proteinase from Acanthamoeba healyi. Korean J Parasitol. 2002; 40: 17-24.
  13. Alfieri SC, Correia CE, Motegi SA, Pral EM. Proteinase activities in total extracts and in medium conditioned by Acanthamoeba polyphaga J Parasitol. 2000; 86: 220-227.
  14. Yun HC, Kim KY, Park SY, Park SK, Park H, Hwang UW, et al. Cloning of a cysteine proteinase gene from Acanthamoeba culbertsoni. Mol Cells. 1999; 9: 491-496.
  15. Aqeel Y, Siddiqui R, Manan Z, Khan NA. The role of G protein coupled receptor-mediated signaling in the biological properties of Acanthamoeba castellanii of the T4 genotype. Microb Pathog. 2015; 81: 22-27.
  16. Ren MY, Wu XY. Toll-like receptor 4 signalling pathway activation in a rat model of Acanthamoeba keratitis. Parasit immunol. 2011; 33: 25-33.
  17. Sissons J, Kim KS, Stins M, Jayasekera S, Alsam S, Khan NA. Acanthamoeba castellanii induces host cell death via a phosphatidylinositol 3-kinase-dependent mechanism. Infect Immun. 2005; 73: 2704-2708.
  18. Tripathi T, Abdi M, Alizadeh H. Role of phospholipase A(2) (PLA(2)) inhibitors in attenuating apoptosis of the corneal epithelial cells and mitigation of Acanthamoeba keratitis. Exp Eye Res. 2013; 113: 182-191.
  19. Cagnol S, Chambard JC. ERK and cell death: mechanisms of ERK-induced cell death--apoptosis, autophagy and senescence. The FEBS journal. 2010; 277: 2-21.
  20. Aqeel Y, Siddiqui R, Iftikhar H, Khan NA. The effect of different environmental conditions on the encystation of Acanthamoeba castellanii belonging to the T4 genotype. Exp Parasitol. 2013; 135: 30–35.
  21. Na BK, Kang JM, Sohn WM. CsCF-6, a novel cathepsin F-like cysteine protease for nutrient uptake of Clonorchis sinensis. Int J Parasitol. 2008; 38: 493-502.
  22. Na BK, Bae YA, Zo YG, Choe Y, Kim SH, Desai PV, et al. Biochemical properties of a novel cysteine protease of Plasmodium vivax, vivapain-4. PLoS Negl Trop Dis. 2010; 4: e849.
  23. Koehsler M, Leitsch D, Duchene M, Nagl M, Walochnik J. Acanthamoeba castellanii: growth on human cell layers reactivates attenuated properties after prolonged axenic culture. FEMS Microbiol Lett. 2009; 299: 121-127.
  24. Kim JH, Matin A, Shin HJ, Park H, Yoo KT, Yuan XZ, et al. Functional roles of mannose-binding protein in the adhesion, cytotoxicity and phagocytosis of Acanthamoeba castellanii. Exp Parasitol. 2012; 132: 287-292.
  25. Wu D, Qiao K, Feng M, Fu Y, Cai J, Deng Y, et al. Apoptosis of Acanthamoeba castellanii trophozoites induced by oleic acid. J Eukaryot Microbiol. 2018, 65: 191-199.
  26. Serrano-Luna JJ, Cervantes-Sandoval I, Calderon J, Navarro-Garcia F, Tsutsumi V, Shibayama M. Protease activities of Acanthamoeba polyphaga and Acanthamoeba castellanii. Can J Microbiol. 2006; 52: 16-23.
  27. Alsam S, Kim KS, Stins M, Rivas AO, Sissons J, Khan NA. Acanthamoeba interactions with human brain microvascular endothelial cells. Microb Pathog. 2003; 35: 235-241.
  28. Feng D, Wang B, Ma Y, Shi W, Tao K, Zeng W, et al. The Ras/Raf/Erk pathway mediates the subarachnoid hemorrhage-induced apoptosis of hippocampal neurons through phosphorylation of p53. Mol Neurobiol. 2016; 53: 5737-5748.
  29. Bruchhaus I, Loftus BJ, Hall N, Tannich E. The intestinal protozoan parasite Entamoeba histolytica contains 20 cysteine protease genes, of which only a small subset is expressed during in vitro Eukaryot Cell. 2003; 2: 501-509.
  30. McCubrey JA, Steelman LS, Chappell WH, Abrams SL, Wong EW, Chang F, et al. Roles of the Raf/MEK/ERK pathway in cell growth, malignant transformation and drug resistance. Biochim Biophys Acta. 2007; 1773: 1263-1284.
  31. Sajid M, Mckerrow JH. Cysteine proteases of parasitic organisms. Mol Biochem Parasitol. 2002; 120: 1-21.
  32. Mazur T, Hadas E. The effect of the passages of Acanthamoeba strains through mice tissues on their virulence and its biochemical markers. Parasitol Res. 1994; 80: 431-434.
  33. Na BK, Shenai BR, Sijwali PS, Choe Y, Pandey KC, Singh A, et al. Identification and biochemical characterization of vivapains, cysteine proteases of the malaria parasite Plasmodium vivax. Biochem J. 2004; 378: 529-538.
  34. Ren M, Gao L, Wu X. TLR4: the receptor bridging Acanthamoeba challenge and intracellular inflammatory responses in human corneal cell lines. Immunol Cell Biol. 2010; 88: 529-536.
  35. Tripathi T, Abdi M, Alizadeh H. Protease-activated receptor 2 (PAR2) is upregulated by Acanthamoeba plasminogen activator (aPA) and induces proinflammatory cytokine in human corneal epithelial cells. Invest Ophthalmol Vis Sci. 2014; 55: 3912-3921.
  36. Chusattayanond AD, Boonsilp S, Kasisit J, Boonmee A, Warit S. Thai Acanthamoeba isolate (T4) induced apoptotic death in neuroblastoma cells via the Bax-mediated pathway. Parasitol Int. 2010; 59: 512-516.
  37. Zhang X, Shan P, Sasidhar M, Chupp GL, Flavell RA, Choi AMK, et al. Reactive oxygen species and extracellular signal-regulated kinase 1/2 mitogen-activated protein kinase mediate hyperoxia-induced cell death in lung epithelium. Am J Respir Cell Mol Biol. 2003; 28: 305-315.
  38. Glotin AL, Calipel A, Brossas JY, Faussat AM, Treton J, Mascarelli F. Sustained versus transient ERK1/2 signaling underlies the anti- and proapoptotic effects of oxidative stress in human RPE cells. Invest Ophthalmol Vis Sci. 2006; 47: 4614-4623.
  39. Yang R, Piperdi S, Gorlick R. Activation of the RAF/mitogen-activated protein/extracellular signal-regulated kinase kinase/extracellular signal-regulated kinase pathway mediates apoptosis induced by chelerythrine in osteosarcoma. Clin Cancer Res. 2008; 14: 6396-6404.
  40. Wu Z, Wu LJ, Tashiro S, Onodera S, Ikejima T. Phosphorylated extracellular signal-regulated kinase up-regulated p53 expression in shikonin-induced HeLa cell apoptosis. Chin Med J (Engl). 2005; 118: 671-677.
  41. Cox AD, Der CJ. Ras history: The saga continues. Small GTPases. 2010; 1: 2-27.
  42. Sonoda K, Sakamoto T, Yoshikawa H, Ashizuka S, Ohshima Y, Kishihara K, et al. Inhibition of corneal inflammation by the topical use of Ras farnesyltransferase inhibitors: selective inhibition of macrophage localization. Invest Ophthalmol Vis Sci. 1998; 39: 2245-2251.
  43. Liang Q, Mohan RR, Chen L, Wilson S. Signaling by HGF and KGF in corneal epithelial cells: Ras/MAP kinase and Jak-STAT pathways. Invest Ophthalmol Vis Sci. 1998; 39: 1329-1338.
  44. Troll W, Frenkel K, Wiesner R. Protease inhibitors as anticarcinogens. J Natl Cancer Inst. 1984; 73: 1245-1250.