Improved Therapeutic Efficacy of Mammalian Expressed-Recombinant Interferon Gamma Against Ovarian Cancer Cells

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Experimental Cell Research xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Experimental Cell Research


journal homepage: www.elsevier.com/locate/yexcr

Improved therapeutic efficacy of mammalian expressed-recombinant


interferon gamma against ovarian cancer cells
Ali Razaghia, Carina Villacrésb, Vincent Jungb, Narges Mashkoura, Michael Butlerb,

Leigh Owensa, Kirsten Heimanna,
a
Centre for Biodiscovery and Molecular Development of Therapeutics, James Cook University, Townsville QLD 4811, Australia
b
Department of Microbiology, University of Manitoba, Winnipeg, MB, Canada R3T 2N2

A R T I C L E I N F O A BS T RAC T

Keywords: Human interferon gamma (hIFNγ) affects tumour cells and modulates immune responses, showing promise as
Interferon gamma an anti-cancer biotherapeutic. This study investigated the effect of glycosylation and expression system of
Glycosylation recombinant hIFNγ in ovarian carcinoma cell lines, PEO1 and SKOV3. The efficacy of E. coli- and mammalian-
Ovarian cancer expressed hIFNγ (hIFNγ-CHO and HEK293, glycosylated/de-glycosylated) on cytostasis, cell death (MTT, and
SKOV3
Guava-ViaCount® flow-cytometry) and apoptotic signalling (Western blot of Cdk2, histone H3, procaspase-3,
FADD
Therapeutic efficacy
FADD, cleaved PARP, and caspase-3) was examined. Hydrophilic Interaction Liquid Chromatography
determined the structure of N-linked glycans present in HEK293-expressed hIFNγ (hIFNγ-HEK). PEO1 was
more sensitive to hIFNγ than SKOV3, but responses were dose-dependent and expression platform/glycosyla-
tion status-independent, whereas SKOV3 responded to mammalian-expressed hIFNγ in a dose-independent
manner, only. Complex-type oligosaccharides dominated the N-glycosylation pattern of hIFNγ-HEK with some
terminal sialylation and core fucosylation. Cleaved PARP and cleaved caspase-3 were not detected in either cell
line, but FADD was expressed in SKOV3 with levels increased following treatment. In conclusion, hIFNγ did not
induce apoptosis in either cell line. Mammalian- expressed hIFNγ increased cell death in the drug-resistant
SKOV3. The presence of FADD in SKOV3, which may inhibit apoptosis through activation of NF-κB, could serve
as a novel therapeutic target.

1. Introduction and fucose. In contrast, the glycans at N97 are non-fucosylated hybrid
high-mannose structures with a sugar composition of N-acetylneur-
Human interferon-gamma (hIFNγ) is a cytokine with immunomo- aminic acid, galactose, mannose, and N-acetylglucosamine [2,4].
dulatory properties, vital for innate and adaptive immunity against Commercial hIFNγ is bacterially derived from Escherichia coli
viral/microbial infections and exhibits cytotoxic/cytostatic activity (human interferon gamma-1b (hIFNγ-1b), tradename:
against cancer cells. Native hIFNγ is naturally synthesised by CD4+ T ACTIMMUNE®), approved for clinical treatment of chronic granulo-
helper cell type 1 (Th1) lymphocytes, CD8+ cytotoxic lymphocytes and matous disease and malignant osteopetrosis, with growing prospect for
natural killer (NK) cells [1,2]. cancer immunotherapy [2]. A short half-life stemming from a lack of
The active form of native hIFNγ is a soluble homodimer with two N- glycosylation limits the efficacy of hIFNγ-1b [5] and production is
glycosylation sites on asparagines (N25 & N97) on the surface of each costly due to the formation of inclusion bodies and endotoxin
dimer [1–3]. The glycans at N25 are fucosylated and are mainly contamination [6]. More cost-efficient production is being explored
complex-type sialylated oligosaccharides with the sugar composition in eukaryotic systems such as yeast (e.g. Pichia pastoris) [7], protozoa
of N-acetylneuraminic acid, galactose, mannose, N-acetylglucosamine, (e.g. Leishmania sp.), and mammalian cells (e.g. Chinese hamster

Abbreviations: 2-AB, 2-aminobenzoic acid; 3D, three-dimensional structure; AU, arbitrary unit; BSA, bovine serum albumin; CHO, Chinese hamster ovary; hIFNγ-CHO, CHO-
expressed hIFNγ; deglyco-hIFNγ-HEK, deglycosylated HEK293-expressed hIFNγ; delgyco-hIFNγ-CHO, deglycosylated CHO-expressed hIFNγ; FBS, foetal bovine serum; hIFNγ-
HEK, HEK293-expressed hIFNγ; hIFNγ-R, hIFNγ receptor; HEK293, human embryonic kidney 293; hIFNγ-1b, human interferon gamma-1b;; hIFNγ, human interferon-gamma;
HILIC, hydrophilic interaction liquid chromatography; IRF-1, interferon regulatory factor-1; IL1β, interleukin 1 beta; NK, natural killer cells; NF-κB, nuclear factor kappa-light-
chain-enhancer of activated B cells; PNGase F, peptide N-glycosidase F; FADD, Fas-Associated death domain; PARP, poly-ADP-ribose polymerase; RIPK3, receptor-interacting
kinase 3; STAT1, signal transducer & activator of transcription-1; TNFα, tumour necrosis factor alpha

Corresponding author.
E-mail address: Kirsten.heimann@jcu.edu.au (K. Heimann).

http://dx.doi.org/10.1016/j.yexcr.2017.08.014
Received 17 May 2017; Received in revised form 5 August 2017; Accepted 8 August 2017
0014-4827/ © 2017 Elsevier Inc. All rights reserved.

Please cite this article as: Razaghi, A., Experimental Cell Research (2017), http://dx.doi.org/10.1016/j.yexcr.2017.08.014
A. Razaghi et al. Experimental Cell Research xxx (xxxx) xxx–xxx

ovary (CHO), human embryonic kidney 293 (HEK293), mice, rat) [2], [16], increased activation of the anti-apoptotic transcription factor,
with best results to date in mammalian expression systems (higher nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB),
productivities, similarity to native hIFNγ in glycosylation and protein a biomarker for aggressive phenotypes in lymphomas, lung and color-
folding) [2]. Differences exist in the N-glycan structure of hIFNγ ectal carcinomas (Fig. 1) [17,18]. hIFNγ, through induction of the Fas
expressed in CHO; i.e. non-human-like sialylation potentially induces pathway and export of FasL [19], can activate FADD (Fig. 1), which
immunogenicity and affects stability [2,8]. may explain h-IFNγ resistance in cancer cell lines like SKOV3.
Glycan residues provide protease resistance, and the type of glycan Therefore FADD levels following treatment with hIFNγ were also
affects half-life (i.e. high mannose-type hIFNγ expressed in insect cells investigated here in PEO1 and SKOV3.
had a shorter half-life) and pharmacokinetics of hIFNγ [3]. Detailed
knowledge of the effect of glycosylation on the therapeutic efficacy of
hIFNγ is limited, although glycosylation has been proven to be 2. Material & methods
essential for the therapeutic efficacy of protein drugs [9]. This study
explored the efficacy of glycosylation status of recombinant hIFNγ from 2.1. Ovarian carcinoma cell lines & cultivation
three different expression systems (E. coli, CHO and HEK293) against
the ovarian cancer cell lines, PEO1 and SKOV3, to evaluate whether or Two ovarian carcinoma cell lines, PEO1 (passage № ≥ 40)
not mammalian expressed recombinant hIFNγ can be a superior (Catalogue № 10032308) and SKOV3 (passage № ≥ 20) (Catalogue
substitute for the prokaryotic product. № 91091004), were purchased from the European Collection of
Ovarian cancer is the 5th deadliest cancer in women and treatment Authenticated Cell Cultures, UK. Cell lines were maintained at 37 °C
of ovarian cancer with hIFNγ-1b underwent phase I, II & III clinical in a humidified growth chamber in 5% CO2 in RPMI-1640 medium
trials with mixed results. Some of the identified obstacles were tumour supplemented with L-glutamine and sodium bicarbonate (Sigma,
insensitivity to hIFNγ and inability to deliver hIFNγ locally [10–12]. In R8758, Castle Hill, NSW 1765, Australia) and addition of 10% foetal
contrast, preclinical in vitro & in vivo studies on ovarian cancer cell bovine serum (FBS) (Sigma, F4135; USA origin); penicillins
lines all showed a degree of sensitivity to hIFNγ-1b except for SKOV3 (100 U mL−1) and streptomycin (100 µg mL−1). Cells were subcultured
(Table 1). when confluent. For the passage, cells were detached through addition
Anti-cancer efficacy of hIFNγ in ovarian cancer cells was shown to of Gibco® Trypsin-EDTA (0.25%) solution (ThermoFisher Scientific,
be due to cytostasis, through activation of p53 and p21 leading to cell Newstead, QLD 4006, Australia). Subsequently, trypsin was deacti-
cycle arrest (Fig. 1) and/or cytotoxicity, causing cell death. The latter vated by addition of FBS-containing complete medium. Subcultures
was subdivided into pyroptosis through activation of interferon reg- were seeded with 50 ×103 cells per vented T-25 flask (25 cm2, Orange
ulatory factor-1 (IRF-1) and caspase-1, and intrinsic apoptotic signal- Scientific, Sydney, NSW, Australia). Culture medium was changed
ling via activation of IRF-1, caspase-8, cytochrome-c release from every three days.
mitochondria, the caspase cascade and inhibition of poly-ADP-ribose
polymerase (PARP) (Fig. 1; Table 1). Treatment of OVCAR3 with
hIFNγ-1b led to both, apoptosis and cytostasis. Prior work on PEO1 2.2. Recombinant hIFNγ
also showed the possible engagement of both, cytostasis and apoptosis
[13,14], which had been suggested based on signal transducer & Products from three different protein expression platforms were
activator of transcription-1 (STAT1) activation and p53 expression purchased; hIFNγ-1b from Sigma (SRP3058), CHO-expressed (hIFNγ-
[15]. Therefore, this research aimed to investigate other signalling CHO) from SinoBiologicals (11725-HNAS, Beijing, China) (Purity: >
molecules (cleaved PARP and caspase-3) to provide more conclusive 92%) and HEK293-expressed (hIFNγ-HEK) from Acrobiosystems
evidence of hIFNγ-induced apoptosis. In contrast, high levels of (IFG-H4211, Newark, DE 19711, USA). All recombinant hIFNγ
phosphorylated Fas-Associated Death Domain (FADD), a regulator of products were provided as a lyophilized powder from PBS (pH 7.4)
cell cycle progression, proliferation, tumorigenesis and necroptosis with a purity > 92%, and were tested for endotoxin content (< 0.1
EU μg−1).

Table 1
Summary of preclinical treatments of ovarian cancer cell lines with hIFNγ-1b.

Cell Line Mechanism of Action Signalling Molecules Detected Ref.

2774 Apoptosis Nitric oxidea [40]


Pyroptosis Induction of IRF-1 & activation of caspase-1 [41]
HOC7 Apoptosis Nitric oxidea [40]
OAW42 Apoptosis PARP inhibition [14]
OVCAR3 Apoptosis Nitric oxidea [40]
Less sensitive Induction of IRF-1 [41]
Apoptosis PARP inhibition [14]
Cytostasis/Apoptosis Cell cycle arrest; G1, G2, and S phases; [42]
Induction of p53, Bax, and caspase-3b
Cytostasis Increasing p21 mRNA [15]
OVCAR4, OVCAR5 Cytostasis/Apoptosis PARP inhibition [14]
PA1 Pyroptosis Induction of IRF-1 & activation of caspase-1 [41]
PEO1 Apoptosis/Cytostasis Depolarization of mitochondrial membrane, release of cytochrome C & activation of caspase-9 [13]
Induction of caspase-8 & 9 [14]
Cytostasis Increasing STAT1 & [15]
p21 mRNA
PEO14, PEO16 Cytostasis/Apoptosis PARP inhibition [15]
SKOV3 Insensitive Initial expression of p21, IRF-1 mRNA detected. However, later the pattern of expression was reduced to the [14,15,41]
level of untreated cells.
SW626 Cytostasis/Apoptosis PARP inhibition [14]

a
Polytherapy of hIFNγ, interleukin 1 beta (IL1β), and tumour necrosis factor alpha (TNFα).
b
Polytherapy of hIFNγ plus TNFα.

2
A. Razaghi et al. Experimental Cell Research xxx (xxxx) xxx–xxx

Fig. 1. hIFnγ-induced signal transduction in ovarian carcinoma cells. Involvement of the FADD pathway is hypothetical. The figure has been composed based on information obtained
from [13–15,17–19,25,29,30,34,36,41,43–47]. Bax, Bcl-2-associated X protein, Bid, BH3 interacting domain death agonist, CASP1, 3, 7, 8, 9, Caspase1, 3, 7, 8, 9; CYT-C,
Cytochrome-C; c-FLIP, Cellular FLICE (FADD-like IL-1β-converting enzyme)-inhibitory protein; FADD, Fas-Associated Death Domain Protein; Fas, Cell surface death receptor;
FasL, Fas Ligand; GAS, Gamma interferon-activated sequence; IRF-1, Interferon-regulated factor-1; Jak, Janus kinase; NF-κB, Nuclear factor kappa-light-chain-enhancer of
activated B cells; PARP, Poly-(ADP-ribose) polymerase; STAT1, Signal transducer & activator of transcription-1; TRAIL, TNF-related apoptosis-inducing ligand; DR, Death receptor.

2.3. Deglycosylation leading to the determination of total cell counts. ii) The “viability dye”
stains dead cells.
N-glycans are positioned on the surface of the hIFNγ dimmer [1,3] The cytotoxic effect was determined as a percentage based on the
and are thus readily available targets for enzymatic removal. hIFNγ- division of total dead cells to estimated absolute cell number. The
CHO and hIFNγ-HEK were deglycosylated (delgyco-hIFNγ-CHO; percentile of the cytostatic effect of each treatment was calculated as
deglyco-hIFNγ-HEK293) using the GlycoProfile™ II Enzymatic In- per equation 1 (Eq. (1)):
solution N-Deglycosylation Kit (Sigma, PP0201) following the manu-
(Control[total cells]−Treatment[total cells])
facturer's instructions with the following modification: β-mercap- Cytostasis(%) = ×100
toethanol, a denaturant, was omitted due to potential cytotoxic effects Control[total cells] (1)
on cancer cells and heat-inactivation (100 °C) of hIFNγ, as temperature
of 40–64 °C inactivates hIFNγ.
2.5.2. Cell cycle analysis
2.4. In-vitro treatments The percentile of cells in G0/G1, S, and G2 phases was determined
based on DNA content. Before treatment, cells were synchronised in G0
50 ×103 cells were inoculated per T25-flask (n = 3) and incubated by culturing for 24 h in RPMI medium without FBS. Following
for 24 h. 5 mL of complete RPMI-1640 medium supplemented with treatment, the Guava® Cell Cycle Reagent (Merck, 4500-0220) was
100 ng mL−1 of each six treatment conditions (hIFNγ-1b, hIFNγ-CHO, used following the manufacturer's protocol.
deglyco-hIFNγ-CHO, hIFNγ-HEK and deglyco-hIFNγ-HEK) and in-
cubated for 72 h. Cells cultured in the same medium but in the absence
of recombinant hIFNγs served as negative controls. 2.6. Protein extraction, determination & analysis

2.5. Cytotoxic & cytostatic measurements Following detachment at confluency after 72 h treatment, cells were
washed with PBS then lysed by adding 0.5 mL ice-cold lysis buffer
A Guava® easyCyte 8HT Benchtop Flow Cytometer (Merck, 0500– (10 mM Tris-HCl, 100 mM NaCl, pH 8, 25 mM EDTA, pH 8, 0.5% SDS,
4008) was used for conducting assays as follows; 1 Protease inhibitor cocktail tablet (Boehringer, 1697498, Mannheim,
Germany) per 10 mL) and vigorous pipetting. Afterwards, lysates were
2.5.1. Cell counting & viability frozen at −20 °C overnight, before thawing and centrifugation at
Absolute cell counts, identification of dead cells, and viability were 12,000g for 5 min at 4 °C. The total protein content of the cell lysate
determined using the Guava® ViaCount® Reagent (Merck, 4000-0040, supernatant was determined spectrophotometrically (Enspire®
USA) on cell suspensions. This assay distinguishes viable cells from Multimode Plate Reader (PerkinElmer, Glen Waverly, VIC,
dead ones based on the differential permeability of cell membranes to Australia)) using the Pierce™ BCA Protein Assay Kit (ThermoFisher,
the DNA-binding dyes used. i) The “nuclear dye” stains nucleated cells 23225) using bovine serum albumin (BSA) as a standard.

3
A. Razaghi et al. Experimental Cell Research xxx (xxxx) xxx–xxx

2.6.1. SDS-PAGE & Western blot analysis SDS-PAGE gels (Biorad, Hercules, USA) stained with Coomassie
To denature samples; 20 µg of total protein was added to Laemmli brilliant blue. Protein bands were finally chopped, washed and treated
sample buffer (200 mM Tris-Cl, (pH 6.8), 400 mM dithiothreitol, 8% overnight with peptide N-glycosidase F (PNGase F) (Roche
SDS, 0.4% bromophenol blue, 40% glycerol) followed by heating at Diagnostics, Indianapolis, USA) to release glycans. The released N-
95 °C for 5 min. Samples were loaded on 12% Mini-PROTEAN® TGX™ glycans were labelled with 2-aminobenzoic acid (2-AB).
Precast Gels (Biorad, 4561044, Gladesville, NSW 2111, Australia).
Electrophoresis was conducted at 200 V for 30 min using 1X running 2.9.1. HPLC profiling of 2-AB- labelled N-glycans
buffer (25 mM Tris, 192 mM glycine, and 0.1% SDS). Precision Plus Previously 2-AB-labelled N-glycans were analysed by hydrophilic
Protein™ Standard (Biorad, 1610373) was used to determine product interaction liquid chromatography (HILIC) with fluorescence detection
size. Proteins were transferred to PVDF membranes using Trans-Blot® using an amide HPLC Waters X-Bridge column (4.6 × 250 mm;
Turbo™ Mini PVDF Transfer Packs (Biorad, 1704156) and a Trans- 3.5 µm) and a constant flow rate of 0.86 mL min−1. The buffers used
Blot® Turbo™ Transfer System (Biorad) using the manufacturer's 3- were 50 mM ammonium formate (pH 4.4, buffer A) and acetonitrile
min protocol. Blots were incubated in blocking buffer (5% BSA in Tris- (ACN, buffer B) with a linear gradient of 20 – 50% applied over 48 min,
buffered saline with Tween 20 (TBST) buffer; 20 mM Tris base, pH 7.5, followed by an increase from 50% to 100%. Glycan structures were
150 mM NaCl, 0.1% Tween 20) at room temperature with agitation for assigned using Empower Software (Waters™, USA) where GU values
1 h. Antibodies (see below) were diluted in blocking buffer. Blot were obtained by using a 2-AB labelled dextran ladder standard and
membranes were exposed to the primary antibody at 4 °C overnight compared with NIBRT glycan database (https://glycobase.nibrt.ie).
and to the secondary antibody with agitation at room temperature for
1 h. Clarity Max™ Western ECL Substrate (Biorad, 1705062) and the 2.10. Statistics
Syngene G: Box XRQ chemiluminescent system were used for detection
and imaging, respectively. Data (n = 3 replicates) were analysed via one-way ANOVA using S-
Densitometry using the Gel Analyser function in ImageJ (version PLUS® software, version 8.0 with α set to 0.01. Homogeneity of
1.51i, USA) was used to quantitate relative intensity, as per equation 2 variance and normality was ascertained using the Leven's and the
(Eq. (2)): Shapiro-Wilks tests, respectively. Tukey-Kramer HSD posthoc tests
were performed to determine the source of significance at p < 0.01.
Relative intensity
Blot of interest [arbitrary unit ] 3. Results
=
Corresponding blot of the loading control (β − actin )[arbitrary unit ]
(2) 3.1. Effects of recombinant hIFNγ on PEO1

Treatments with all recombinant hIFNγ products caused 60–65%


2.6.2. Antibodies
dead cells through cytotoxicity as determined by ViaCount® (Fig. 2A).
Apoptosis Western Blot Cocktail kit (pro/p17-caspase 3, cleaved-
Statistical analyses confirmed a significant effect of treatment com-
PARP, β-actin) (Abcam, ab136812, Melbourne, VIC, Australia) was
pared to controls, but there was no statistically significant effect (p <
used with dilution for primary antibody cocktail 1:250 and 1:100 for
0.01) between treatments (i.e. expression source or glycosylation
goat HRP-conjugated secondary antibody cocktail. Cell Cycle and
status) (Fig. 2A). Neither cleaved PARP nor cleaved caspase-3 was
Apoptosis WB Cocktail kit (Cdk2 pTyr15 (pCdk), Histone H3 pSer10
detected in Western blots, however, a slight reduction in relative
(pHH3) /β-actin/PARP) (Abcam, ab139417) was used with dilution for
intensity of procaspase-3 expression (an inactive zymogen of caspase
primary antibody cocktail 1:250 and 1:2500 for secondary antibody
3) was observed after treatment with all recombinant hIFNγ products,
cocktail. Anti-FADD antibody (Abcam, ab24533) was used with dilu-
as quantified by Image-J as a ratio of 0.9–1.1, compared to a ratio of
tion for primary antibody 1:1000 and 1:100 for goat HRP-conjugated
1.3 in controls (Fig. 3D). Expression of FADD was not detected in
secondary antibody.
PEO1 cells.
Based on the number of PEO1 cells, all recombinant hIFNγ
2.7. Dose-response assay products caused 54–57% of cytostasis with no significant difference
between treatments; except, a slightly stronger and statistically sig-
Dose-response was evaluated using the CellTiter 96® Non- nificant effect ~62% (p < 0.01) observed in cells treated with hIFNγ-
Radioactive Cell Proliferation (MTT) Assay (Promega, G4000, HEK (Fig. 2B).
Alexandria, NSW 2015, Australia). Approximately 3000 cells were Treatment with all recombinant hIFNγ products, irrespective of
inoculated per 100 μL volume of complete RPMI-1640 medium per source and glycosylation, affected the cell cycle of PEO1 shown by an
well of a 96-well assay plate in dilution series of hIFNγ-1b and hIFNγ- increase of cells in G2 phase, 28–30%, compared to controls, ~22% (p
HEK (0–500 ng mL−1). Following the manufacturer's protocol, absor- < 0.01) (Fig. 2C). Western blot analysis confirmed reduced relative
bance was recorded spectrophotometrically (EnSpire®). intensities of Cdk2 expression, as quantified by Image-J as a ratio of
0.5–0.7, compared to a ratio of 1.1 in controls (Fig. 3E). Relative
2.8. Microscopy intensities of Histone H3, the M-phase indicator, were comparable
between controls and treatments (Fig. 3E).
Cultured cells were imaged with an inverted digital microscope The growth of PEO1 was strongly inhibited in a dose-dependent but
(Olympus IX53, Japan) in T25 flasks after 72 h of incubation with source-independent manner following treatment with HIFNγ-1b and
hIFNγ and untreated controls using phase contrast at 100× magnifica- hIFNγ-HEK (Fig. 2E). The calculated IC50 was ~200 ng mL−1 for both
tion. (Fig. 2E).

2.9. Glycan analysis 3.2. Effects of recombinant hIFNγ on SKOV3

The structure of N-linked glycans was determined following Royle, Effect of treatment on induction of cell death was significant (p <
Radcliffe, Dwek and Rudd [20]. The hIFNγ-HEK sample was recon- 0.01) with cytotoxic efficacy declining: hIFNγCHO (~28%) =
stituted in 300 μL of Milli-Q water and denatured with 15 μL of beta- hIFNγHEK (~24%) = deglycol-hIFNγHEK (~23%) > deglycol-
mercaptoethanol for 5 min at 95 °C for analysis by gradient (8–16%) hIFNγCHO (~17%) ≥ hIFNγ1b (~11%) ≈ control (~5%) (Fig. 2A).

4
A. Razaghi et al. Experimental Cell Research xxx (xxxx) xxx–xxx

Fig. 2. 72 h-treatment effects of recombinant hIFNγ from three different expression systems and their deglycosylated forms on PEO1 and SKOV3. (A) Cytotoxic effect (B) cytostatic
effect (calculated as per Eq. (1)). Cell cycle analysis of (C) PEO1 and (D) SKOV3. 48 h-dose-response to treatment with hIFNγ-1b and hIFNγ-HEK on the growth of (E) PEO1 and (F)
SOKV3. For dose-response effect; growth is expressed as a fraction of control values (Mean ± SD, n = 3), p < 0.01.

Procaspase-3 was detected, but treatments had no effect on relative Cell cycle analysis of SKOV 3 cells showed a significant percent of
expression intensity (Fig. 3A), cleaved PARP and caspase-3 were not cells arrested in S-phase following treatment with all recombinant
detected. Western blotting detected FADD expression in all treatments hIFNγ products, 68–70%, compared to controls (~31%) (Fig. 2D) (p <
and controls. Notably, relative expression was increased after treat- 0.01). Source of hIFNγ or glycosylation status had no significant
ment, as quantified by Image-J as a ratio of 0.9–2.1, compared to a impact. The largest cytostatic effects were observed for hIFNγ-
ratio of 0.7 in controls; with highest levels being observed after HEK293 (glycosylated ~43% and deglycosylated ~44%), hIFNγ-CHO
treatment with hIFNγ-CHO (ratio of 2.1) (Fig. 3C). (glycosylated ~44%), deglyco-hIFNγ-CHO (~31%) was less effective,

5
A. Razaghi et al. Experimental Cell Research xxx (xxxx) xxx–xxx

Fig. 3. Western blot analysis of recombinant hIFNγ-induced signalling molecules in SKOV3 and PEO1. (A, D) procaspase-3, (C) FADD, (B, E) Cdk2 (as an indication of G1/S phase),
Histone H3 (as a biomarker of M phase), Untreated cells (controls) were used to determine un-induced signalling molecule levels and β-actin, a housekeeping protein, was used as a
loading control to obtain relative intensity histograms with Image-J (Eq. (2)). * C had the same loading controls as B.

and hIFNγ-1b (~24%) was the least effective (Fig. 2B) (p < 0.01). hIFNγ could not be calculated for this cell line.
Western blot analysis confirmed an increase in relative expression of Furthermore, SKOV3 cells exposed to hIFNγ-CHO, deglyco-hIFNγ-
Cdk2 after treatment, as quantified by Image-J as a ratio of 1–1.6, CHO, hIFNγ-HEK, and deglyco-hIFNγ-HEK showed changes in cell
compared to a ratio of 0.7 in controls. In contrast, relative expression of morphology (Fig. 4), with cells becoming thin and elongated more
Histone H3 was reduced in treatments, as quantified by Image-J as a frequently, while such alterations were not observed for PEO1.
ratio of 0.2–0.4 compared to a ratio of 0.6 in controls (Fig. 3B).
The growth of SKOV3 was inhibited in a dose-independent but
source-dependent manner following treatment with HIFNγ-1b and 3.3. Glycan analysis
hIFNγ-HEK (Fig. 2F). Treatment with 15–500 ng mL−1 inhibited
growth by approximately 15% and 25% for hIFNγ-1b and hIFNγ- The glycan analysis of hIFNγ-HEK showed that the protein is
HEK, respectively. Since no dose response was observed, an IC50 of heterogeneously glycosylated containing bi-, tri-, and tetra-antennary
structures dominated by complex-type oligosaccharides, many of which

6
A. Razaghi et al. Experimental Cell Research xxx (xxxx) xxx–xxx

Fig. 4. Recombinant hIFNγ-induced cell elongation in SKOV3 cells. Phase contrast micrographs of SKOV3 cells with elongated thin shapes after 72 h treatment. (A) Control (B)
hIFNγ-1b (C) hIFNγ-CHO (D) deglyco-hIFNγ-CHO (E) hIFNγ-HEK (F) deglyco-hIFNγ-HEK. White arrows point to elongated thin-shaped cells.

Fig. 5. NP-HPLC profile of 2-AB-labelled hIFNγ-HEK N-glycans.

contained terminal sialylation and core fucosylation. Mannose, galac- 4. Discussion


tose, fucose and sialic acid were major constituents of the sugar
composition (Fig. 5). High heterogeneity in cancer cells regarding genomic mutations,
DNA copy number, oncogene bypass, and epigenetic alterations is the

7
A. Razaghi et al. Experimental Cell Research xxx (xxxx) xxx–xxx

leading cause of inconsistent responses to chemotherapy and drug- 4.2. hIFNγ efficacy in SKOV3
resistance [21]. This can explain the observed differences in cytotoxi-
city, cytostasis, IC50 and therapeutic efficacy of the recombinant SKOV3 has been reported as insensitive to treatment with hIFNγ-
hIFNγs used here in PEO1 and SKOV3. 1b [14]. However, this study showed a degree of sensitivity to
mammalian expressed hIFNγ. Deglycosylation had a marginal impact
on the cytotoxic efficacy but a significant impact on the cytostatic
efficacy in CHO-, but not HEK-expressed hIFNγ. This suggests that
4.1. hIFNγ efficacy in PEO1 other factors like the proper folding of the native protein in mammalian
expression systems are as important as glycosylation for drug potency
PEO1 was highly susceptible to treatment with hIFNγ regardless of and pharmacodynamics [28].
expression platform and glycosylation. Although key apoptotic signal- Signalling molecules typically involved in apoptosis (cleaved PARP
ling molecules (cleaved PARP and caspase-3) were not detected by and caspase3) were not detected by Western blot analysis, ruling out
Western blot, levels of procaspase-3 decreased slightly. Caspase-3 is an apoptosis as the principal mechanism of cell death. Although FADD
executioner of apoptosis through the proteolytic processing of down- was present in SKOV3 at detectable levels in controls and treatments,
stream proteins such as PARP. Caspase-3 (active p17 and p12 FADD is unlikely an initiator of the extrinsic apoptotic pathway, as no
subunits) is activated through cleavage of procaspase-3, which is a cleaved PARP and caspase-3 were detected. Overall cell death was low
zymogen (inactive precursor). Thus, monitoring the altered proportion in SKOV3, which together with the presence of FADD in control cells,
of procaspase-3 to activated/cleaved caspase-3, i.e. a decrease of the may indicate that FADD inhibits apoptosis through the activation of
procaspase-3 or an increase of caspase 3, can be used as a biomarker NF-κB (Fig. 1) [17] and autophagy [29]. FADD-induced necroptotic
for apoptosis [22]. These results seem to oppose apoptotic signal cell death via mitochondrial dysfunction requires further investigation
transduction as a consequence of hIFNγ treatment in PEO1, contrary on increased levels of proteins participating in the formation of
to previous reports based on flow-cytometric measurements of induced complex II (Fig. 1) [25]. Of all cell death mechanisms, necroptosis is
levels of caspase 8 and 9 and Western blot detection of cleaved PARP, the most probable in SKOV 3, an interpretation supported by hIFNγ-
caspase-8 and 9 [13,14]. Some differences in cell numbers (not given induced cell death mechanisms in other cancer types [16,30], but direct
vs. 10,000 cells mL−1), hIFNγ concentrations used (≈ 250 vs. evidence via detection of the core signalling molecule, receptor-
100 ng mL−1), treatment duration (96 vs. 72 h) and particularly interacting kinase 3 (RIPK3) [25], is required.
confluence status of the culture (100% vs. 40% used here) and passage FADD levels in SKOV3 cells could potentially correlate with
number between their and our study limits comparability. Apoptosis in tumorigenesis, anti-apoptotic behaviour, and resistance against antic-
in vitro culture conditions has been reported to be strongly cell density- ancer drugs, which has also been reported for lymphomas, lung and
dependent e.g. high cell density induced apoptosis in an autocrine colorectal carcinomas [18]. This could make FADD a novel target to
manner following the release of an unknown low molecular weight overcome drug-resistance in this ovarian cancer cell line (Fig. 1) [17].
peptide-containing factor in leukaemia HL-60 cells [23]. In contrast, a In SKOV3, all three types of hIFNγ caused cytostasis through cell cycle
high passage number has been implicated in causing resistance to arrest in S phase accompanied by an increase in Cdk2 and a decrease in
apoptosis [24], which is a possible explanation why apoptosis was not Histone 3. This result is supported by previous studies showing hIFNγ-1b-
observed in this study. Since cytostatic effects and mechanisms of cell induced elevated levels of p21 - and IRF-1 mRNA in SKOV3 [15]. Other
death of hIFNγ were not comparable with previous publications on cancer types sensitive to hIFNγ-1b also showed a similar trend, e.g. the
PEO1, there is an urgent need for standardisation of toxicity assays induction of cell cycle arrest in human mesothelioma with hIFNγ-1b
(e.g. confluency, passage number, duration of treatment and use of seems to depend on cyclin regulation through p21WAF1/CIP1 and p27Kip1-
bioactivity units), especially when aiming to unravel signalling cascades independent mechanisms [31] and relies on p53 and p21WAF1/clp1/Sdl1
and treatment-induced cell death mechanisms. pathways in hepatocytes [32] (Fig. 1).
Necroptotic cell death in PEO1 appears unlikely since activation of
FADD is essential for induction of necroptosis [25]. Autophagy remains 4.3. Differences in responses of PEO1 & SKOV3 to treatment with
a possibility for an hIFNγ-induced cell death mechanism in this high hIFNγ
passage number PEO1 cell line, as hIFNγ has been shown to induce
autophagy signalling pathways with growth inhibition and cell death in In general, comparison of cytostasis following treatment with
other cancer cell lines like hepatocellular carcinoma [26]. Thus, it is recombinant hIFNγ showed that PEO1 was 10–30% more sensitive
recommended to investigate the presence of key autophagic cell death than SKOV3 (Fig. 2B). Furthermore, cytotoxicity of hIFNγ was greater
proteins e.g. LC3 (microtubule-associated protein 1 light chain 3), (~70% dead cells) compared to SKOV3 (max. ~30% dead cells)
Beclin-1 or Atg5 (Autophagy protein 5) proteins [26]. (Fig. 2A).
Treatment with recombinant hIFNγ led to cytostasis caused by cell Treatment with mammalian-expressed hIFNγ induced higher levels
cycle arrest in G2 phase, correlating well with previous reports on of cell death compared to untreated controls and hIFNγ-1b treatment
hIFNγ-1b-induced elevated levels of p21 mRNA and cytostasis [15]. only in SKOV3. While glycosylation can affect the three-dimensional
Decreased levels of the Cdk2 protein in treatments correlate with lower structure (3D) and binding affinity of glycoproteins in general [33],
cell numbers in G1/S phases in comparison to controls [27]. Low deglycosylation possibly did not influence the 3D structure of hIFNγ,
histone H3 levels in controls and treatments may indicate a very short since the N-glycosylation sites are on the external surface of the
M phase. homodimer, explaining the no difference in efficacy in treatments with
hIFNγ-1b dose response experiments on growth and cell death in glycosylated and deglycosylated hIFNγ-HEK in SKOV3. In contrast,
ovarian cancers show high variability depending on cell lines used [14]. deglycosylation of hIFNγ-CHO decreased cell death. This contrasts
For example, reported IC50s were 10 units mL−1 ( = 0.5 ng mL−1) and with the observation of hIFNγ-induced cell death in PEO1, where
> 5000 units mL−1 (> 250 ng mL−1) for OVCAR4 and SW626, respec- source and glycosylation did not result in different outcomes. This
tively. This study is the first examining dose-responses for SKOV3 and seems to indicate that the hIFNγ receptor (hIFNγ-R) in PEO1 and
PEO1, which were dose-independent and dose-dependent, respec- SKOV3 may be different in their ability to bind hIFNγ. The SKOV3
tively. Comparisons of IC50s between our study and previous studies hIFNγ-R could bind de-glycosylated/glycosylated hIFNγ better than
is hampered due to significant differences in experimental conditions bacterial unglycosylated hIFNγ-1b, if the latter has a different 3D
(e.g. cell density, the difference in cell lines, and duration of treatment, structure, while the hIFNγ-R interaction in PEO1 cells could be more
see above). tolerant of differences in 3D structure. This hypothesis is supported by

8
A. Razaghi et al. Experimental Cell Research xxx (xxxx) xxx–xxx

studies that show mutations in hIFNγ-Rs affect binding to the receptor Acknowledgment
[19,34–36]. However, it may be possible that deglycosylation efficiency
could have been different for the two sources of hIFNγ, which could be The authors wish to acknowledge the provision of a James Cook
validated via SDS-PAGE in future studies. Future detailed studies on University Postgraduate Research Scholarship (JCUPRS) to Ali
amino acid sequence differences in the receptors and crystallography to Razaghi. The study was financed by the Higher Degree Research
unravel changes in the 3D structure of hIFNγ forms and binding sites Enhancement Scheme (HDR ES) grant allocated to AR. We acknowl-
of the receptor are also required. Furthermore, induction of FADD in edge A/Prof. Patrick Schaeffer for his assistance with protein analysis.
SKOV3, but not PEO1, could indicate hIFNγ-initiated Fas-signalling in
general with increasing efficacy by the mammalian forms, especially the References
CHO-derived glycosylated form of hIFNγ. This conclusion is supported
by regulation of Fas-ligand expression via expression of hIFNγ in [1] Ha Young, K.J. Hardy, Role of interferon-y in immune cell regulation, J. Leukoc.
human CD4+ T-lymphocytes [19]. Biol. 58 (1995) 373–381.
[2] A. Razaghi, L. Owens, K. Heimann, Review of the recombinant human interferon
gamma as an immunotherapeutic: impacts of production platforms and glycosy-
4.4. Glycan analysis lation, J. Biotechnol. 240 (2016) 48–60.
[3] T. Sareneva, K. Cantell, L. Pyhala, J. Pirhonen, I. Julkunen, Effect of carbohydrates
on the pharmacokinetics of human interferon-gamma, J. Interferon Res. 13 (1993)
In CHO, N-glycosylation of hIFNγ has been reported to carry 267–269.
terminal Gal-α1-3-Gal and to be highly sialylated [8]. Additionally, [4] T. Sareneva, J. Pirhonen, K. Cantell, I. Julkunen, N-glycosylation of human
interferon-y: glycans at Asn-25 are critical for protease resistance, Biochem. J. 308
the composition of the culture medium and production platform have (1995) 9–14.
been shown to influence the heterogeneity of N-glycosylation (of both [5] N. Miyakawa, M. Nishikawa, Y. Takahashi, M. Ando, M. Misaka, Y. Watanabe,
N-, one N-site and non-glycosylated) [37,38]. While the N-glycan Y. Takakura, Prolonged circulation half-life of interferon gamma activity by gene
delivery of interferon gamma-serum albumin fusion protein in mice, J. Pharm. Sci.
oligosaccharide structure has been examined in hIFNγ-CHO, reporting
100 (2011) 2350–2357.
high levels of sialylation [39], our study is the first to report on the N- [6] A. Razaghi, E. Tan, L.H. Lua, L. Owens, O.P. Karthikeyan, K. Heimann, Is Pichia
glycan structure in hIFNγ-HEK, which validates earlier reports on the pastoris a realistic platform for industrial production of recombinant human
absence of terminal Gal-α1-3-Gal and low levels of sialylation in interferon gamma?, Biologicals 45 (2017) 52–60.
[7] A. Razaghi, R. Huerlimann, L. Owens, K. Heimann, Increased expression and
human-expressed hIFNγ. secretion of recombinant hIFNγ through amino acid starvation-induced selective
In summary, hIFNγ is currently undergoing clinical trials, however pressure on the adjacent HIS4 gene in Pichia pastoris, Eur. Pharm. J. 62 (2015)
knowledge on how different sources of recombinant hIFNγ, with diver- 43–50.
[8] D. Ghaderi, M. Zhang, N. Hurtado-Ziola, A. Varki, Production platforms for
gent glycosylation profiles, affect cell proliferation, death and signalling is biotherapeutic glycoproteins. Occurrence, impact, and challenges of non-human
limited. This study also investigated effects of differences in glycosylation sialylation, Biotechnol. Genet. Eng. Rev. 28 (2012) 147–176.
on the efficacy of recombinant hIFNγ by comparing the efficacy of [9] R.J. Sola, K. Griebenow, Glycosylation of therapeutic proteins: an effective strategy
to optimize efficacy, BioDrugs 24 (2010) 9–21.
recombinant hIFNγ from two mammalian expression systems (CHO [10] G.P. Dunn, C.M. Koebel, R.D. Schreiber, Interferons, immunity and cancer
and HEK293) in their glycosylated and deglycosylated forms against immunoediting, Nat. Rev. Immunol. 6 (2006) 836–848.
hIFNγ-1b for the potential treatment of ovarian cancers, using a sensitive [11] G.H. Windbichler, H. Hausmaninger, W. Stummvoll, A.H. Graf, C. Kainz,
J. Lahodny, U. Denison, E. Muller-Holzner, C. Marth, Interferon-gamma in the
cell line (PEO1) and a resistant cell line (SKOV3). As all recombinant first-line therapy of ovarian cancer: a randomized phase III trial, Br. J. Cancer 82
hIFNγ products had a purity > 92%, and were tested for endotoxin (2000) 1138–1144.
content (< 0.1 EU μg−1), similar bioactivities were assumed; however, a [12] D.S. Alberts, C. Marth, R.D. Alvarez, G. Johnson, M. Bidzinski, D.R. Kardatzke,
W.Z. Bradford, J. Loutit, D.H. Kirn, M.C. Clouser, M. Markman,
contribution to the observed effects due vendor-specific differences in
G.C.T. Consortium, Randomized phase 3 trial of interferon gamma-1b plus
purification methods cannot be completely ruled out. standard carboplatin/paclitaxel versus carboplatin/paclitaxel alone for first-line
treatment of advanced ovarian and primary peritoneal carcinomas: results from a
prospectively designed analysis of progression-free survival, Gynecol. Oncol. 109
5. Conclusion (2008) 174–181.
[13] C. Barton, D. Davies, F. Balkwill, F. Burke, Involvement of both intrinsic and
Through investigation of potential apoptotic signalling pathways, a extrinsic pathways in IFN-gamma-induced apoptosis that are enhanced with
cisplatin, Eur. J. Cancer 41 (2005) 1474–1486.
new potential drug target was discovered in SKOV3, FADD, being [14] L. Wall, F. Burke, C. Barton, J. Smyth, F. Balkwill, IFN-gamma induces apoptosis in
expressed in SKOV3 but not in PEO1. Fas-associated signalling could ovarian cancer cells in vivo and in vitro, Clin. Cancer Res. 9 (2003) 2487–2496.
be responsible for treatment-resistance of this form of ovarian cancer. [15] F. Burke, P.D. Smith, M.R. Crompton, C. Upton, F.R. Balkwill, Cytotoxic response
of ovarian cancer cell lines to IFN-gamma is associated with sustained induction of
Overall, based on cytostatic and cytotoxic results achieved in SKOV3, IRF-1 and p21 mRNA, Br. J. Cancer 80 (1999) 1236–1244.
hIFNγ expressed in HEK293, but not E. coli shows promise for the [16] L. Tourneur, G. Chiocchia, FADD: a regulator of life and death, Trends Immunol.
treatment of tumorigenic and anti-cancer drug-resistant ovarian can- 31 (2010) 260–269.
[17] K.A. Schinske, S. Nyati, A.P. Khan, T.M. Williams, T.D. Johnson, B.D. Ross,
cers. Based on these promising results, we suggest that a continuum of R.P. Tomas, A. Rehemtulla, A novel kinase inhibitor of FADD phosphorylation
preclinical research is undertaken to investigate responsivity of more chemosensitizes through the inhibition of NF-kappaB, Mol. Cancer Ther. 10 (2011)
ovarian cancer cell lines, other cancer types and recombinant glyco- 1807–1817.
[18] S. Patel, D. Murphy, E. Haralambieva, Z.A. Abdulla, K.K. Wong, H. Chen, E. Gould,
proteins to generalise the idea that mammalian/human expression
G. Roncador, C. Hatton, A.P. Anderson, A.H. Banham, K. Pulford, Increased
platforms can enhance the therapeutic efficacy of anti-cancer biother- expression of phosphorylated FADD in anaplastic large cell and other T-Cell
apeutics. lymphomas, Biomark. Insights 9 (2014) 77–84.
[19] D. Boselli, G. Losana, P. Bernabei, D. Bosisio, P. Drysdale, R. Kiessling, J.S. Gaston,
D. Lammas, J.L. Casanova, D.S. Kumararatne, F. Novelli, IFN-gamma regulates Fas
Conflict of interest ligand expression in human CD4+ T lymphocytes and controls their anti-myco-
bacterial cytotoxic functions, Eur. J. Immunol. 37 (2007) 2196–2204.
[20] L. Royle, C.M. Radcliffe, R.A. Dwek, P.M. Rudd, Detailed structural analysis of N-
All authors declare no conflict of interest. The funders had no role glycans released from glycoproteins in SDS-PAGE gel bands using HPLC combined
in study design, data collection and analysis, decision to publish or with exoglycosidase array digestions, Methods Mol. Biol. 347 (2006) 125–143.
preparation of the manuscript. [21] X. Hu, Z. Zhang, Understanding the genetic mechanisms of cancer drug resistance
using genomic approaches, Trends Genet. 32 (2016) 127–137.
[22] K.S. Putt, G.W. Chen, J.M. Pearson, J.S. Sandhorst, M.S. Hoagland, J.T. Kwon,
Ethical approval S.K. Hwang, H. Jin, M.I. Churchwell, M.H. Cho, D.R. Doerge, W.G. Helferich,
P.J. Hergenrother, Small-molecule activation of procaspase-3 to caspase-3 as a
personalized anticancer strategy, Nat. Chem. Biol. 2 (2006) 543–550.
This article does not contain any studies with human participants or [23] K. Saeki, A. Yuo, M. Kato, K. Miyazono, Y. Yazaki, F. Takaku, Cell density-
animals performed by any of the authors. dependent apoptosis in HL-60 cells, which is mediated by an unknown soluble

9
A. Razaghi et al. Experimental Cell Research xxx (xxxx) xxx–xxx

factor, is inhibited by transforming growth factor 1 and overexpression of Bcl-2, J. Am. J. Pathol. 168 (2006) 2054–2063.
Biol. Chem. 272 (1997) 20003–20010. [36] M.D.S. Jean, Fi Brignole, G. Feldmann, A. Goguel, C. Baudouin, Interferon-γ
[24] L. Pronsato, A. La Colla, A.C. Ronda, L. Milanesi, R. Boland, A. Vasconsuelo, High induces apoptosis and expression of inflammation-related proteins in Chang
passage numbers induce resistance to apoptosis in C2C12 muscle cells, Biocell: Off. conjunctival cells, Invest Ophthalmol. Vis. Sci. 40 (1999) 2199–2212.
J. Soc. Latinoam. De. Microsc. Electron. … et al. 37 (2013) 1–9. [37] C. Pm, I. Ap, H. Pm, The macroheterogeneity of recombinant human interferon‐
[25] M. Pasparakis, P. Vandenabeele, Necroptosis and its role in inflammation, Nature gamma produced by Chinese‐hamster ovary cells is affected by the protein and lipid
517 (2015) 311–320. content of the culture medium, Biotechnol. Appl. Biochem. 21 (1995) 87–100.
[26] P. Li, Q. Du, Z. Cao, Z. Guo, J. Evankovich, W. Yan, Y. Chang, L. Shao, D.B. Stolz, [38] N. Jenkins, P. Castro, S. Menon, A. Ison, A. Bull, Effect of lipid supplements on the
A. Tsung, D.A. Geller, Interferon-gamma induces autophagy with growth inhibition production and glycosylation of recombinant interferon-gamma expressed in CHO
and cell death in human hepatocellular carcinoma (HCC) cells through interferon- cells, Cytotechnology 15 (1994) 209–215.
regulatory factor-1 (IRF-1), Cancer Lett. 314 (2012) 213–222. [39] D.C. James, R.B. Freedman, M. Hoare, O.W. Ogonah, B.C. Rooney, O.A. Larionov,
[27] Q. Liu, X. Liu, J. Gao, X. Shi, X. Hu, S. Wang, Y. Luo, Overexpression of DOC-1R V.N. Dobrovolsky, O.V. Lagutin, N. Jenkins, N-glycosylation of recombinant human
inhibits cell cycle G1/S transition by repressing CDK2 expression and activation, interferon-y produced in different animal expression systems, Biotechnology 13
Int. J. Biol. Sci. 9 (2013) 541–549. (1995) 64771.
[28] S. Frokjaer, D.E. Otzen, Protein drug stability: a formulation challenge, Nat. Rev. [40] J. Rieder, R. Jahnke, M. Schloesser, M. Seibel, M. Czechowski, C. Marth,
Drug Discov. 4 (2005) 298–306. G. Hoffmann, Nitric oxide-dependent apoptosis in ovarian carcinoma cell lines,
[29] B.D. Bell, S. Leverrier, B.M. Weist, R.H. Newton, A.F. Arechiga, K.A. Luhrs, Gynecol. Oncol. 82 (2001) 172–176.
N.S. Morrissette, C.M. Walsh, FADD and caspase-8 control the outcome of [41] E.J. Kim, J.M. Lee, S.E. Namkoong, S.J. Um, J.S. Park, Interferon regulatory
autophagic signaling in proliferating T cells, Proc. Natl. Acad. Sci. 105 (2008) factor-1 mediates interferon-gamma-induced apoptosis in ovarian carcinoma cells,
16677–16682. J. Cell Biochem. 85 (2002) 369–380.
[30] R.J. Thapa, S.H. Basagoudanavar, S. Nogusa, K. Irrinki, K. Mallilankaraman, [42] Y.-Q. Guan, Z. Zheng, L. Liang, Z. Li, L. Zhang, J. Du, J.-M. Liu, The apoptosis of
M.J. Slifker, A.A. Beg, M. Madesh, S. Balachandran, NF-kappaB protects cells from OVCAR-3 induced by TNF-α plus IFN-γ co-immobilized polylactic acid copolymers,
gamma interferon-induced RIP1-dependent necroptosis, Mol. Cell Biol. 31 (2011) J. Mater. Chem. 22 (2012) 14746.
2934–2946. [43] E.C. Alappat, C. Feig, B. Boyerinas, J. Volkland, M. Samuels, A.E. Murmann,
[31] C. Vivo, F. Levy, Y. Pilatte, J. Fleury-Feith, P. Chretien, I. Monnet, L. Kheuang, A. Thorburn, V.J. Kidd, C.A. Slaughter, S.L. Osborn, A. Winoto, W.J. Tang,
M.C. Jaurand, Control of cell cycle progression in human mesothelioma cells M.E. Peter, Phosphorylation of FADD at serine 194 by CKIalpha regulates its
treated with gamma interferon, Oncogene 20 (2001) 1085–1093. nonapoptotic activities, Mol. Cell 19 (2005) 321–332.
[32] A. Kano, Y. Watanabe, N. Takeda, T.S.-i. Aizawa, Analysis of IFN-γ-induced cell [44] E.W. Lee, J. Seo, M. Jeong, S. Lee, J. Song, The roles of FADD in extrinsic apoptosis
cycle arrest and cell death in hepatocytes, J. Biochem. 121 (1997) 677–683. and necroptosis, BMB Rep. 45 (2012) 496–508.
[33] X. Huang, J.J. Barchi Jr., F.D. Lung, P.P. Roller, P.L. Nara, J. Muschik, [45] L.C. Li, S. Jayaram, L. Ganesh, L. Qian, J. Rotmensch, A.V. Maker, B.S. Prabhakar,
R.R. Garrity, Glycosylation affects both the three-dimensional structure and Knockdown of MADD and c-FLIP overcomes resistance to TRAIL-induced apop-
antibody binding properties of the HIV-1IIIB GP120 peptide RP135, Biochemistry tosis in ovarian cancer cells, Am. J. Obstet. Gynecol. 205 (362) (2011) e312–e325.
36 (1997) 10846–10856. [46] J.O. Pyo, M.H. Jang, Y.K. Kwon, H.J. Lee, J.I. Jun, H.N. Woo, D.H. Cho, B. Choi,
[34] X. Xu, X.Y. Fu, J. Plate, A.S. Chong, IFN-gamma induces cell growth inhibition by H. Lee, J.H. Kim, N. Mizushima, Y. Oshumi, Y.K. Jung, Essential roles of Atg5 and
Fas-mediated apoptosis: requirement of STAT1 protein for up-regulation of Fas FADD in autophagic cell death: dissection of autophagic cell death into vacuole
and FasL expression, Cancer Res. 58 (1998) 2832–2837. formation and cell death, J. Biol. Chem. 280 (2005) 20722–20729.
[35] D. Rosner, V. Stoneman, T. Littlewood, N. McCarthy, N. Figg, Y. Wang, G. Tellides, [47] S.Y. Park, J.W. Seol, Y.J. Lee, J.H. Cho, H.S. Kang, I.S. Kim, S.H. Park, T.H. Kim,
M. Bennett, Interferon-gamma induces Fas trafficking and sensitization to apop- J.H. Yim, M. Kim, T.R. Billiar, D.W. Seol, IFN-gamma enhances TRAIL-induced
tosis in vascular smooth muscle cells via a PI3K- and Akt-dependent mechanism, apoptosis through IRF-1, Eur. J. Biochem. 271 (2004) 4222–4228.

10

You might also like