26.01.2014 Views

PARP and RIP-1 are required for autophagy induced by 11 ...

PARP and RIP-1 are required for autophagy induced by 11 ...

PARP and RIP-1 are required for autophagy induced by 11 ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Basic Research Paper<br />

Autophagy 7:6, 1-15; June 20<strong>11</strong>; © 20<strong>11</strong> L<strong>and</strong>es Bioscience<br />

Basic Research Paper<br />

<strong>PARP</strong> <strong>and</strong> <strong>RIP</strong>-1 <strong>are</strong> <strong>required</strong> <strong>for</strong> <strong>autophagy</strong><br />

<strong>induced</strong> <strong>by</strong> <strong>11</strong>'-deoxyverticillin A, which precedes<br />

caspase-dependent apoptosis<br />

This manuscript has been published online, prior to printing. Once the issue is complete <strong>and</strong> page numbers have been assigned, the citation will change accordingly.<br />

Nan Zhang, 1,2,† Yali Chen, 3,† Ruixuan Jiang, 4 Erwei Li, 3 Xiuling Chen, 2 Zhijun Xi, 1 Yinglu Guo, 1 Xingzhong Liu, 3 Yuguang Zhou, 2,5<br />

Yongsheng Che 3,5, * <strong>and</strong> Xuejun Jiang 2,3, *<br />

1<br />

National Research Center of Urological Cancer; Institute of Urology; Peking University First Hospital; 2 Biological Resource Center; 3 Key Laboratory of Systematic Mycology<br />

& Lichenology; 5 State Key Laboratory of Microbial Resources;Institute of Microbiology; Chinese Academy of Sciences; 4 Capital Medical University;<br />

6<br />

Beijing Institute of Pharmacology & Toxicology; Beijing, China<br />

The epipolythiodioxopiperazines (ETPs) <strong>are</strong> fungal secondary metabolites proven to trigger both apoptotic <strong>and</strong> necrotic<br />

cell death of tumor cells. However, the underlying mechanism of their regulatory role in macro<strong>autophagy</strong> <strong>and</strong> the<br />

interplay between <strong>autophagy</strong> <strong>and</strong> apoptosis initiated <strong>by</strong> the ETPs, remain unexplored. In the current work, we found that<br />

<strong>11</strong>'-deoxyverticillin A (C42), a member of the ETPs, induces autophagosome <strong>for</strong>mation, accumulation of microtubuleassociated<br />

protein 1 light chain 3-ii (LC3-ii) <strong>and</strong> degradation of sequestosome 1 (SQSTM1/p62). In addition, the LC3-ii<br />

accrual <strong>and</strong> p62 degradation occur prior to caspase activation <strong>and</strong> coincide with <strong>PARP</strong> activation. Inhibition of <strong>autophagy</strong><br />

©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

<strong>by</strong> either chemical inhibitors or <strong>by</strong> RNA interference single knockdown of essential autophagic genes partially reduces<br />

the cell death <strong>and</strong> the cleavage of both caspase 3 <strong>and</strong> <strong>PARP</strong>. Necrostatin-1, a specific inhibitor of necroptosis, inhibits both<br />

the augmentation of LC3-ii <strong>and</strong> the cleavage of caspase 3, which was confirmed <strong>by</strong> depletion of receptor-interacting<br />

protein 1 (<strong>RIP</strong>-1), a crucial necrostatin-1-targeted adaptor kinase mediating cell death <strong>and</strong> survival. Moreover, inhibition<br />

of <strong>PARP</strong> <strong>by</strong> either chemical inhibitors or RNA interference provides obvious protection <strong>for</strong> cell viability <strong>and</strong> suppresses<br />

Donotdistribute.<br />

the LC3-ii accretion caused <strong>by</strong> C42 treatment. Interestingly, double silencing of LC3 <strong>and</strong> p62 completely suppressed<br />

<strong>PARP</strong> cleavage <strong>and</strong> concurrently <strong>and</strong> maximally augmented the PAR <strong>for</strong>mation <strong>induced</strong> <strong>by</strong> C42. Collectively, we have<br />

demonstrated that C42 enhances the cellular autophagic process, which requires both <strong>PARP</strong> <strong>and</strong> <strong>RIP</strong>-1 participation,<br />

preceding <strong>and</strong> possibly augmenting, the caspase-dependent apoptotic cell death.<br />

Introduction<br />

Macro<strong>autophagy</strong> (hereafter referred to as <strong>autophagy</strong>) was originally<br />

described as a cellular adaptation to starvation involving<br />

the vesicular sequestration of long-lived cytoplasmic proteins <strong>and</strong><br />

organelles such as mitochondria. 1,2 The newly created doublemembrane<br />

autophagosome consequently fuses with lysosomes to<br />

<strong>for</strong>m an autolysosome, in which the engulfed cytoplasmic material<br />

is hydrolyzed through the action of acid-dependent enzymes,<br />

ensuring that the amino acids <strong>and</strong> other macromolecular precursors<br />

can be recycled. 3 Although <strong>autophagy</strong> serves as a cell survival<br />

mechanism, if overactivated <strong>and</strong> allowed to go to excess, it will<br />

eventually lead to cell death <strong>by</strong> depletion of the cell’s organelles<br />

<strong>and</strong> critical proteins. 4-6<br />

Cell death plays essential roles in organ development, tissue<br />

homeostasis <strong>and</strong> degenerative diseases. Three major types of cell<br />

death have been described—apoptosis, 7 <strong>autophagy</strong>, 8,9 <strong>and</strong> necrosis,<br />

10,<strong>11</strong> based on their distinct cell morphology. The apoptotic cell<br />

†<br />

These authors contributed equally to this work.<br />

Key words: <strong>11</strong>'-deoxyverticillin A, <strong>autophagy</strong>, apoptosis, <strong>PARP</strong>, <strong>RIP</strong><br />

death is characterized <strong>by</strong> the activation of caspases <strong>and</strong> <strong>for</strong>mation<br />

of apoptotic bodies, 12,13 whereas the autophagic one is differentiated<br />

<strong>by</strong> large-scale sequestration of portions of the cytoplasm<br />

in autophagosomes, giving the cell a characteristic vacuolated<br />

appearance. 14 Accumulating evidence suggests the existence of<br />

several molecular connections among apoptosis, <strong>autophagy</strong> <strong>and</strong><br />

necrosis. 15-17 In response to specific perturbations, the same input<br />

signal can cause cells to change from one cell death manifestation<br />

to another, 17,18 with a mixed type of cell death also observed<br />

in some cases. 19,20 In those cells undergoing persistent <strong>autophagy</strong>,<br />

hallmarks of apoptosis, such as caspase activation, necrotic cell<br />

death, organelle swelling <strong>and</strong> plasma membrane rupture, <strong>are</strong><br />

often observed. 21 In nutrient-deprived or growth factor-withdrawn<br />

cells, <strong>autophagy</strong> protects the cells <strong>and</strong> allows their survival<br />

<strong>by</strong> inhibiting apoptosis. 22 In the presence of caspase inhibitors,<br />

<strong>autophagy</strong> also protects the cells from caspase-independent<br />

death. 16 Depending on the cellular setting, the same proteins can<br />

regulate both autophagic <strong>and</strong> apoptotic processes. For example,<br />

*Correspondence to: Yongsheng Che <strong>and</strong> Xuejun Jiang; Email: cheys@im.ac.cn <strong>and</strong> jiangxj@im.ac.cn<br />

Submitted: 09/22/10; Revised: 01/22/<strong>11</strong>; Accepted: 02/09/<strong>11</strong><br />

DOI:<br />

www.l<strong>and</strong>esbioscience.com Autophagy 1


p53, a potent inducer of apoptosis, also activates <strong>autophagy</strong><br />

through its target gene, DRAM (damage-regulated modulator<br />

of <strong>autophagy</strong>). 23 Beclin 1, <strong>required</strong> <strong>for</strong> <strong>for</strong>mation of the autophagic<br />

vesicle, also interacts with both Bcl-2 <strong>and</strong> Bcl-x L<br />

. 24 Calpainmediated<br />

cleavage of ATG5 is a critical pro-apoptotic event that<br />

enables the activation of caspase-dependent cell death. 25 Cho <strong>and</strong><br />

Kim et al. have identified ATG6 as a link between apoptotic <strong>and</strong><br />

autophagic cell death. 26 Induction of <strong>autophagy</strong> <strong>by</strong> ATG1 inhibits<br />

cell growth <strong>and</strong> induces apoptotic cell death. 27 It has been suggested<br />

that <strong>autophagy</strong> relates to apotosis <strong>by</strong> acting as a partner,<br />

an antagonist or an enabler of apoptosis depending on cell type,<br />

stimulus <strong>and</strong> environment. 28,29<br />

<strong>11</strong>'-deoxyverticillin A, a natural product isolated from the<br />

Cordyceps-colonizing fungus Gliocladium sp., 30 is a member of<br />

a class of fungal secondary metabolites known as epipolythiodioxopiperazines<br />

(ETPs). 31 The ETPs have been previously<br />

reported to have a broad range of biological activities including<br />

antimicrobial, antiparasitic, antiviral, immunosuppressive <strong>and</strong>/<br />

or anti-inflammatory effects. 31 Gliotoxin is the first ETP reported<br />

<strong>and</strong> the best characterized one. Its toxicity to animal cell cultures<br />

has been studied extensively, <strong>and</strong> it is thought to be mediated in<br />

at least two ways: conjugation to, <strong>and</strong> consequent inactivation of,<br />

proteins with susceptible thiol residues <strong>and</strong> generation of reactive<br />

oxygen species (ROS) via redox cycling. 32 Results from previous<br />

studies have demonstrated that gliotoxin causes the apoptotic <strong>and</strong><br />

colon carcinoma cell line HCT<strong>11</strong>6 <strong>and</strong> adenocarcinoma colon<br />

cell lines SW480 <strong>and</strong> SW620. As shown in Figure 1A, C42<br />

caused cell viability loss of HCT<strong>11</strong>6 in a time- <strong>and</strong> concentration-dependent<br />

manner <strong>and</strong> similar inhibitory effects were also<br />

observed in SW480 <strong>and</strong> SW620 cells (Sup. Fig. 1A <strong>and</strong> B). In<br />

addition, flow cytometry data indicated that the C42-<strong>induced</strong><br />

cell death of HCT<strong>11</strong>6 could be either apoptotic or necrotic (may<br />

be either necrosis or secondary necrosis) (Fig. 1B).<br />

Since the pan-caspase inhibitor Z-VAD-FMK (Z-V-FMK)<br />

can block caspase-dependent apoptosis, 36 we next examined<br />

the C42-dependent cell death in the presence of this inhibitor.<br />

Interestingly, pre-treatment of HCT<strong>11</strong>6 (Fig. 1C) with the<br />

inhibitor provided only partial protection against C42-<strong>induced</strong><br />

cell death. This finding was confirmed <strong>by</strong> colony growth in a<br />

survival assay, in which the inhibitor was nearly unable to rescue<br />

the C42-<strong>induced</strong> cell death (Fig. 1D). Results from additional<br />

experiments confirmed that porcine aortic endothelial<br />

(PAE) cells overexpressing the anti-apoptotic protein Bcl-x L<br />

(Sup. Fig. 1C, 2A <strong>and</strong> 2B) <strong>and</strong> p53 -/- HCT<strong>11</strong>6 cells lacking the<br />

pro-apoptotic protein p53 (Sup. Fig. 1D) were both sensitive<br />

to C42, indicating that caspase-independent or non-apoptotic<br />

mechanisms also contribute to C42-<strong>induced</strong> cell death. In addition,<br />

we also found that 3-methyladenine (3-MA), a widely used<br />

inhibitor of <strong>autophagy</strong>, 37 inhibited the C42-<strong>induced</strong> cell death<br />

©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

(Fig. 1E). To investigate the relationship between <strong>autophagy</strong><br />

necrotic death of animal cells. 31,33 In one study, cell death caused<br />

<strong>by</strong> gliotoxin switched from apoptosis to necrosis at concentrations<br />

over 10 μM. 33 Leptosins, the ETPs isolated from a marine<br />

fungus, were cytotoxic to various tumor cells <strong>by</strong> inactivation of<br />

<strong>and</strong> C42-dependent cell death, two known activators of <strong>autophagy</strong>,<br />

rapamycin <strong>and</strong> trehalose, were used to promote the cellular<br />

autophagic processes. 38,39 As expected, the combination of<br />

Donotdistribute.<br />

either rapamycin or trehalose with C42 showed a measurable <strong>and</strong><br />

the Akt/protein kinase B pathway or <strong>by</strong> inhibition of DNA topoisomerases<br />

I <strong>and</strong> II. 34 Chen <strong>and</strong> Ding reported that in HCT-<strong>11</strong>6<br />

human colon cancer cells, either the p53 pathway or p38 MAPK<br />

signaling plays a role in mediation of <strong>11</strong>,<strong>11</strong>'-dideoxyverticillin A<br />

<strong>induced</strong> G 2<br />

/M arrest. 35<br />

Although the above-mentioned natural products <strong>are</strong> cytotoxic<br />

against various tumor cells <strong>by</strong> inducing apoptosis or necrosis, their<br />

effects on the autophagic process have not been investigated. The<br />

purpose of this study is to underst<strong>and</strong> the molecular mechanism<br />

<strong>for</strong> C42’s cytotoxicity to human tumor cells <strong>by</strong> exploring the two<br />

major categories of cell death. We found that it induces <strong>for</strong>mation<br />

of intracellular vacuoles in human colon carcinoma cells <strong>and</strong><br />

enhances the molecular hallmarks of <strong>autophagy</strong>, such as LC3<br />

processing <strong>and</strong> p62 degradation, while concomitantly activating<br />

caspases <strong>and</strong> leading to cleavage of <strong>PARP</strong>. Inhibition of <strong>PARP</strong><br />

or <strong>RIP</strong> <strong>by</strong> either the chemical inhibitors or RNA interference<br />

knockdown reduces the LC3-II accrual, <strong>PARP</strong> activation <strong>and</strong><br />

the cleavage of apoptotic hallmarks as well as cell death. Thus,<br />

activation of both <strong>PARP</strong> <strong>and</strong> <strong>RIP</strong> is <strong>required</strong> <strong>for</strong> C42-<strong>induced</strong><br />

<strong>autophagy</strong>, which is closely associated with caspase-dependent<br />

cell death.<br />

Results<br />

<strong>11</strong>'-deoxyverticillin A (C42) induces cell death of human colon<br />

cell lines which is partially inhibited <strong>by</strong> 3-MA. In this study,<br />

the C42-<strong>induced</strong> cell death was first examined using the human<br />

more potent inhibitory effect on cell viability (Sup. Fig. 1E) <strong>and</strong><br />

both agents were able to attenuate cell viability <strong>by</strong> themselves,<br />

suggesting that enhanced <strong>autophagy</strong> may directly result in cell<br />

death (Sup. Fig. 1E).<br />

C42 enhances <strong>autophagy</strong> in colon cells. Electron microscopy,<br />

which is believed to be one of the most convincing approaches, 40<br />

was used to determine whether C42 induces <strong>autophagy</strong> or not.<br />

Comp<strong>are</strong>d to the control, an obvious accrual of membrane vacuoles<br />

was found in the C42-treated HCT<strong>11</strong>6 cells <strong>and</strong> cytosolic<br />

components or organelles were sequestered in some of those<br />

vacuoles. Autophagosome-like vacuoles with double-membrane<br />

structures (high magnification) <strong>and</strong> a segment of the doublemembrane<br />

being <strong>for</strong>med between a vacuole <strong>and</strong> mitochondrion<br />

(black arrowhead) were also observed (Fig. 2A). A swollen (white<br />

arrow) <strong>and</strong> degenerating mitochondrion (white arrow) <strong>and</strong> mitochondria<br />

inside of vacuole (black arrow) were also noted (Fig.<br />

2A <strong>and</strong> Sup. Fig. 2C). In addition, characteristics of apoptosis,<br />

such as the condensation <strong>and</strong> margination of chromation on<br />

the nuclear membrane, were found in these cells (Fig. 2A <strong>and</strong><br />

white arrowhead), indicating that C42 <strong>induced</strong> <strong>autophagy</strong> <strong>and</strong><br />

apoptosis in the same cell <strong>and</strong> directly <strong>induced</strong> the degradation<br />

of mitochondria. To further measure autophagosome <strong>for</strong>mation<br />

in the C42-treated cells, HCT<strong>11</strong>6 cells were transfected with a<br />

green fluorescent protein (GFP) <strong>and</strong> LC3 fusion protein (a specific<br />

marker of autophagosomes), 41 <strong>and</strong> visualized <strong>by</strong> confocal<br />

microscopy. In contrast to the C42-exposed cells, which showed<br />

increased punctate staining of GFP-LC3, GFP-LC3 staining<br />

2 Autophagy Volume 7 Issue 6


©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

Donotdistribute.<br />

Figure 1. <strong>11</strong>'-deoxyverticillin A (C42) induces programmed cell death <strong>and</strong> 3-MA application results in resistance to the C42-<strong>induced</strong> cell death. (A)<br />

hcT<strong>11</strong>6 cells were treated with C42 (0.05–0.5 μM) <strong>for</strong> up to 48 h; cell viability was analyzed <strong>by</strong> MTS assay as described in Materials <strong>and</strong> Methods. Data<br />

<strong>are</strong> presented as mean ± SD <strong>and</strong> <strong>are</strong> representatives of three independent experiments. (B) Following treatment of HCT<strong>11</strong>6 cells with C42 (0.25, 0.5<br />

μM) <strong>for</strong> 12 h, the apoptosis <strong>and</strong> necrosis <strong>induced</strong> were determined <strong>by</strong> flow cytometry. Apoptotic: AV-positive <strong>and</strong> PI-negative; necrotic: PI-positive;<br />

AV: annexin-V. The data <strong>are</strong> presented as mean ± SD from three independent experiments (C) HCT<strong>11</strong>6 cells were treated with C42 (0.1–0.5 μM) <strong>for</strong> 24<br />

h in the presence or absence of Z-V-FMK (20 μM) be<strong>for</strong>e detection <strong>by</strong> MTS assay. (D) Colony survival assays in HCT<strong>11</strong>6 cells were per<strong>for</strong>med following<br />

the treatment of HCT<strong>11</strong>6 cells with C42 (0.1 μM) <strong>for</strong> 6 h in the presence or absence of Z-V-FMK. Data represent the mean ± SD of three experiments,<br />

each per<strong>for</strong>med in triplicate. (E) HCT<strong>11</strong>6 cells were treated with C42 (0.1–0.5 μM) <strong>for</strong> 24 h in the presence or absence of 3-MA (2 mM), cell viability was<br />

analyzed <strong>by</strong> MTS assay. Ctrl: cells with equal amount of DMSO.<br />

remained diffuse in the control cells (Fig. 2B) (p < 0.01). Since<br />

the ratio of LC3-II to actin is an accurate indicator of <strong>autophagy</strong>,<br />

42 the expression of LC3-II in HCT<strong>11</strong>6 (Fig. 2C), SW480<br />

(Sup. Fig. 2D) <strong>and</strong> SW620 (Sup. Fig. 2E) cells was analyzed.<br />

Treatment with C42 increased the ratio of LC3-II to actin relative<br />

to the controls in all the cells tested.<br />

www.l<strong>and</strong>esbioscience.com Autophagy 3


©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

Donotdistribute.<br />

Figure 2. C42 induces accumulation of LC3-ii <strong>and</strong> enhances autophagic flux. (A) Electron microscopy was per<strong>for</strong>med on vehicle- (Ctrl) <strong>and</strong> C42-treated<br />

(0.25 μM, 6 h) HCT<strong>11</strong>6 cells as described in Materials <strong>and</strong> Methods. White arrowheads: the condensation <strong>and</strong> margination of the chromatin on nuclear<br />

membrane; left of lower part: swollen <strong>and</strong> degenerating mitochondria (white arrows), a vacuole with mitochondria inside (black arrow), a segment of<br />

double-membrane structure between a vacuole <strong>and</strong> mitochondrion (black arrowhead); right of lower part: high-contrast image of the cell region indicated<br />

<strong>by</strong> white rectangle. (B) HCT<strong>11</strong>6 cells were transfected with a plasmid expressing GFP-LC3. After 24 h, the cells were incubated <strong>for</strong> 6 h at 37°C in<br />

DMEM medium with DMSO (Ctrl) or C42 (0.25 μM). Following fixation, cells were stained with DAPI <strong>and</strong> visualized <strong>by</strong> confocal microscopy. The number<br />

of punctate GFP-LC3 in each cell was counted, <strong>and</strong> at least 50 cells were included <strong>for</strong> each group. The data were normally distributed <strong>and</strong> statistically<br />

analyzed using one-way ANOVA. The double asterisk denotes the group is statistically different from the control groups (p < 0.01). HCT<strong>11</strong>6 cells were<br />

treated with C42 (0.05–0.5 μM) <strong>for</strong> 12 h (C <strong>and</strong> D) or in the presence of Baf A1 (10 nM) (E) <strong>and</strong> cell lysates were prep<strong>are</strong>d <strong>and</strong> analyzed <strong>by</strong> immunoblotting<br />

using the indicated antibodies. Densitometry was per<strong>for</strong>med <strong>for</strong> quantification <strong>and</strong> the ratios of LC3-ii, p62, phosphorylated mTOR <strong>and</strong> phosphorylated<br />

p70S6K (T389) to actin <strong>are</strong> presented below the blots. The ratios <strong>are</strong> representative of at least three independent experiments.<br />

Mammalian target of rapamycin (mTOR) inhibits <strong>autophagy</strong><br />

<strong>and</strong> its kinase activity can be inferred <strong>by</strong> measuring the levels<br />

of phosphorylation of its substrates, such as p70S6 kinase. 43 To<br />

study whether or not C42-<strong>induced</strong> <strong>autophagy</strong> was associated<br />

with this known pathway, the phosphorylation state of mTOR<br />

<strong>and</strong> p70S6 kinase was examined. As expected, the phosphorylation<br />

of mTOR <strong>and</strong> p70S6K was attenuated in a concentrationdependent<br />

manner in response to the agent (Fig. 2D). To detect<br />

4 Autophagy Volume 7 Issue 6


autophagic flux, the level of LC3-II was measured in<br />

the presence of bafilomycin A1 (Baf A1), which acts<br />

as a specific inhibitor of the vacuolar type H + -ATPase<br />

in cells, thus blocking acidification of organelles <strong>and</strong><br />

subsequent fusion of the autophagosome with the<br />

lysosome. 44 Baf A1 addition resulted in further accumulation<br />

of LC3-II in the cells tested (Fig. 2E). Since<br />

p62/SQSTM1 is considered to be a selective substrate<br />

of <strong>autophagy</strong> <strong>and</strong> will be accumulated when <strong>autophagy</strong><br />

is inhibited, 43 we also examined the levels of p62<br />

upon challenge with C42. As shown in Figure 2C<br />

<strong>and</strong> Supplemental Figure 2D <strong>and</strong> E, the degradation<br />

of p62 <strong>induced</strong> <strong>by</strong> C42 was concentration-dependent,<br />

indicating that C42 actually enhances the autophagic<br />

process.<br />

The mRNA level of LC3 was also measured <strong>by</strong><br />

semiquantified RT-PCR <strong>and</strong> it was found that C42<br />

increased the level only at 0.5 μM comp<strong>are</strong>d to the<br />

control. Nevertheless, the augmentation of LC3<br />

mRNA is not significant <strong>and</strong> could be due to the<br />

reduced actin mRNA (Sup. Fig. 2F). There<strong>for</strong>e, C42<br />

may affect the LC3 expression at high concentrations<br />

or at some time point via an unknown mechanism.<br />

Autophagy <strong>induced</strong> <strong>by</strong> C42 precedes caspase<br />

©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

activation. Caspase activation was examined to<br />

explore the interplay between <strong>autophagy</strong> <strong>and</strong> apoptosis<br />

upon C42 challenge of HCT<strong>11</strong>6 cells. After 1 h of<br />

treatment, we found that C42 <strong>induced</strong> the accumula-<br />

Donotdistribute.<br />

tion of LC3-II <strong>and</strong> degradation of p62 in a time- <strong>and</strong><br />

concentration-dependent manner (Fig. 3A). Using<br />

antibodies against the early apoptosis hallmarks, 26,29<br />

caspase 9 <strong>and</strong> <strong>PARP</strong>-1, we noticed that LC3-II<br />

accrual <strong>and</strong>/or p62 degradation preceded the cleavage<br />

of caspases <strong>and</strong> <strong>PARP</strong> (Fig. 3A <strong>and</strong> Sup. Fig. 3A<br />

<strong>and</strong> B). In contrast to caspase activation <strong>and</strong> <strong>PARP</strong>-1<br />

cleavage detected at 4 h, the C42-<strong>induced</strong> <strong>for</strong>mation<br />

of poly(ADP-ribose) (PAR) polymers appe<strong>are</strong>d to<br />

occur concomitantly with the autophagic process from an early<br />

time-point of treatment (Fig. 3A). PAR, a direct result of <strong>PARP</strong>-1<br />

activation, is readily activated in response to DNA damage <strong>and</strong><br />

well-associated with necrotic cell death. 10 Thus, we assume that<br />

C42, in addition to inducing caspase-dependent apoptosis, is also<br />

a powerful activator of <strong>PARP</strong>-1 <strong>and</strong> may induce caspase-independent<br />

or necrosis-like cell death in those cells if the stimulus<br />

is continually presented. On the other h<strong>and</strong>, activation of <strong>PARP</strong><br />

is a cellular response to C42 stress <strong>and</strong> may be <strong>required</strong> <strong>for</strong> C42-<br />

<strong>induced</strong> <strong>autophagy</strong>. Results from additional experiment indicated<br />

that the mRNA level of <strong>PARP</strong>-1 is increased only at 0.5 μM of<br />

C42 as measured <strong>by</strong> semiquantified RT-PCR (Sup. Fig. 3C).<br />

In PAE cells (in which caspase 3 activation is easily detected<br />

comp<strong>are</strong>d to EGFR expressing cells), Bcl-x L<br />

overexpression attenuated<br />

the C42-<strong>induced</strong> caspase 3 cleavage (Fig. 3B), but failed<br />

to block its activation (Sup. Fig. 3D). Additionally, the <strong>for</strong>ced<br />

overexpression of this anti-apoptotic protein showed no sign of<br />

affecting the C42-dependent LC3-II accumulation (Fig. 3B).<br />

Thus, Bcl-x L<br />

alone is insufficient to prevent the caspase activation<br />

Figure 3. C42 induces accumulation of LC3-ii or degradation of p62 that occurs prior<br />

to the activation of caspases <strong>and</strong> cleavage of <strong>PARP</strong>. HCT<strong>11</strong>6 (A), PAE-EGFR (B) or PAE-<br />

Bcl-x L<br />

(B) cells were treated <strong>for</strong> up to 4 h; cell lysates were prep<strong>are</strong>d <strong>and</strong> analyzed <strong>by</strong><br />

immunoblotting using the indicated antibodies. The asterisk indicates a nonspecific<br />

b<strong>and</strong>.<br />

<strong>and</strong> cell death <strong>and</strong> the overexpression of Bcl-x L<br />

rescues autophagosome<br />

<strong>for</strong>mation. Notably, we were unable to observe a clear<br />

accumulation of LC3-II in C42-treated PAE-GFP cells (Sup.<br />

Fig. 3E), which <strong>are</strong> more sensitive to the agent because their caspase<br />

3 activation can be readily detected at an earlier stage (data<br />

not shown). We assume that C42 mainly activates the apoptosis<br />

pathway in those cells <strong>and</strong> hyperactivated apoptosis impairs<br />

autophagosome <strong>for</strong>mation, whereas Bcl-x L<br />

<strong>and</strong> EGFR overexpression<br />

restore <strong>autophagy</strong> in C42-treated PAE cells. These findings<br />

provide supplementary evidence that C42-<strong>induced</strong> <strong>autophagy</strong><br />

could be associated with the apoptotic process.<br />

3-MA partially inhibits the C42-<strong>induced</strong> activation of both<br />

caspases <strong>and</strong> <strong>PARP</strong>. To explore the possible association of C42-<br />

<strong>induced</strong> <strong>autophagy</strong> with apoptosis, both inhibitors <strong>and</strong> enhancers<br />

of <strong>autophagy</strong> were evaluated in the following experiments. 3-MA<br />

(Fig. 4A) inhibited (or delayed) cellular <strong>autophagy</strong>, whereas<br />

rapamycin (Sup. Fig. 4A) <strong>and</strong> Z-V-FMK (Fig. 4C) promoted<br />

the process in the cells as evidenced <strong>by</strong> both LC3-II accrual <strong>and</strong><br />

p62 degradation. Notably, 3-MA inhibited both the cleavage <strong>and</strong><br />

www.l<strong>and</strong>esbioscience.com Autophagy 5


Figure 4. 3-MA inhibits C42-<strong>induced</strong> cleavage of <strong>PARP</strong>. HCT<strong>11</strong>6 cells<br />

were treated with C42 (0.1 <strong>and</strong> 0.25 μM) in the presence or absence of<br />

3-MA (A <strong>and</strong> B, 2 mM) or Z-V-FMK (C, 20 μM) <strong>for</strong> 24 h be<strong>for</strong>e analysis <strong>by</strong><br />

immunoblotting with the indicated antibodies. The ratios represent the<br />

results of three similar experiments.<br />

activation of <strong>PARP</strong> (Fig. 4A <strong>and</strong> Sup. Fig. 4B), whereas rapamycin<br />

showed opposite effects (Sup. Fig. 4A <strong>and</strong> C). Moreover, the<br />

inhibitory effect of 3-MA on activation of caspase 3 was concentration-dependent<br />

(Fig. 4B <strong>and</strong> Sup. Fig. 4E). Interestingly,<br />

application of Z-V-FMK remarkably increased the LC3-II level<br />

in a dose- <strong>and</strong> time-dependent manner in both HCT<strong>11</strong>6 <strong>and</strong> the<br />

PAE-Bcl-x L<br />

overexpressing cells (Fig. 4C <strong>and</strong> Sup. Fig. 4E <strong>and</strong><br />

F). Although Z-V-FMK <strong>induced</strong> accumulation of both p62 <strong>and</strong><br />

LC3-II in a short time, it failed to block the drop of the C42-<br />

<strong>induced</strong> p62 level over an extended period (Fig. 4C <strong>and</strong> Sup.<br />

Fig. 4F). In addition, we found that the presence of Z-V-FMK<br />

enhanced C42-<strong>induced</strong> PAR <strong>for</strong>mation (Sup. Fig. 4D). These<br />

findings suggest that this pan-caspase inhibitor could retard <strong>and</strong><br />

transiently inhibit C42-<strong>induced</strong> <strong>autophagy</strong> in a short period but<br />

enhance the process in the long run, probably via a feedback<br />

mechanism. Under these circumstances, enhanced <strong>autophagy</strong><br />

may directly lead to cell death or shift to caspase-independent<br />

apoptosis or necrosis-like cell death <strong>by</strong> hyperactivating <strong>PARP</strong>.<br />

©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

Both <strong>PARP</strong>-1 <strong>RIP</strong> play a role in the C42-<strong>induced</strong> <strong>autophagy</strong>.<br />

It is well known that hyperactivated <strong>PARP</strong>-1 causes depletion<br />

of NAD + <strong>and</strong> ATP, eventually leading to irreversible cellular<br />

energy failure <strong>and</strong> necrotic cell death. 10 In addition, recent stud-<br />

Donotdistribute.<br />

ies have shown that <strong>PARP</strong> also participates in mediation of <strong>autophagy</strong>.<br />

45,46 We also observed that the <strong>PARP</strong> activation concurred<br />

with the C42-dependent <strong>autophagy</strong>, implying that the activation<br />

of <strong>PARP</strong> may be <strong>required</strong> <strong>for</strong> the C42-<strong>induced</strong> <strong>autophagy</strong>. To<br />

explore <strong>PARP</strong>’s regulatory role in the C42-triggered <strong>autophagy</strong>,<br />

both chemical inhibitor <strong>and</strong> RNA interference approaches were<br />

employed. The presence of 1,5-dihydroxyisoquinoline (ISO),<br />

a potent inhibitor of poly(ADP-ribose) synthetase, 47 provided<br />

app<strong>are</strong>nt protection <strong>for</strong> the C42-treated cells in a concentration-dependent<br />

manner (Fig. 5A). Upon treatment with 3-aminobenzamide<br />

(3-AB), an inhibitor of <strong>PARP</strong>-1, 48 a similar cell<br />

viability rescue was observed (Sup. Fig. 5A). Pretreatment with<br />

ISO remarkably inhibited PAR <strong>for</strong>mation <strong>and</strong> decreased LC3-II<br />

levels <strong>and</strong> caspase 3 activation <strong>induced</strong> <strong>by</strong> C42 (Fig. 5B <strong>and</strong><br />

Figure 5 (See opposite page). <strong>PARP</strong> <strong>and</strong> <strong>RIP</strong> inhibition attenuate<br />

C42-<strong>induced</strong> accumulation of LC3-ii <strong>and</strong> result in resistance to C42-<br />

dependent cell death. (A) HCT<strong>11</strong>6 cells were treated with C42 (0.1, 0.25<br />

<strong>and</strong> 0.5 μM) in the presence or absence of necrostatin-1 (Nec-1, 30 μM)<br />

or 1,5-isoquinolinediol (isO, 30 μM) <strong>for</strong> 12 h be<strong>for</strong>e detection <strong>by</strong> MTS<br />

assay. (B) Immunoblotting analysis was per<strong>for</strong>med with the indicated<br />

antibodies following exposure to C42 (0.25 μM) in the presence or<br />

absence of Nec-1 (30 μM) or ISO (30 μM) <strong>for</strong> 12 h. HCT<strong>11</strong>6 cells were<br />

transfected with the control (Mock), <strong>PARP</strong> (C) or <strong>RIP</strong> siRNA (D); after 36<br />

h of transfection, cells were treated with C42 (0.25 μM) <strong>for</strong> 12 h be<strong>for</strong>e<br />

immunoblotting analysis with the indicated antibodies. (E) Viability of<br />

cells (C <strong>and</strong> D) was analyzed <strong>by</strong> MTS assay. The data were statistically<br />

analyzed using Student’s t-test, which demonstrated that the difference<br />

between control siRNA <strong>and</strong> <strong>PARP</strong> siRNA (p = 0.02) or <strong>RIP</strong>-1 siRNA (p<br />

= 0.03) was significant. *indicates a significant difference between the<br />

groups at level p < 0.05.<br />

6 Autophagy Volume 7 Issue 6


©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

Donotdistribute.<br />

Figure 5. For figure legend, see page 6.<br />

www.l<strong>and</strong>esbioscience.com Autophagy 7


Figure 6. NAC application <strong>and</strong> <strong>RIP</strong>-3 depletion suppress C42-<strong>induced</strong><br />

LC3-ii accumulation <strong>and</strong> <strong>PARP</strong> cleavage. (A) HCT<strong>11</strong>6 cells were transfected<br />

with the control (Mock) <strong>and</strong> <strong>RIP</strong>-3 siRNA, respectively; after 36<br />

h of transfection, cells were treated with C42 (0.25 μM) <strong>for</strong> 12 h be<strong>for</strong>e<br />

immunoblotting analysis with the indicated antibodies. (B) Immunoblotting<br />

analysis was per<strong>for</strong>med with the indicated antibodies following<br />

exposure to C42 (0.25 μM) in the presence or absence of NAC (5 mM)<br />

<strong>for</strong> 12 h. Densitometry was per<strong>for</strong>med <strong>for</strong> quantification <strong>and</strong> the ratios<br />

of LC3-ii to actin <strong>are</strong> presented in the graphs below the parts. The<br />

ratios <strong>are</strong> representative of two experiments. The asterisk indicates a<br />

nonspecific b<strong>and</strong>.<br />

suggested that both <strong>RIP</strong>-1 <strong>and</strong> <strong>PARP</strong>-1 <strong>are</strong> implicated in the<br />

C42-<strong>induced</strong> <strong>autophagy</strong>, through which they may influence the<br />

caspase activation <strong>and</strong> cell death.<br />

<strong>RIP</strong>-3, a protein kinase with an N-terminal kinase domain<br />

similar to <strong>RIP</strong>-1, has been reported to participate in the mediation<br />

of apoptosis <strong>and</strong> necrosis. 51,52 To test whether this kinase<br />

is also involved in the C42-dependent <strong>autophagy</strong> <strong>and</strong> apoptosis<br />

or not, RNA interference was per<strong>for</strong>med in HCT<strong>11</strong>6 cells.<br />

Abrogation of <strong>RIP</strong>-3 with siRNA revealed an inhibitory effect<br />

on the C42-<strong>induced</strong> LC3-II accumulation <strong>and</strong> <strong>PARP</strong> cleavage<br />

(Fig. 6A). Interestingly, depletion of <strong>RIP</strong>-3 alone also increased<br />

the basal level of LC3-II, whereas abrogation of both <strong>RIP</strong>-1 <strong>and</strong><br />

<strong>RIP</strong>-3 decreased both the basal <strong>and</strong> C42-<strong>induced</strong> LC3-II content<br />

©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

(Fig. 6A Sup. Fig. 5C). Thus, <strong>RIP</strong>-1 <strong>RIP</strong>-3 may coordinate<br />

in mediating the C42-dependent autophagic process.<br />

Since ROS has been proven to mediate <strong>autophagy</strong> in multiple<br />

contexts <strong>and</strong> cell types, 53-56 <strong>and</strong> the EPTs <strong>are</strong> capabe of generat-<br />

Donotdistribute.<br />

ing ROS in the cell culture system, 32 C42 may affect <strong>autophagy</strong><br />

Sup. Fig. 5B). Using a RNA interference approach, we found<br />

that knockdown of <strong>PARP</strong>-1 remarkably reduced the <strong>for</strong>mation of<br />

PAR, the level of LC3-II <strong>and</strong> cleavage of <strong>PARP</strong> (Fig. 5C). The<br />

remaining PAR, we assume, could be due to incomplete RNA<br />

interference or the activation of another <strong>PARP</strong> family member.<br />

There<strong>for</strong>e, the activation of <strong>PARP</strong> contributes to the C42-<br />

<strong>induced</strong> <strong>autophagy</strong>.<br />

Receptor-interacting protein (<strong>RIP</strong>), a protein with diverse<br />

functions, participates in the regulation of apoptosis, necrosis<br />

<strong>and</strong> autophagic cell death. 49 Application of Nec-1, a necroptosis<br />

inhibitor that blocks <strong>RIP</strong> kinase, 50 partially attenuated the C42-<br />

<strong>induced</strong> cell death, LC3-II accrual, caspase 3 cleavage <strong>and</strong> <strong>PARP</strong><br />

activation (Fig. 5A <strong>and</strong> B). Abrogation of <strong>RIP</strong>-1 with siRNA<br />

showed a similar inhibitory effect on the a<strong>for</strong>ementioned cleavage<br />

<strong>and</strong> activation (Fig. 5D). Moreover, silencing either <strong>PARP</strong>-1<br />

or <strong>RIP</strong>-1 provided partial, yet significant, protection <strong>for</strong> the<br />

C42-treated cells (p < 0.05) (Fig. 5E). Notably, <strong>RIP</strong>-1 depletion<br />

also affected the basal level of LC3, but failed to reduce the total<br />

LC3 content comp<strong>are</strong>d to the control (Fig. 5D). These findings<br />

via ROS production. To confirm this hypothesis, the widely used<br />

antioxidant, N-acetylcysteine (NAC), was used. As expected,<br />

NAC suppressed the C42-<strong>induced</strong> LC3-II accumulation <strong>and</strong><br />

<strong>PARP</strong> cleavage (Fig. 6B), suggesting that ROS mediates the<br />

C42-<strong>induced</strong> <strong>autophagy</strong> <strong>and</strong> apoptosis.<br />

Knockdown of <strong>autophagy</strong>-related genes partially attenuates<br />

the C42-<strong>induced</strong> cleavage of caspase <strong>and</strong> <strong>PARP</strong> as well as<br />

cell death. To verify the association between the C42-dependent<br />

cell death <strong>and</strong> <strong>autophagy</strong>, siRNA was per<strong>for</strong>med <strong>for</strong> the key<br />

autophagic genes LC3 <strong>and</strong> beclin 1. Knockdown of either LC3<br />

or beclin 1 inhibited the accumulation of LC3-II <strong>and</strong> degradation<br />

of p62 (Fig. 7A), confirming the interfering efficiency.<br />

Furthermore, the lack of expression of either LC3 or beclin 1<br />

resulted in a partial decrease in cell death (Fig. 7B), suggesting<br />

that <strong>autophagy</strong> is closely associated with or may augment,<br />

the C42-<strong>induced</strong> cell death. Notably, deletion of either LC3 or<br />

beclin 1 showed an inhibitory effect on <strong>PARP</strong>-1 cleavage, indicating<br />

that the C42-dependent <strong>autophagy</strong> is associated with<br />

caspase-dependent apoptosis (Fig. 7C <strong>and</strong> D <strong>and</strong> Sup. Fig. 6A).<br />

Although we noticed that beclin 1 depletion usually affected the<br />

level of LC3, the LC3 knockdown cells often maintained a relatively<br />

high content of Beclin 1 in a cell-type-dependent manner<br />

(Fig. 7A <strong>and</strong> Sup. Fig. 6B). The efficiency of siRNA could be<br />

the reason why the lack of beclin 1 exhibited a greater protective<br />

effect against the C42-<strong>induced</strong> cell death (Fig. 7B). In addition,<br />

double-perturbation of LC3 <strong>and</strong> beclin 1 offered greater protection<br />

against the C42-<strong>induced</strong> cell death (Fig. 7B). As expected,<br />

8 Autophagy Volume 7 Issue 6


©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

Donotdistribute.<br />

Figure 7. Autophagic genes silencing partially attenuates C42-dependent <strong>PARP</strong> cleavage <strong>and</strong> offers partial protection <strong>for</strong> the C42-treated cells. (A)<br />

hcT<strong>11</strong>6 cells transfected with the control (Mock) or LC3 (siLC3) or Beclin 1 (siBec1) siRNA were exposed to C42 (0.25 μM) <strong>for</strong> 12 h <strong>and</strong> cells were lysed<br />

<strong>and</strong> analyzed <strong>by</strong> immunoblotting with the antibodies indicated to confirm siRNA efficiency. (B) MTS assays were used to determine cell viability (A).<br />

Immunoblotting was per<strong>for</strong>med to detect the cleavage of <strong>PARP</strong> following siRNA knockdown of LC3 (C) <strong>and</strong> beclin 1 (D). Densitometry was per<strong>for</strong>med<br />

<strong>for</strong> quantification <strong>and</strong> the ratios of c<strong>PARP</strong> (cleaved <strong>PARP</strong>) to actin <strong>are</strong> presented in the graphs besides the parts, while the ratios of LC3-ii to actin <strong>are</strong><br />

shown below the parts. The ratios <strong>are</strong> representatives of three experiments.<br />

silencing of both LC3 <strong>and</strong> beclin 1 resulted in a greater inhibitory<br />

effect on <strong>PARP</strong> cleavage (Sup. Fig. 6A) comp<strong>are</strong>d to the singular<br />

beclin 1 perturbation. These findings were further confirmed in<br />

the PAE-Bcl-x L<br />

cells (Sup. Fig. 6B <strong>and</strong> C). In addition to inhibiting<br />

the activation of caspase 3, knockdown of either LC3 or beclin<br />

1 also affected the amount of <strong>RIP</strong>, suggesting that <strong>RIP</strong> is closely<br />

associated with the autophagic process (Sup. Fig. 6B).<br />

ATG5 is an essential protein <strong>required</strong> <strong>for</strong> <strong>autophagy</strong> at the stage<br />

of autophagosome precursor synthesis <strong>and</strong> deletion of the ATG5<br />

gene in yeast or mammalian cells usually blocks <strong>autophagy</strong> efficiently.<br />

57 In addition, ATG5 has been observed to be implicated<br />

in the apoptotic process. 25 There<strong>for</strong>e, the possible involvement<br />

of ATG5 in the C42-dependent cell death was also investigated.<br />

Indeed, silencing of the ATG5 gene with siRNA remarkably<br />

reduced both the basal <strong>and</strong> C42-stimulated accumulation of<br />

LC3-II (Fig. 8A). The ATG5 lacking cells markedly decreased<br />

<strong>PARP</strong>-1 cleavage <strong>and</strong> were partially resistant to the C42-<strong>induced</strong><br />

cell viability loss (Fig. 8A <strong>and</strong> B). In addition, these cells were<br />

defective in PAR <strong>for</strong>mation when exposed to C42. There<strong>for</strong>e,<br />

we assumed that the C42-<strong>induced</strong> <strong>autophagy</strong> could be associated<br />

with either the caspase-dependent or -independent apoptosis,<br />

likely with the involvement of ATG5 in both processes. These<br />

findings <strong>are</strong> consistent with previous reports that, in addition<br />

to its canonical function in <strong>autophagy</strong>, ATG5 is connected to a<br />

pathway that activates the apoptotic cascade.<br />

Although p62/SQSTM1 is considered as a selective substrate<br />

of <strong>autophagy</strong>, as a multidomain protein it has been found to interact<br />

with LC3. 58,59 Hence, the signaling adaptor p62 may actually<br />

participate in <strong>autophagy</strong>. Furthermore, p62 was reported to associate<br />

with the polyubiquitinated caspase 8 <strong>and</strong> trigger the extrinsic<br />

apoptosis pathway. 60 To examine whether p62 is involved in<br />

the C42-<strong>induced</strong> LC3 conversion <strong>and</strong> cell death, the cells were<br />

transfected with p62 siRNA. The singular perturbation of the<br />

p62 gene expression resulted in a decrease in LC3-II <strong>and</strong> cleavage<br />

of <strong>PARP</strong> (Fig. 8C) <strong>and</strong> resistance to the C42-triggered cell<br />

death (Fig. 8D). To test whether the cell death in either the LC3-<br />

or beclin 1-depleted setting is due to the accrual of p62, LC3<br />

<strong>and</strong> beclin 1 were each knocked down in combination with p62.<br />

Although singular knockdown of each gene offered significant<br />

protection <strong>for</strong> the cells, double-perturbation of LC3 <strong>and</strong> p62 or<br />

beclin 1 <strong>and</strong> p62, was unable to provide greater protection <strong>for</strong> the<br />

cells than p62 solitary silencing (Fig. 9A <strong>and</strong> B). Noticeably, the<br />

www.l<strong>and</strong>esbioscience.com Autophagy 9


©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

Donotdistribute.<br />

The ratios of PAR <strong>and</strong> c<strong>PARP</strong> to actin <strong>are</strong> presented in the graphs (left: c<strong>PARP</strong> to actin; right: PAR to actin). (B) MTS assays were used to analyze cell<br />

Figure 8. p62 is involved in both accumulation of LC3 <strong>and</strong> cell death <strong>induced</strong> <strong>by</strong> C42. (A) HCT<strong>11</strong>6 cells were transfected with the control (Mock) or<br />

ATG5 siRNA; after 36 h of transfection, cells were treated with C42 (0.25 μM) <strong>for</strong> 12 h be<strong>for</strong>e immunoblotting analysis with the indicated antibodies.<br />

viability after depletion of ATG5. (C) HCT<strong>11</strong>6 cells were transfected with the control (Mock) or p62 siRNA; after 36 h of transfection, cells were treated<br />

with C42 (0.25 μM) <strong>for</strong> 12 h be<strong>for</strong>e immunoblotting analysis with the indicated antibodies. The ratios LC3-ii or c<strong>PARP</strong> to actin <strong>are</strong> presented below the<br />

parts. (D) MTS assays were per<strong>for</strong>med to analyze cell viability after silencing of p62.<br />

combined knockdown of LC3 <strong>and</strong> p62 offered a greater inhibitory<br />

effect on <strong>PARP</strong>-1 cleavage but resulted in the most increase<br />

in PAR <strong>for</strong>mation (Fig. 9C); in contrast, double-depletion of<br />

beclin 1 <strong>and</strong> p62 did not show additional inhibition on <strong>PARP</strong><br />

cleavage, but resulted in a higher suppression of <strong>PARP</strong> activation<br />

(at least in a short period) (Fig. 9D). Although knockdown<br />

of either LC3 or beclin 1 inhibited the cleavage of <strong>PARP</strong>, they<br />

affected the activation of <strong>PARP</strong> in different manners (Fig. 9C<br />

<strong>and</strong> D). Unlike LC3, knockdown of beclin 1 alone or in combination<br />

with p62 may temporally inhibit the caspase-independent<br />

apoptotic or necrotic pathway. Notably, it seems that there is an<br />

inverse correlation between the C42-dependent <strong>PARP</strong> cleavage<br />

<strong>and</strong> activation under <strong>autophagy</strong>-defective circumstances. Taken<br />

together, the above findings suggest that p62 is involved in<br />

mediating the C42-dependent LC3 processing <strong>and</strong> participates<br />

in cell death under this scenario, which is consistent with a previous<br />

report on p62 in activating the caspase pathways. 60 Since<br />

knockdown of the autophagic genes cannot completely prevent<br />

the C42-<strong>induced</strong> cell death, we assume that blocking <strong>autophagy</strong><br />

only temporarily slows down or delays caspase-dependent apoptosis<br />

<strong>and</strong> may switch the cell to caspase-independent or necrosislike<br />

death.<br />

Discussion<br />

Despite having long been recognized as the inducers of apoptosis<br />

<strong>and</strong> necrosis, the ETPs’ roles in <strong>autophagy</strong> remain unclear<br />

<strong>and</strong> the interplay between different cell death pathways <strong>induced</strong><br />

Figure 9 (See opposite page). Combined silencing of LC3 <strong>and</strong> p62 switch the mode of cell death. (A) HCT<strong>11</strong>6 cells were transfected with the control<br />

(Mock) or a variety of siRNAs (singular or combined), after transfection <strong>for</strong> 36 h, cells were treated with C42 (0.25 μM) <strong>for</strong> 24 h be<strong>for</strong>e immunoblotting<br />

analysis with anti-LC3 antibody. (B) MTS assays were per<strong>for</strong>med to detect viability of cells (A). The data were statistically analyzed <strong>by</strong> using Student’s<br />

t-test. It revealed that the difference between the control siRNA <strong>and</strong> siRNAs of a variety of genes was significant. *indicates significant difference<br />

between the mock <strong>and</strong> various groups at level p < 0.05; **means significant difference at level p < 0.01. HCT<strong>11</strong>6 cells were transfected with the control<br />

(Mock), LC3 (C), Beclin 1 (B) or p62 (C <strong>and</strong> D) siRNAs solely or combined, after 36 h of transfection, cells were treated with C42 (0.25 μM) <strong>for</strong> upon to 24<br />

h be<strong>for</strong>e immunoblotting analysis with the indicated antibodies. The ratios of PAR <strong>and</strong> <strong>PARP</strong> to actin were shown in the graph right of <strong>and</strong> below the<br />

parts in (C <strong>and</strong> D), respectively. The ratios <strong>are</strong> representatives of two similar experiments.<br />

10 Autophagy Volume 7 Issue 6


©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

Donotdistribute.<br />

www.l<strong>and</strong>esbioscience.com Autophagy <strong>11</strong>


have found that the activation of <strong>PARP</strong> plays a regulatory role<br />

in apoptosis induction, revealing that the <strong>PARP</strong> inhibitor 3-AB<br />

provides transient protection <strong>by</strong> switching necrosis to apoptosis<br />

in 1-methyl-3-nitro-1-nitrosoguanidinium (MNNG)-treated<br />

cells. In contrast to MNNG, preincubation with 3-AB results in<br />

decreased membrane blebbing <strong>and</strong> reduced <strong>for</strong>mation of apoptotic<br />

bodies in melphalan-treated HL60 cells. 68 A recent study<br />

has shown that the inhibition of <strong>PARP</strong>-1 leads to prevention of<br />

the NAD + depletion <strong>and</strong> resistance against paclitaxel-<strong>induced</strong><br />

caspase 3 activation <strong>and</strong> apoptotic cell death <strong>by</strong> activating the<br />

PtdIns3K/Akt pathway. 69 We have noted that the C42-<strong>induced</strong><br />

<strong>for</strong>mation of PAR was cell-type-dependent <strong>and</strong> the inhibition of<br />

<strong>PARP</strong> also yielded diverse results in a context-dependent manner.<br />

Nevertheless, perturbation of some autophagic genes obviously<br />

reduced the <strong>for</strong>mation of PAR in all C42-treated cells, suggesting<br />

that <strong>autophagy</strong> may have a feedback effect on the activation<br />

of <strong>PARP</strong>. In other words, <strong>autophagy</strong> may also be involved in the<br />

C42-triggered caspase-independent or necrosis-like cell death.<br />

The pan-caspase inhibitor Z-V-FMK can block apoptotic<br />

cell death, but enhance death receptor-mediated necrotic cell<br />

death. 62 Wu <strong>and</strong> Shen reported that <strong>autophagy</strong> plays a protective<br />

role in Z-V-FMK-<strong>induced</strong> necrotic cell death. 16 We found that<br />

Z-V-FMK alone increased the level of p62 but decreased that of<br />

LC3-II. However, it failed to block the C42-dependent degra-<br />

©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

dation of p62. There<strong>for</strong>e, we propose that Z-V-FMK enhances<br />

<strong>by</strong> the ETPs has not been explored. In this study, the <strong>for</strong>mation<br />

of autophagosomes was identified <strong>by</strong> fluorescence <strong>and</strong><br />

electron microscopy upon treatment of the targeted cells with<br />

<strong>11</strong>'-deoxyverticillin A (C42). The EM photographs revealed the<br />

presence of the autophagic vacuoles <strong>and</strong> the characteristics of<br />

apoptosis. The C42-<strong>induced</strong> <strong>autophagy</strong> was further confirmed<br />

<strong>by</strong> immunoblotting <strong>for</strong> lipidation of LC3 <strong>and</strong> <strong>by</strong> monitoring the<br />

autophagic flux <strong>and</strong> was found to be closely associated with the<br />

activation of caspases <strong>and</strong> <strong>PARP</strong>, consistent with the reports in<br />

the literature that the accumulation of autophagic vacuoles can<br />

occur be<strong>for</strong>e cell death. 61,62 Our data demonstrated that the C42-<br />

<strong>induced</strong> <strong>autophagy</strong> preceded the caspase-dependent apoptosis in<br />

the cells tested. Inhibition of apoptosis enhanced the induction of<br />

<strong>autophagy</strong> markers but did not completely prevent the cell death<br />

from happening, whereas suppression of <strong>autophagy</strong> <strong>by</strong> chemical<br />

inhibitors or silencing the essential autophagic genes partially<br />

prevented or delayed the cleavage of caspases <strong>and</strong> <strong>PARP</strong>-1. Thus,<br />

we assumed that the C42-<strong>induced</strong> <strong>autophagy</strong>, once overactivated,<br />

could independently lead to cell death in addition to association<br />

with apoptotic cell death. We found that both <strong>PARP</strong> activation<br />

<strong>and</strong> <strong>RIP</strong>-1 expression were implicated in the C42-<strong>induced</strong> autophagic<br />

process, in agreement with previous reports in reference<br />

45, 46 <strong>and</strong> 63. It is interesting to note that <strong>RIP</strong>-1 also plays a<br />

role in mediating the cleavage <strong>and</strong> activation of <strong>PARP</strong> caused <strong>by</strong><br />

C42. We also found that <strong>RIP</strong>-3, another <strong>RIP</strong> kinase participating<br />

in mediation of apoptosis <strong>and</strong> necrosis, 51,52 also played a role<br />

in the C42-dependent <strong>autophagy</strong> <strong>and</strong> apoptosis. Considering<br />

these results, we assume that both apoptosis <strong>and</strong> <strong>autophagy</strong> may<br />

cooperate to result in cell death.<br />

the C42-<strong>induced</strong> autophagic process, which may eventually lead<br />

to cell death. In addition, we assumed that, once the apoptotic<br />

pathway was inhibited, the cells would shift back to autophagic<br />

Donotdistribute.<br />

cell death. Furthermore, the overactivation of <strong>PARP</strong> caused <strong>by</strong><br />

As the founding member of the <strong>PARP</strong> family, 64 <strong>PARP</strong>-1 is<br />

implicated in the apoptotic pathway <strong>and</strong> its cleavage can be<br />

evaluated to confirm the activation of caspase 3. In addition,<br />

<strong>PARP</strong>-1 can convert β-nicotinamide adenine dinucleotide<br />

(NAD + ) to polymers of PAR, which participate in regulating<br />

nuclear homeostasis upon DNA damage stress. 65 Such damage<br />

can cause overactivation of <strong>PARP</strong>, leading to the depletion of<br />

intracellular NAD + , suppression of ATP production <strong>and</strong> finally,<br />

cell death, especially the necrotic cell death. 10 However, several<br />

studies have shown that <strong>PARP</strong> activation also occurs during<br />

apoptosis <strong>and</strong> inhibition of PAR <strong>for</strong>mation impairs the activation<br />

of apoptotic machinery, resulting in release of apoptosis-inducing<br />

factor (AIF). <strong>11</strong>,63 Here, we clearly showed that both the PAR <strong>for</strong>mation<br />

<strong>and</strong> C42-<strong>induced</strong> <strong>autophagy</strong> preceded the appearance of<br />

hallmarks of apoptosis <strong>and</strong> inhibitors of <strong>autophagy</strong> or silencing<br />

of essential autophagic genes suppressed both the cleavage <strong>and</strong><br />

activation of <strong>PARP</strong>. Thus, <strong>autophagy</strong> app<strong>are</strong>ntly can act as a<br />

partner in the C42-dependent apoptotic cell death. Interestingly,<br />

chemical inhibition of <strong>PARP</strong> activation not only decreased the<br />

content of PAR, but also significantly attenuated caspase 3 activation,<br />

suggesting that the caspase-dependent <strong>and</strong> -independent<br />

apoptosis could be triggered <strong>by</strong> <strong>PARP</strong> overactivation upon stimulation<br />

<strong>by</strong> C42. Additionally, overactivated <strong>PARP</strong> <strong>induced</strong> <strong>by</strong><br />

C42 could also cause the necrotic cell death. In HaCaT cells,<br />

inhibition of <strong>PARP</strong> influences the mode of sulfur mustard<strong>induced</strong><br />

cell death. 66 Induction of apoptosis <strong>by</strong> <strong>PARP</strong> inhibitors<br />

is also observed in HeLa cells. 67 Pogrebniak <strong>and</strong> Hasmann<br />

Z-V-FMK could switch the cells to caspase-independent or type<br />

III cell death upon treatment with C42.<br />

Nec-1, a specific inhibitor of <strong>RIP</strong>-1 kinase, inhibits programmed<br />

necrosis (necroptosis). 50 Results from a recent study<br />

showed that the presence of Nec-1 shifts shikonin-<strong>induced</strong><br />

necroptosis to apoptosis. 70 However, we found that Nec-1 only<br />

partially decreased the LC3-II content <strong>and</strong> inhibited caspase<br />

cleavage <strong>and</strong> <strong>PARP</strong> activation <strong>induced</strong> <strong>by</strong> C42. Reduced <strong>RIP</strong>-1<br />

expression <strong>by</strong> RNAi also decreased the C42-<strong>induced</strong> LC3-II<br />

amount, <strong>PARP</strong> cleavage <strong>and</strong> cell death. Thus, C42 may act to<br />

directly affect <strong>autophagy</strong>, apoptosis <strong>and</strong> necrosis via the <strong>RIP</strong>-1<br />

protein or the <strong>RIP</strong>-1 complex <strong>for</strong>med with other proteins,<br />

although the precise mechanism(s) remains to be elucidated. In<br />

other words, <strong>RIP</strong>-1 or its complex is <strong>required</strong> <strong>for</strong> C42-<strong>induced</strong><br />

apoptosis <strong>and</strong> Nec-1 may indirectly inhibit caspase-dependent<br />

apoptosis via the <strong>RIP</strong>-1 kinase. These findings <strong>are</strong> consistent<br />

with the notion that <strong>RIP</strong>-1 is a diversified functional protein,<br />

participating in regulation of apoptosis, necrosis <strong>and</strong> autophagic<br />

cell death. 71,72<br />

One intriguing finding of this study is that suppression of<br />

<strong>autophagy</strong> <strong>by</strong> chemical inhibitors or RNAi knockdown of the<br />

critical autophagic genes only partially inhibits the cleavage <strong>and</strong><br />

activation of <strong>PARP</strong>. Conversely, inhibition of <strong>PARP</strong> activation<br />

also affects the cellular autophagic process. Thus, these findings<br />

suggest that a close link among different types of cell death may<br />

exist. We further assume that there may be a transfer between<br />

different modes of cell death <strong>induced</strong> <strong>by</strong> C42, as knockdown<br />

12 Autophagy Volume 7 Issue 6


of both LC3 <strong>and</strong> p62 completely suppressed <strong>PARP</strong> cleavage <strong>and</strong><br />

maximally augmented PAR <strong>for</strong>mation. Accordingly, cell death<br />

may transfer from the caspase-dependent mode to caspase-independent<br />

apoptosis or necrosis under this scenario. Singular gene<br />

silencing merely provided the cell with partial protection, indicting<br />

that other related genes could have taken over the role to<br />

execute <strong>and</strong> complete the death process when one protein was<br />

missing <strong>and</strong> the stimulus continued to be presented. The above<br />

hypothesis was confirmed <strong>by</strong> combined gene silencing. Depletion<br />

of LC3 or beclin 1 usually concurred with defective degradation<br />

of p62, thus, the cells might utilize p62 to relay the death signaling<br />

in the continual presence of C42.<br />

ROS <strong>are</strong> highly reactive oxygen free radicals or nonradical<br />

molecules generated <strong>by</strong> multiple mechanisms in cells <strong>and</strong> the nicotinamide<br />

adenine dinucleotide phosphate (NADPH) oxidases<br />

<strong>and</strong> mitochondria <strong>are</strong> their major cellular sources. 73 To date, ROS<br />

<strong>are</strong> widely recognized as being implicated in various biological<br />

functions <strong>and</strong> diseases. 74 Depending on the cellular context, ROS<br />

can either promote cell survival or cell death. 54 Growing evidence<br />

indicates that ROS control <strong>autophagy</strong> in multiple contexts <strong>and</strong><br />

cell types <strong>and</strong> changes in ROS <strong>and</strong> <strong>autophagy</strong> regulation contribute<br />

to cancer commencement <strong>and</strong> development. 53-56 Since<br />

the ETPs were reported to generate ROS via redox cycling, the<br />

regulatory effect of C42 on <strong>autophagy</strong> <strong>and</strong> apoptosis is likely to<br />

associate with those important signaling molecules. There<strong>for</strong>e,<br />

(2762). Anti-PAR (551813) <strong>and</strong> anti-<strong>RIP</strong>-1 (610458) antibody<br />

were purchased from BD Pharmingen; antibodies to β-actin (sc-<br />

1616), p62 (sc-28359), GFP (sc-81045) <strong>and</strong> Beclin 1 (sc-<strong>11</strong>427)<br />

were purchased from Santa Cruz Biotechnology. Anti-<strong>RIP</strong>-3<br />

(ab56164) antibody was from Abcam.<br />

Fungal material <strong>and</strong> extracts preparation. The strain of<br />

the Cordyceps-colonizing fungus was isolated <strong>by</strong> one of the<br />

authors (X.L.) from the samples of Cordyceps sinensis (Berk.)<br />

Sacc. collected in Linzhi, Tibet, in March, 2004. The fungus<br />

was identified as Gliocladium sp. <strong>and</strong> assigned the accession No.<br />

XZC04-CC-302 in X.L.’s culture collection at the Institute of<br />

Microbiology, Chinese Academy of Sciences, Beijing. The crude<br />

extract was prep<strong>are</strong>d as described previously in reference 75.<br />

Plasmids <strong>and</strong> siRNAs. The GFP-LC3 plasmid is a kind gift of<br />

Dr. Tamotsu Yoshimori (Osaka University, Japan).<br />

The siRNA specific <strong>for</strong> human MAP1LC3 (sc-43390),<br />

<strong>PARP</strong>-1 (sc-29437), <strong>RIP</strong>-1 (sc-36426), <strong>RIP</strong>-3 (sc-61482), ATG5<br />

(sc-41445) <strong>and</strong> p62 (sc-29679) were purchased from Santa Cruz<br />

Biotechnology along with the control siRNA. siGENOME<br />

SMART pool BECN1 (Dharmacon, 8678) targeting beclin 1 was<br />

bought from Dharmacon.<br />

Cell culture <strong>and</strong> western blot analysis. St<strong>and</strong>ard single-cell<br />

cloning <strong>and</strong> G418 selection procedures were used to establish<br />

stable cell lines of porcine aortic endothelial cells, that express<br />

©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

wild-type (WT) EGFR as described previously in reference 76.<br />

underst<strong>and</strong>ing of the mechanisms of C42-<strong>induced</strong> <strong>autophagy</strong> via<br />

ROS generation will provide significant implications <strong>for</strong> development<br />

of the ETPs into therapeutic anticancer agents.<br />

In summary, the ETPs <strong>are</strong> an interesting class of fungal<br />

The WT GFP-EGFR- <strong>and</strong> GFP-Bcl-x L<br />

-expressing PAE cell<br />

lines were grown in F12 medium containing 10% fetal bovine<br />

serum, antibiotics <strong>and</strong> glutamine <strong>and</strong> supplemented with<br />

Donotdistribute.<br />

G418. The human adenocarcinoma of the colon cells SW480<br />

toxins that inhibit farnesyl transferase, an enzyme <strong>required</strong> <strong>for</strong><br />

normal functioning of the Ras oncogene. Compounds sharing<br />

such a property have been explored as putative anticancer agents.<br />

Here, we demonstrate that <strong>11</strong>'-deoxyverticillin A (C42) can activate<br />

<strong>and</strong> enhance the cellular autophagic process in addition to<br />

triggering apoptosis. These results provide new insights into the<br />

function <strong>and</strong> the mode of action of C42, allowing us to further<br />

underst<strong>and</strong> the molecular mechanisms <strong>by</strong> which C42 exerts its<br />

anticancer activity. Future work in this direction could yield<br />

valuable in<strong>for</strong>mation in the development of these compounds<br />

into effective cancer therapeutic strategies through modulation<br />

of <strong>autophagy</strong>.<br />

Materials <strong>and</strong> Methods<br />

Chemicals <strong>and</strong> antibodies. <strong>11</strong>'-Deoxyverticillin A was isolated<br />

from the solid-substrate fermentation culture of the Cordycepscolonizing<br />

fungus Gliocladium sp. The following reagents were<br />

purchased from Sigma-Aldrich: rapamycin (R0395), trehalose<br />

(T9531), 3-methyladenine (M9281), necrostatin-1 (N9037),<br />

3-aminobenzamide (A0788), N-acetyl-L-cysteine (NAC,<br />

A7250), isoquinolinediol (I138) <strong>and</strong> polyclonal antibodies against<br />

LC3 (L7543) <strong>and</strong> ATG5 (A0731). Z-VAF-FMK (FMK001)<br />

was purchased from R & D systems. The following antibodies<br />

were from Cell Signaling Technology: caspase 3 (9662), cleaved<br />

caspase 3 (9664), caspase 9 (9508), <strong>PARP</strong> (9542), p70S6K<br />

(Thr389; 9205), phospho-mTOR (Ser2448; 2971) <strong>and</strong> Bcl-x L<br />

were maintained in DMEM (HyClone, SH30022. 01B) supplemented<br />

with 10% fetal bovine serum (GIBCO, 16000), plus<br />

antibiotics. Human adenocarcinoma of the colon cells SW620<br />

were maintained in RPMI-1640-medium (HyClone, SH30809.<br />

01B), whereas human colon carcinoma cells HCT<strong>11</strong>6 were<br />

grown in McCoy’s 5A (KeyGEN, KGM5ASY) supplemented<br />

with FBS <strong>and</strong> antibiotics. Cells were split overnight <strong>and</strong> grown<br />

to 50% confluency be<strong>for</strong>e adding <strong>11</strong>'-dideoxyverticillin A<br />

(C42).<br />

For siRNA interference, cells were grown to 40% confluence<br />

in their respective medium without antibiotics <strong>and</strong> transfected<br />

using DharmaFECT (Dharmacon, T2001) according to the<br />

manufacturer’s instructions. After transfection <strong>for</strong> 36 h, cells<br />

were directly treated with C42 without being split. Whole cell<br />

lysates were prep<strong>are</strong>d with lysis using Triton X-100/glycerol buffer,<br />

which contained 50 mM Tris-HCl, pH 7.4, 4 mM EDTA,<br />

2 mM EGTA <strong>and</strong> 1 mM dithiothreitol, supplemented with 1%<br />

Triton X-100 <strong>and</strong> protease inhibitors <strong>and</strong> then separated on a<br />

SDS-PAGE gel (15, 10 or 8% according to the molecular weights<br />

of the proteins of interest) <strong>and</strong> transferred to PVDF membrane.<br />

Western blotting was per<strong>for</strong>med using appropriate primary antibodies<br />

<strong>and</strong> horseradish peroxidase-conjugated suitable secondary<br />

antibodies, followed <strong>by</strong> detection with enhanced chemiluminescence<br />

(Pierce Chemical, 34080).<br />

Several x-ray films were analyzed to verify the linear range of<br />

the chemiluminescence signals <strong>and</strong> the quantifications were carried<br />

out using densitometry.<br />

www.l<strong>and</strong>esbioscience.com Autophagy 13


Cell proliferation assay (MTS). Cells were plated in 96-well<br />

plates (10,000 cells per well) in 100 μl complete culture medium.<br />

After overnight culture, the medium was replaced with complete<br />

medium that was either drug-free or contained C42 or other<br />

chemicals. The cells were cultured <strong>for</strong> various periods <strong>and</strong> cellular<br />

viability was determined with CellTiter 96 ® Aqueous Non-<br />

Radioactive Cell Proliferation Assay (G<strong>11</strong><strong>11</strong>, Promega).<br />

For siRNA, each well was plated with 5,000 cells. After 36 h<br />

of transfection, C42 was directly added to the cells <strong>and</strong> cellular<br />

viability was determined.<br />

Flow cytometry assay. HCT<strong>11</strong>6 cells were treated with C42<br />

(0.25 <strong>and</strong> 0.5 μM, respectively), trypsinized <strong>and</strong> harvested<br />

(keeping all floating cells), washed with PBS buffer, followed <strong>by</strong><br />

incubating with a fluorescein isothiocyanate-labeled annexin V<br />

(FITC) <strong>and</strong> propidium iodide (PI) according to the instructions<br />

of an Annexin-V-FITC Apoptosis Detection Kit (Biovision Inc.,<br />

K101-100) <strong>and</strong> analyzed <strong>by</strong> flow cytometry (FACSAria, Becton<br />

Dickinson). Percentages of the cells with annnexin V-positive<br />

<strong>and</strong> PI-negative stainings were considered as apoptotic, whereas<br />

PI-positive staining was considered to be necrotic.<br />

Colony growth assay. Cells were treated with C42 <strong>for</strong> 6 h,<br />

split, seeded at 300 cells/ml <strong>and</strong> cultured <strong>for</strong> two weeks to allow<br />

colony growth of surviving cells.<br />

Fluorescence microscopy. HCT<strong>11</strong>6 cells were transfected with<br />

the GFP-LC3 expressing plasmid <strong>and</strong> after 24 h, cells were treated<br />

according to the manufacturer’s protocol <strong>and</strong> its integrity was<br />

confirmed <strong>by</strong> electrophoresis on ethidium bromide-stained 1%<br />

agarose gel. Total cellular RNA (1 μg) was reverse transcribed<br />

at 37°C <strong>for</strong> 15 min in 20 μl of PrimeScript TM RT reagent Kit<br />

(TaKaRa, DRR037A). Reactions were stopped <strong>by</strong> heat inactivation<br />

<strong>for</strong> 5 s at 85°C. Primer sequences used <strong>for</strong> amplification were<br />

as fellows: LC3 upstream primer, 5'-ATG CCG TCG GAG AAG<br />

ACC-3', downstream primer, 5'-TTA CAC TGA CAA TTT<br />

CAT CC-3'; <strong>PARP</strong>-1 upstream primer, 5'-CTA AAG GCT CAG<br />

AAC GAC C-3', downstream primer, 5'-GAA GGA GGG CAC<br />

CGA ACA-3'; β-actin upstream primer, 5'-GCC TGA CGG<br />

CCA GGT CAT CAC-3', downstream primer, 5'-CGG ATG<br />

TCC ACG TCA CAC TTC-3'. PCR reactions with LC3 primers<br />

were initiated with a 5 min denaturation at 95°C in a final<br />

volume of 20 μl. The cycle profile was 95°C, 60°C <strong>and</strong> 72°C <strong>for</strong><br />

30 s <strong>for</strong> up to 25 cycles. The PCR amplification products were<br />

determined in DNA agarose gels, which were photographed <strong>and</strong><br />

quantitatively scanned using Photoshop softw<strong>are</strong>. The expression<br />

of β-actin was used as the internal control.<br />

Statistical analysis. Normally distributed data <strong>are</strong> shown as<br />

mean ± SD <strong>and</strong> were analyzed using one-way analysis of variance<br />

<strong>and</strong> the Student-Newman-Keuls post-hoc test.<br />

©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

Acknowledgements<br />

We thank Alex<strong>and</strong>er Sorkin from UCHSC <strong>for</strong> providing EGFR<br />

plasmid, Dr. Tamotsu Yoshimori from Osaka University <strong>for</strong> the<br />

GFP-LC3 plasmid, Dr. Quan Chen from the Institute of Zoology<br />

<strong>for</strong> the Bcl-x L<br />

plasmid <strong>and</strong> Dr. Xin Ye from the Institute of<br />

Donotdistribute.<br />

Microbiology, CAS <strong>for</strong> p53 -/- HCT<strong>11</strong>6 cells. We <strong>are</strong> also grateful<br />

with C42, the fluorescence of GFP-LC3 was viewed <strong>and</strong> the<br />

images were acquired via confocal microscopy (Leica, TCS SP5).<br />

Electron microscopy. Electron microscopy was per<strong>for</strong>med as<br />

described previously in reference 77. Briefly, HCT<strong>11</strong>6 cell samples<br />

were washed three times with PBS, trypsinized <strong>and</strong> collected<br />

<strong>by</strong> centrifuging. The cell pellets were fixed with 4% para<strong>for</strong>maldehyde<br />

overnight at 4°C, post-fixed with 1% OsO 4<br />

in cacodylate<br />

buffer at room temperature <strong>for</strong> 1 h <strong>and</strong> dehydrated stepwise<br />

with ethanol. The dehydrated pellets were rinsed with propylene<br />

oxide <strong>for</strong> 30 min <strong>and</strong> then embedded in Spurr resin <strong>for</strong> sectioning.<br />

Images of thin sections were observed under a transmission<br />

electron microscope (JEM1230, Tokyo, Japan).<br />

RNA extraction <strong>and</strong> RT-PCR analysis. Total cellular RNA<br />

was extracted using TRIzol reagent (Invitrogen, 15596-018)<br />

to Dr. Daniel J. Klionsky from the University of Michigan <strong>for</strong> his<br />

advice in preparing the manuscript. This study was supported <strong>by</strong><br />

grants from the National Natural Science Foundation of China<br />

(NSFC, grants 90408024, 30900731, 30672097, 30870057 <strong>and</strong><br />

30925039) <strong>and</strong> the Ministry of Science <strong>and</strong> Technology of China<br />

(2009CB522302 <strong>and</strong> 2009ZX09302-004).<br />

Note<br />

Supplemental materials can be found at:<br />

www.l<strong>and</strong>esbioscience.com/journals/<strong>autophagy</strong>/article/15103<br />

References<br />

1. Klionsky DJ, Emr SD. Autophagy as a regulated pathway<br />

of cellular degradation. Science 2000; 290:1717-21.<br />

2. Mizushima N. Autophagy: process <strong>and</strong> function. Genes<br />

Dev 2007; 21:2861-73.<br />

3. Levine B, Klionsky DJ. Development <strong>by</strong> self-digestion:<br />

molecular mechanisms <strong>and</strong> biological functions of<br />

<strong>autophagy</strong>. Dev Cell 2004; 6:463-77.<br />

4. Codogno P, Meijer AJ. Autophagy <strong>and</strong> signaling: their<br />

role in cell survival <strong>and</strong> cell death. Cell Death Differ<br />

2005; 12:1509-18.<br />

5. Edinger AL, Thompson CB. Death <strong>by</strong> design: apoptosis,<br />

necrosis <strong>and</strong> <strong>autophagy</strong>. Curr Opin Cell Biol 2004;<br />

16:663-9.<br />

6. Levine B, Kroemer G. Autophagy in the pathogenesis<br />

of disease. Cell 2008; 132:27-42.<br />

7. Jacobson MD, Weil M, Raff MC. Programmed cell<br />

death in animal development. Cell 1997; 88:347-54.<br />

8. Levine B, Yuan J. Autophagy in cell death: an innocent<br />

convict? J Clin Invest 2005; <strong>11</strong>5:2679-88.<br />

9. Mizushima N, Levine B, Cuervo AM, Klionsky DJ.<br />

Autophagy fights disease through cellular self-digestion.<br />

Nature 2008; 451:1069-75.<br />

10. Ha HC, Snyder SH. Poly(ADP-ribose) polymerase is a<br />

mediator of necrotic cell death <strong>by</strong> ATP depletion. Proc<br />

Natl Acad Sci USA 1999; 96:13978-82.<br />

<strong>11</strong>. Yu SW, Wang H, Poitras MF, Coombs C, Bowers WJ,<br />

Federoff HJ, et al. Mediation of poly(ADP-ribose)<br />

polymerase-1-dependent cell death <strong>by</strong> apoptosis-inducing<br />

factor. Science 2002; 297:259-63.<br />

12. Baehrecke EH. How death shapes life during development.<br />

Nat Rev Mol Cell Biol 2002; 3:779-87.<br />

13. Budihardjo I, Oliver H, Lutter M, Luo X, Wang X.<br />

Biochemical pathways of caspase activation during<br />

apoptosis. Annu Rev Cell Dev Biol 1999; 15:269-90.<br />

14. Baehrecke EH. Autophagy: dual roles in life <strong>and</strong> death?<br />

Nat Rev Mol Cell Biol 2005; 6:505-10.<br />

15. Luo S, Rubinsztein DC. Apoptosis blocks Beclin<br />

1-dependent autophagosome synthesis: an effect rescued<br />

<strong>by</strong> Bcl-x L<br />

. Cell Death Differ 2010; 17:268-77.<br />

16. Wu YT, Tan HL, Huang Q, Kim YS, Pan N, Ong<br />

WY, et al. Autophagy plays a protective role during<br />

zVAD-<strong>induced</strong> necrotic cell death. Autophagy 2008;<br />

4:457-66.<br />

17. Yu L, Alva A, Su H, Dutt P, Freundt E, Welsh S,<br />

et al. Regulation of an ATG7-beclin 1 program of<br />

autophagic cell death <strong>by</strong> caspase-8. Science 2004;<br />

304:1500-2.<br />

18. Zong WX, Ditsworth D, Bauer DE, Wang ZQ,<br />

Thompson CB. Alkylating DNA damage stimulates a<br />

regulated <strong>for</strong>m of necrotic cell death. Genes Dev 2004;<br />

18:1272-82.<br />

19. Guillon-Munos A, van Bemmelen MX, Clarke PG.<br />

Role of phosphoinositide-3-kinase in the autophagic<br />

death of serum-deprived PC12 cells. Apoptosis 2005;<br />

10:1031-41.<br />

20. Maiuri MC, Zalckvar E, Kimchi A, Kroemer G. Selfeating<br />

<strong>and</strong> self-killing: crosstalk between <strong>autophagy</strong> <strong>and</strong><br />

apoptosis. Nat Rev Mol Cell Biol 2007; 8:741-52.<br />

21. Zucchini-Pascal N, de Sousa G, Rahmani R. Lindane<br />

<strong>and</strong> cell death: at the crossroads between apoptosis,<br />

necrosis <strong>and</strong> <strong>autophagy</strong>. Toxicology 2009; 256:32-41.<br />

14 Autophagy Volume 7 Issue 6


22. Colell A, Ricci JE, Tait S, Milasta S, Maurer U,<br />

Bouchier-Hayes L, et al. GAPDH <strong>and</strong> <strong>autophagy</strong> preserve<br />

survival after apoptotic cytochrome c release in the<br />

absence of caspase activation. Cell 2007; 129:983-97.<br />

23. Crighton D, Wilkinson S, O’Prey J, Syed N, Smith P,<br />

Harrison PR, et al. DRAM, a p53-<strong>induced</strong> modulator<br />

of <strong>autophagy</strong>, is critical <strong>for</strong> apoptosis. Cell 2006;<br />

126:121-34.<br />

24. Pattingre S, Tassa A, Qu X, Garuti R, Liang XH,<br />

Mizushima N, et al. Bcl-2 antiapoptotic proteins<br />

inhibit Beclin 1-dependent <strong>autophagy</strong>. Cell 2005;<br />

122:927-39.<br />

25. Yousefi S, Perozzo R, Schmid I, Ziemiecki A, Schaffner<br />

T, Scapozza L, et al. Calpain-mediated cleavage of Atg5<br />

switches <strong>autophagy</strong> to apoptosis. Nat Cell Biol 2006;<br />

8:<strong>11</strong>24-32.<br />

26. Cho DH, Jo YK, Hwang JJ, Lee YM, Roh SA, Kim<br />

JC. Caspase-mediated cleavage of ATG6/Beclin-1 links<br />

apoptosis to <strong>autophagy</strong> in HeLa cells. Cancer Lett<br />

2009; 274:95-100.<br />

27. Scott RC, Juhasz G, Neufeld TP. Direct induction of<br />

<strong>autophagy</strong> <strong>by</strong> Atg1 inhibits cell growth <strong>and</strong> induces<br />

apoptotic cell death. Curr Biol 2007; 17:1-<strong>11</strong>.<br />

28. Eisenberg-Lerner A, Bialik S, Simon HU, Kimchi A.<br />

Life <strong>and</strong> death partners: apoptosis, <strong>autophagy</strong> <strong>and</strong><br />

the cross-talk between them. Cell Death Differ 2009;<br />

16:966-75.<br />

29. Zalckvar E, Yosef N, Reef S, Ber Y, Rubinstein AD,<br />

Mor I, et al. A systems level strategy <strong>for</strong> analyzing the<br />

cell death network: implication in exploring the apoptosis/<strong>autophagy</strong><br />

connection. Cell Death Differ 2010;<br />

17:1244-53.<br />

30. Dong JY, He HP, Shen YM, Zhang KQ. Nematicidal<br />

epipolysulfanyldioxopiperazines from Gliocladium<br />

roseum. J Nat Prod 2005; 68:1510-3.<br />

40. Mizushima N. Methods <strong>for</strong> monitoring <strong>autophagy</strong>. Int<br />

J Biochem Cell Biol 2004; 36:2491-502.<br />

41. Kabeya Y, Mizushima N, Ueno T, Yamamoto A,<br />

Kirisako T, Noda T, et al. LC3, a mammalian homologue<br />

of yeast Apg8p, is localized in autophagosome<br />

membranes after processing. EMBO J 2000; 19:5720-8.<br />

42. Klionsky DJ, Abeliovich H, Agostinis P, Agrawal DK,<br />

Aliev G, Askew DS, et al. Guidelines <strong>for</strong> the use <strong>and</strong><br />

interpretation of assays <strong>for</strong> monitoring <strong>autophagy</strong> in<br />

higher eukaryotes. Autophagy 2008; 4:151-75.<br />

43. Hay N, Sonenberg N. Upstream <strong>and</strong> downstream of<br />

mTOR. Genes Dev 2004; 18:1926-45.<br />

44. Yamamoto A, Tagawa Y, Yoshimori T, Moriyama<br />

Y, Masaki R, Tashiro Y. Bafilomycin A1 prevents<br />

maturation of autophagic vacuoles <strong>by</strong> inhibiting fusion<br />

between autophagosomes <strong>and</strong> lysosomes in rat hepatoma<br />

cell line, H-4-II-E cells. Cell Struct Funct 1998;<br />

23:33-42.<br />

45. Huang Q, Wu YT, Tan HL, Ong CN, Shen HM. A<br />

novel function of poly(ADP-ribose) polymerase-1 in<br />

modulation of <strong>autophagy</strong> <strong>and</strong> necrosis under oxidative<br />

stress. Cell Death Differ 2009; 16:264-77.<br />

46. Munoz-Gamez JA, Rodriguez-Vargas JM, Quiles-Perez<br />

R, Aguilar-Quesada R, Martin-Oliva D, de Murcia<br />

G, et al. <strong>PARP</strong>-1 is involved in <strong>autophagy</strong> <strong>induced</strong> <strong>by</strong><br />

DNA damage. Autophagy 2009; 5:61-74.<br />

47. Wieler S, Gagne JP, Vaziri H, Poirier GG, Benchimol<br />

S. Poly(ADP-ribose) polymerase-1 is a positive regulator<br />

of the p53-mediated G 1<br />

arrest response following<br />

ionizing radiation. J Biol Chem 2003; 278:18914-21.<br />

48. Ding W, Liu W, Cooper KL, Qin XJ, de Souza Bergo<br />

PL, Hudson LG, et al. Inhibition of poly(ADP-ribose)<br />

polymerase-1 <strong>by</strong> arsenite interferes with repair of oxidative<br />

DNA damage. J Biol Chem 2009; 284:6809-17.<br />

49. Yu L, Wan F, Dutta S, Welsh S, Liu Z, Freundt E,<br />

59. Pankiv S, Clausen TH, Lamark T, Brech A, Bruun<br />

JA, Outzen H, et al. p62/SQSTM1 binds directly to<br />

Atg8/LC3 to facilitate degradation of ubiquitinated<br />

protein aggregates <strong>by</strong> <strong>autophagy</strong>. J Biol Chem 2007;<br />

282:24131-45.<br />

60. Jin Z, Li Y, Pitti R, Lawrence D, Pham VC, Lill JR, et<br />

al. Cullin3-based polyubiquitination <strong>and</strong> p62-dependent<br />

aggregation of caspase-8 mediate extrinsic apoptosis<br />

signaling. Cell 2009; 137:721-35.<br />

61. Espert L, Denizot M, Grimaldi M, Robert-Hebmann<br />

V, Gay B, Varbanov M, et al. Autophagy is involved in<br />

T cell death after binding of HIV-1 envelope proteins<br />

to CXCR4. J Clin Invest 2006; <strong>11</strong>6:2161-72.<br />

62. Gonzalez-Polo RA, Boya P, Pauleau AL, Jalil A,<br />

Larochette N, Souquere S, et al. The apoptosis/<strong>autophagy</strong><br />

paradox: autophagic vacuolization be<strong>for</strong>e apoptotic<br />

death. J Cell Sci 2005; <strong>11</strong>8:3091-102.<br />

63. Yu SW, Andrabi SA, Wang H, Kim NS, Poirier GG,<br />

Dawson TM, et al. Apoptosis-inducing factor mediates<br />

poly(ADP-ribose) (PAR) polymer-<strong>induced</strong> cell death.<br />

Proc Natl Acad Sci USA 2006; 103:18314-9.<br />

64. Ame JC, Spenlehauer C, de Murcia G. The <strong>PARP</strong><br />

superfamily. Bioessays 2004; 26:882-93.<br />

65. Schreiber V, Dantzer F, Ame JC, de Murcia G.<br />

Poly(ADP-ribose): novel functions <strong>for</strong> an old molecule.<br />

Nat Rev Mol Cell Biol 2006; 7:517-28.<br />

66. Kehe K, Raithel K, Kreppel H, Jochum M, Worek F,<br />

Thiermann H. Inhibition of poly(ADP-ribose) polymerase<br />

(<strong>PARP</strong>) influences the mode of sulfur mustard<br />

(SM)-<strong>induced</strong> cell death in HaCaT cells. Arch Toxicol<br />

2008; 82:461-70.<br />

67. Ghosh U, Bhattacharyya NP. Induction of apoptosis <strong>by</strong><br />

the inhibitors of poly(ADP-ribose)polymerase in HeLa<br />

cells. Mol Cell Biochem 2009; 320:15-23.<br />

©20<strong>11</strong>L<strong>and</strong>esBioscience.<br />

68. Pogrebniak A, Schemainda I, Pelka-Fleischer R, Nussler<br />

31. Gardiner DM, Waring P, Howlett BJ. The epipolythiodioxopiperazine<br />

(ETP) class of fungal toxins: distribution,<br />

mode of action, functions <strong>and</strong> biosynthesis.<br />

Microbiology 2005; 151:1021-32.<br />

32. Bernardo PH, Brasch N, Chai CL, Waring P. A novel<br />

redox mechanism <strong>for</strong> the glutathione-dependent reversible<br />

uptake of a fungal toxin in cells. J Biol Chem 2003;<br />

278:46549-55.<br />

33. Beaver JP, Waring P. Lack of correlation between early<br />

intracellular calcium ion rises <strong>and</strong> the onset of apoptosis<br />

in thymocytes. Immunol Cell Biol 1994; 72:489-99.<br />

34. Yanagihara M, Sasaki-Takahashi N, Sugahara T,<br />

Yamamoto S, Shinomi M, Yamashita I, et al. Leptosins<br />

isolated from marine fungus Leptoshaeria species inhibit<br />

DNA topoisomerases I <strong>and</strong>/or II <strong>and</strong> induce apoptosis<br />

<strong>by</strong> inactivation of Akt/protein kinase B. Cancer Sci<br />

2005; 96:816-24.<br />

35. Chen Y, Miao ZH, Zhao WM, Ding J. The p53 pathway<br />

is synergized <strong>by</strong> p38 MAPK signaling to mediate<br />

<strong>11</strong>,<strong>11</strong>'-dideoxyverticillin-<strong>induced</strong> G 2<br />

/M arrest. FEBS<br />

Lett 2005; 579:3683-90.<br />

36. Ekert PG, Silke J, Vaux DL. Caspase inhibitors. Cell<br />

Death Differ 1999; 6:1081-6.<br />

37. Seglen PO, Gordon PB. 3-Methyladenine: specific<br />

inhibitor of autophagic/lysosomal protein degradation<br />

in isolated rat hepatocytes. Proc Natl Acad Sci USA<br />

1982; 79:1889-92.<br />

38. Heitman J, Movva NR, Hall MN. Targets <strong>for</strong> cell cycle<br />

arrest <strong>by</strong> the immunosuppressant rapamycin in yeast.<br />

Science 1991; 253:905-9.<br />

39. Sarkar S, Davies JE, Huang Z, Tunnacliffe A,<br />

Rubinsztein DC. Trehalose, a novel mTOR-independent<br />

<strong>autophagy</strong> enhancer, accelerates the clearance of<br />

mutant huntingtin <strong>and</strong> alpha-synuclein. J Biol Chem<br />

2007; 282:5641-52.<br />

et al. Autophagic programmed cell death <strong>by</strong> selective<br />

catalase degradation. Proc Natl Acad Sci USA 2006;<br />

103:4952-7.<br />

50. Degterev A, Hitomi J, Germscheid M, Ch’en IL,<br />

Korkina O, Teng X, et al. Identification of <strong>RIP</strong>-1 kinase<br />

as a specific cellular target of necrostatins. Nat Chem<br />

V, Hasmann M. Poly ADP-ribose polymerase (<strong>PARP</strong>)<br />

inhibitors transiently protect leukemia cells from alkylating<br />

agent <strong>induced</strong> cell death <strong>by</strong> three different effects.<br />

Eur J Med Res 2003; 8:438-50.<br />

69. Szanto A, Hellebr<strong>and</strong> EE, Bognar Z, Tucsek Z, Szabo<br />

Biol 2008; 4:313-21.<br />

51. Feng S, Yang Y, Mei Y, Ma L, Zhu DE, Hoti N, et al.<br />

Cleavage of <strong>RIP</strong>3 inactivates its caspase-independent<br />

apoptosis pathway <strong>by</strong> removal of kinase domain. Cell<br />

Signal 2007; 19:2056-67.<br />

52. Zhang DW, Shao J, Lin J, Zhang N, Lu BJ, Lin SC, et<br />

al. <strong>RIP</strong>3, an energy metabolism regulator that switches<br />

TNF-<strong>induced</strong> cell death from apoptosis to necrosis.<br />

Science 2009; 325:332-6.<br />

53. Huang J, Brumell JH. NADPH oxidases contribute to<br />

<strong>autophagy</strong> regulation. Autophagy 2009; 5:887-9.<br />

54. Mazure NM, Pouyssegur J. Hypoxia-<strong>induced</strong> <strong>autophagy</strong>:<br />

cell death or cell survival? Curr Opin Cell Biol<br />

2010; 22:177-80.<br />

55. Scherz-Shouval R, Shvets E, Fass E, Shorer H, Gil<br />

L, Elazar Z. Reactive oxygen species <strong>are</strong> essential <strong>for</strong><br />

<strong>autophagy</strong> <strong>and</strong> specifically regulate the activity of Atg4.<br />

EMBO J 2007; 26:1749-60.<br />

56. Zhang H, Kong X, Kang J, Su J, Li Y, Zhong J, et al.<br />

Oxidative stress induces parallel <strong>autophagy</strong> <strong>and</strong> mitochondria<br />

dysfunction in human glioma U251 cells.<br />

Toxicol Sci 2009; <strong>11</strong>0:376-88.<br />

57. Cecconi F, Levine B. The role of <strong>autophagy</strong> in mammalian<br />

development: cell makeover rather than cell death.<br />

Dev Cell 2008; 15:344-57.<br />

58. Gao Z, Gammoh N, Wong PM, Erdjument-Bromage<br />

H, Tempst P, Jiang X. Processing of autophagic protein<br />

LC3 <strong>by</strong> the 20S proteasome. Autophagy 2010; 6:126-37.<br />

Donotdistribute.<br />

A, Gallyas F Jr, et al. <strong>PARP</strong>-1 inhibition-<strong>induced</strong> activation<br />

of PI-3-kinase-Akt pathway promotes resistance<br />

to taxol. Biochem Pharmacol 2009; 77:1348-57.<br />

70. Han W, Xie J, Li L, Liu Z, Hu X. Necrostatin-1 reverts<br />

shikonin-<strong>induced</strong> necroptosis to apoptosis. Apoptosis<br />

2009; 14:674-86.<br />

71. Declercq W, V<strong>and</strong>en Berghe T, V<strong>and</strong>enabeele P. <strong>RIP</strong><br />

kinases at the crossroads of cell death <strong>and</strong> survival. Cell<br />

2009; 138:229-32.<br />

72. Festjens N, V<strong>and</strong>en Berghe T, Cornelis S, V<strong>and</strong>enabeele<br />

P. <strong>RIP</strong>-1, a kinase on the crossroads of a cell’s decision<br />

to live or die. Cell Death Differ 2007; 14:400-10.<br />

73. Trachootham D, Lu W, Ogasawara MA, Nilsa RD,<br />

Huang P. Redox regulation of cell survival. Antioxid<br />

Redox Signal 2008; 10:1343-74.<br />

74. Giannoni E, Buricchi F, Grimaldi G, Parri M, Cialdai<br />

F, Taddei ML, et al. Redox regulation of anoikis:<br />

reactive oxygen species as essential mediators of cell<br />

survival. Cell Death Differ 2008; 15:867-78.<br />

75. Liu L, Liu S, Jiang L, Chen X, Guo L, Che Y.<br />

Chloropupukeananin, the first chlorinated pupukeanane<br />

derivative <strong>and</strong> its precursors from Pestalotiopsis<br />

fici. Org Lett 2008; 10:1397-400.<br />

76. Jiang X, Huang F, Marusyk A, Sorkin A. Grb2 regulates<br />

internalization of EGF receptors through clathrin-coated<br />

pits. Molecular biology of the cell 2003; 14:858-70.<br />

77. Li M, Jiang X, Liu D, Na Y, Gao GF, Xi Z. Autophagy<br />

protects LNCaP cells under <strong>and</strong>rogen deprivation conditions.<br />

Autophagy 2008; 4:54-60.<br />

www.l<strong>and</strong>esbioscience.com Autophagy 15

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!