JVI Accepted Manuscript Posted Online 18 February 2015 J. Virol. doi:10.1128/JVI.00213-15 Copyright © 2015, American Society for Microbiology. All Rights Reserved.

1

Adenovirus replaces mitotic checkpoint controls

Roberta L. Turner,a Peter Groitl,b Thomas Dobner,b David A. Ornellesa#

Department of Microbiology and Immunology, Wake Forest School of Medicine, WinstonSalem, North Carolina, USAa; Heinrich Pette Institute, Leibniz Institute for Experimental Virology, Hamburg, Germanyb

# Address correspondence to: David A. Ornelles, [email protected]

Running Head: Adenovirus replaces mitotic checkpoints Abstract word count: 197 Manuscript word count: 7892

2 ABSTRACT 8 9

Infection with adenovirus triggers the cellular DNA damage response, elements of which include cell death and cell cycle arrest. Early adenoviral proteins including the E1B-55K and

10

E4orf3 proteins inhibit signaling in response to DNA damage. A fraction of cells infected with

11

an adenovirus mutant unable to express the E1B-55K and E4orf3 genes appeared to arrest in a

12

mitotic-like state. Cells infected early in G1 of the cell cycle were predisposed to arrest in this

13

state at late times of infection. This arrested state, which displays hallmarks of mitotic

14

catastrophe, was prevented by expression of either the E1B-55K or the E4orf3 genes. However,

15

E1B-55K-mutant virus-infected cells became trapped in a mitotic-like state in the presence of the

16

microtubule poison colcemid, suggesting that the two viral proteins restrict entry into mitosis or

17

facilitate exit from mitosis in order to prevent infected cells from arresting in mitosis. The

18

E1B-55K protein appeared to prevent inappropriate entry into mitosis through its interaction with

19

the cellular tumor suppressor protein p53. The E4orf3 protein facilitated exit from mitosis by

20

possibly mislocalizing and functionally inactivating cyclin B1. When expressed in non-infected

21

cells, E4orf3 overcame the mitotic arrest caused by the degradation-resistant R42A cyclin B1

22

variant. IMPORTANCE

23

Cells that are infected with adenovirus type 5 early in G1 of the cell cycle are predisposed

24

to arrest in a mitotic-like state in a p53-dependent manner. The adenoviral E1B-55K protein

25

prevents entry into mitosis. This newly described activity for the E1B-55K protein appears to

26

depend on the interaction between the E1B-55K protein and the tumor suppressor p53. The

27

adenoviral E4orf3 protein facilitates exit from mitosis, possibly by altering the intracellular

28

distribution of cyclin B1. By preventing entry into mitosis and by promoting exit from mitosis,

3 29

these adenoviral proteins act to prevent the infected cell from arresting in a mitotic-like state. INTRODUCTION

30

Adenoviral infection and the ensuing replication of the viral double-stranded DNA

31

genome activates the host DNA-damage response (1, 2). Early adenoviral proteins collaborate to

32

dampen this host response (reviewed in 3). The initial phase of the DNA-damage response

33

proceeds through a phosphorylation cascade, while subsequent recruitment of effector proteins

34

also depends on the conjugation of ubiquitin and the related small ubiquitin like modifier SUMO

35

(4). Signals initiated by the three apical kinases or DNA-dependent protein kinase (DNA-PK)

36

(5), ataxia telangiectasia mutated protein (ATM) (6), and ATM- and Rad3-related protein (ATR)

37

(7) trigger downstream consequences of DNA-damage such as DNA repair, cell cycle arrest, and

38

cell death. The tumor suppressor protein p53 is centrally positioned in the cellular response to

39

DNA-damage. Numerous branches of the DNA-damage response are controlled by p53

40

including cell cycle arrest, cell death, senescence, autophagy, and cell proliferation (8). Not

41

surprisingly, viruses that elicit a robust DNA-damage response inevitably target p53. For

42

adenovirus, the transcriptional activity of p53 is suppressed by the E1B-55K protein (9-11), the

43

stability of p53 is decreased by a ubiquitin protein-ligase formed by the E1B-55K and E4orf6

44

protein (12-14) and the expression of p53-responsive genes is epigenetically dampened by the

45

E4orf3 protein (15).

46

Cell cycle arrest mediated by p53 following DNA-damage typically occurs at the G1/S

47

border (16). However p53 also inhibits cell cycle progression immediately before mitosis. p53

48

can prevent entry into mitosis by inhibiting a kinesin involved in the arrangement of condensed

49

chromosomes (17). Polo-like kinase 1 (Plk1) promotes the transition from G2 into mitosis. The

50

inhibition of Plk1 uncovers p53-dependent outcomes in response to mitotic stress. In p53-

4 51

deficient cells, Plk1 inhibitors and microtubule poisons elicit mitotic catastrophe and greater

52

DNA-damage than in p53-proficient cells (18). This may reflect the absence of p53-dependent

53

apoptosis that would normally eliminate cells arrested in mitosis. It has been suggested that p53-

54

dependent cell-cycle-arrest at the G2/M border is the key factor in determining whether a cell

55

undergoes mitotic catastrophe or apoptosis (19).

56

Although progression through the cell cycle can be stopped at many stages, the intricately

57

orchestrated process of mitosis proceeds once the antephase checkpoint has been cleared or

58

bypassed (20), despite the persistence of damaged DNA (21). Mitosis is regulated by the

59

appropriate localization of cellular proteins and their timely degradation by the anaphase-

60

promoting complex or cyclosome (APC/C). During the G2 phase of the cell cycle, levels of

61

cyclin B1 rise, which associates with Cdk1 to form the major mitotic kinase (22). Entry into

62

mitosis begins with the activating phosphorylation of the Cdc25C phosphatase and components

63

of the APC/C as well as the inactivating phosphorylation of the Wee1 and Myt1 kinase by polo-

64

like kinases (23). The cyclin B1-Cdk1 complex is believed to shuttle in and out of the nucleus,

65

with hyperphosphorylation of cyclin B1 inhibiting nuclear export of the complex leading to an

66

intranuclear increase in cyclin B1-Cdk1 (24, 25). Within the nucleus, this kinase directs mitotic

67

progression by phosphorylating numerous targets (26) such as the nuclear lamins in order

68

promote nuclear envelope breakdown (27) and condensin II to initiate condensation of the

69

chromosomes (28). Exit from mitosis requires the degradation of proteins ubiquitinated by the

70

APC/C (29). Key mitotic targets of the APC/C are cyclin B1, securin, and Plk1. With the

71

degradation of cyclin B1 and securin, separase is free to cleave cohesin from sister chromatids,

72

leading to the precipitous separation of chromatids and progression out of mitosis (30).

73

Cell death by mitotic catastrophe was initially described as a caspase-dependent,

5 74

apoptotic death triggered by aberrant mitosis and the persistence of active mitotic kinases (31). It

75

has been proposed that mitotic catastrophe be considered an oncosuppressive mechanism leading

76

to cell death resulting from the disorder of mitotic machinery that is typified by a period of

77

aberrant mitotic arrest (32). Additional hallmarks of mitotic catastrophe include the formation of

78

multipolar spindles and the appearance of a cleaved form of cyclin B1 that is unable to be

79

degraded and is therefore presumed to sustain MPF activity (33, 34). Aberrant mitotic arrest

80

requires that cells enter mitosis but halt prior to anaphase. Examples of viruses that arrest cells in

81

mitosis and can elicit mitotic catastrophe include adenovirus, adenovirus-associated virus (AAV)

82

and chicken anemia virus. Expression of the adenovirus E4orf4 protein to high levels in H1299

83

cells lead to the accumulation of cells with 4N and greater DNA content followed by cell death

84

characterized as mitotic catastrophe (35). Exposure of p53-deficient osteosarcoma cells to UV-

85

inactivated AAV leads to centriole overduplication and mitotic arrest (19). Another example of

86

mitotic catastrophe occurs in response to apoptin, a viral protein from chicken anemia virus.

87

Apoptin directly inhibits the APC/C, thereby preventing exit from mitosis, resulting in mitotic

88

arrest and subsequent apoptosis (36). This arrest was found to be independent of p53 and resulted

89

from inhibition of the metaphase-to-anaphase transition (33).

90

We show here that the E1B-55K and E4orf3 genes are sufficient to prevent mitotic arrest

91

in the adenovirus-infected cell. The E1B-55K protein circumvents entry into mitosis in a p53-

92

dependent manner. The E4orf3 protein circumvents mitotic arrest by facilitating exit from

93

mitosis, perhaps by mislocalizing cyclin B1. These early adenoviral proteins may prevent an

94

untimely death that would occur in response to inappropriate mitotic arrest. MATERIALS AND METHODS

95

Chemicals. Hydroxyurea (HU) from Sigma/Aldrich (St. Louis, MO) was used at a

6 96

concentration of 2 mM from a stock solution of 1 M in water. KaryoMAX Colcemid from

97

Gibco/Invitrogen (Gaithersburg, MD) was used at a concentration of 0.2 μg per mL from a stock

98

solution of 10 μg per mL in HBSS.

99

Cells. All cell culture media, supplements, and sera were obtained from Invitrogen

100

(Gaithersburg, MD) or Lonza (Hopkinton, MA) through the Tissue Culture and Virus Vector

101

Core Laboratory of the Comprehensive Cancer Center of Wake Forest University. HeLa cells

102

were maintained in Dulbecco's modified Eagle medium (DMEM) supplemented with 10%

103

newborn calf serum. H1299 cells were maintained in DMEM supplemented with 10% fetal

104

bovine serum. Both cell lines were cultured in a 5% CO2 atmosphere at 37oC by passaging twice

105

weekly at a 1:10 dilution. For all experiments, cells were plated at a density of 5×104 cells per

106

mL. For high-resolution immunofluorescence microscopy, cells were grown on nitric acid-

107

cleaned, sterilized glass coverslips in 6-well plates or 35-mm dishes. For cell cycle profile

108

analysis or protein lysate collection, cells were grown in 60-mm dishes.

109

Viruses. The phenotypically wild-type virus in this study contains several deletions and

110

a substitution within the E3B region. This virus, dl309, displays wild-type characteristics in

111

cultured cells (37). The virus referred to as the double-mutant virus is 3112, a previously

112

described (38) viral recombinant of the dl1520 virus that has a 827 bp deletion in the E1B-55K

113

open reading frame (39) and the dl341 virus that contains an E4orf3 deletion (40). Other E1B-

114

55K deletion mutants include dl338, which bears a 524 bp deletion (41) and dl110, which

115

contains a 472 bp deletion (42). The parental virus for the E1B-55K point mutation mutants

116

listed in Table 2 is H5pg4100, which contains a deletion in the E3 region at nucleotides 28593–

117

30471 and has an added endonuclease restriction site at nucleotide 30955 (BstBI). The

118

adenovirus mutants described in Table 2 were constructed by the method described in (43).

7 119

Briefly, point mutations were introduced into the E1B-55K gene in a shuttle plasmid by site-

120

directed mutagenesis. The appropriate fragment was inserted into the parental plasmid

121

pH5pg4100 and the viral genomes were released from the recombinant plasmids by digestion

122

with PacI. Mutant viruses recovered after transfection of the linear DNA into 293 cells. All

123

viruses were grown in 293 cells and concentrated virus stocks prepared by sequential

124

centrifugation through CsCl gradients as described previously (1).

125

Antibodies. Primary antibodies included a monoclonal mouse antibody against ß-tubulin

126

(Clone TUB 2.1, T4026) from Sigma Aldrich (St. Louis, MO) used at a 1:500 dilution, a

127

monoclonal pantropic mouse antibody against p53 (DO-1) obtained from CalBiochem

128

(Darmstadt, Germany) used at a 1:100 dilution, and a polyclonal rabbit antibody against

129

phospho-histone H3 (Thr11, #9849) from Cell Signaling (Danvers, MA) used at a 1:100 dilution

130

for immunofluorescence. Adenovirus-specific antibodies included the mouse monoclonal

131

antibody Rsa#3 (44) for the E4orf6/7 protein and the rat monoclonal antibody 6A11 (45) used as

132

hybridoma culture supernatant fluid diluted 1:5. Five antibody preparations for cyclin B1 were

133

used for this study. The mouse monoclonal antibody against cyclin B1 (#CC03) was obtained

134

from Oncogene/CalBiochem (Cambridge, MA) and used at a 1:250 dilution for microscopy and

135

1:1000 dilution for western blotting. The remaining cyclin B1 antibodies were obtained from

136

EMD Millipore (Billerica, MA) and included the mouse monoclonal antibody specific for the C-

137

terminus of human cyclin B1 (clone Y106) used at a 1:100 dilution for immunofluorescence and

138

a 1:10,000 dilution in 5% milk for western blotting, a mixture of mouse monoclonal IgGs

139

specific for human cyclin B1 (catalog # 05-373) used at 10 μg per mL for immunofluorescence

140

and 0.33 μg per mL for western blotting, a monoclonal antibody raised against hamster cyclin B1

141

(catalog # MAB3684) used at a 1:100 dilution for immunofluorescence and a 1:5000 dilution in

8 142

5% milk for western blotting and a rabbit monoclonal antibody specific for phosphorylated

143

serine 126 in cyclin B1 (catalog # MABE490) used at a 1:100 dilution for immunofluorescence

144

and a 1:3000 dilution for western blotting. Secondary antibodies used for immunofluorescence

145

microscopy were anti-mouse or anti-rabbit whole IgG conjugated to Alexa Fluor 488 (AF488) or

146

Alexa Fluor 568 (AF568) from Invitrogen used at 2 μg per mL. The secondary antibody used for

147

western blot analysis was an anti-mouse or anti-rabbit whole IgG raised in goats conjugated to

148

horseradish peroxidase from Jackson ImmunoResearch Laboratories (West Grove, PA) and used

149

at a concentration of 0.1 μg per mL.

150

Plasmids and PEI transfection. A plasmid expressing the wild-type human TP53 gene

151

under control of the CMV immediate early promoter and enhancer was generously provided by

152

Guangchao Sui (Wake Forest School of Medicine). Plasmids expressing human cyclin B1 fused

153

to the yellow fluorescent protein Venus (pVenus-N1 Cyclin B1, Addgene plasmid #26062) and

154

the degradation-resistant form of Cylin B1 (pVenus Cyclin B1 R42A, Addgene plasmid #39873)

155

were the gift of Jonathon Pines (Gurdon Institute, Cambridge, England). Cells were seeded on

156

glass coverslips in a 6-well culture dish. For each well, 1 μg of DNA was used in a volume of 0.2

157

mL serum-free media with poly(ethylenimine) (PEI) at a 1:5 dilution from a 7.5 mM stock

158

dissolved in deionized water with gentle heating. Transfections were carried out in a 5% CO2

159

atmosphere at 37°C with gentle rocking and regular rotation for 8 h before being replacing the

160

DNA and PEI mixture with growth medium.

161

Indirect immunofluorescence. Cells were washed twice with phosphate-buffered saline

162

(PBS), fixed for 30 min with 2% paraformaldehyde, and permeabilized for 5 min with 0.2%

163

Triton X-100 in PBS at room temperature. All subsequent washes were performed with Tris-

164

buffered saline with BSA, glycine and Tween-20 (TBS-BGT: 0.137 M NaCl, 0.003 M KCl,

9 165

0.025 M Tris-Cl [pH 8.0], 0.0015 M MgCl2, 0.5% bovine serum albumin, 0.1% glycine, 0.05%

166

Tween 20, and 0.02% sodium azide). Antibodies used for immunofluorescence were diluted in

167

TBS-BGT supplemented with 10% normal goat serum (Invitrogen). Samples were stained for 90

168

min with primary antibody and for 30 min with secondary antibody with multiple washes with

169

TBS-BGT between. Samples were mounted with ProLong Gold mounting media (Invitrogen)

170

containing 4′,6-diamidino-2-phenylindole (DAPI).

171

Microscopy. Micrographs were obtained by standard epifluorescence microscopy with a

172

Nikon TE300 inverted microscope or by confocal laser scanning microscopy with a Nikon TiE

173

inverted microscope fitted with a Nikon C1si system. A 100×/1.4 NA magnification oil-

174

immersion objective was used for all micrographs. Monochromatic images were acquired on the

175

Nikon TE300 microscope with a Retiga EX 1350 digital camera (QImaging Corp., Burnaby,

176

British Columbia, Canada). Confocal images were acquired as a series of 5 sections at 0.2 um

177

intervals in the center of the nucleus using sequential excitation for each of the three

178

fluorochromes. The confocal images are presented as a maximum intensity projection to reduce

179

the three-dimensional information to two-dimensions. Merged color images were prepared by

180

assigning either red, green or blue to the appropriate fluorochrome. For single channel

181

fluorescent images, a pseudocolor scheme was assigned in order to mimic the visual perception

182

with increased saturation at increased fluorescent intensity. Monochromatic images were

183

assigned colors with the open-source program ImageJ (version 1.46). All figures were assembled

184

with the open-source vector graphics editor Inkscape (version 0.48) and raster graphics editor

185

GIMP (version 2.8.4.)

186 187

Flow cytometry. Cells were harvested by trypsinization. EdU-labeled cells were labeled as indicated below and resuspended in FACS buffer (1% BSA in PBS) with propidium iodide

10 188

provided by the Click-iT EdU kit. All other samples were washed twice with PBS and

189

resuspended in 100 μL PBS. Cells were transferred dropwise into 2.5 mL of 70% ethanol with

190

constant mixing. Samples were stored at -20oC for at least 12 h. Ethanol was then removed and

191

the samples resuspended at a density of 106 cells per mL in PI solution diluted in water from a

192

10X stock (1 M NaCl, 0.36 M sodium citrate, 500 μg per mL propidium iodide, 6% NP-40) with

193

0.1 mg per mL RNase A diluted from a stock solution of 10 mg per mL in 0.01 M sodium acetate

194

with 0.1 M Tris-Cl, pH 7.4. Samples were incubated with PI solution for 30 min in the dark at

195

37oC and then transferred to ice before flow cytometric analysis. A Becton Dickinson FACS

196

Calibur instrument was used to acquire the propidium iodide signal in linear mode. Gating on the

197

peak width and peak area was used to select single cells for DNA profile and cell cycle analysis.

198

Western blotting. Cells were washed in PBS supplemented with protease and

199

phosphatase inhibitors (2 mM EDTA, 1 mM NaF, 1 mM sodium pyrophosphate, 1 mM

200

phenylmethylsulfonyl fluoride, 1 mM Na3VO4, and 2 μM leupeptin), harvested by scraping, and

201

suspended in a volume of one-tenth concentrated PBS sufficient to dilute the cells to a density of

202

2×107 cells per mL An equal volume of 2X sodium dodecyl sulfate (SDS) protein sample buffer

203

(4% SDS, 0.135 M Tris [pH 6.8], 20% glycerol, 0.02% bromophenol blue, and 6% β-

204

mercaptoethanol) was added to make the final concentration 107 cells per mL. The cell lysate

205

was heated for 5 min at 95oC and subjected to 3 pulses of sonication lasting 20 sec each. The

206

lysates were separated by SDS-polyacrylamide gel electrophoresis through 10% acrylamide gels.

207

The proteins were then electrophoretically transferred to nitrocellulose (Whatman/GE

208

Healthcare) overnight at 4oC. The nitrocellulose was blocked in TBS-BGT containing 5% nonfat

209

dry milk and sodium azide, stained with primary antibody diluted in TBS-BGT with sodium

210

azide overnight at 4oC, and stained with secondary antibody diluted in TBS-BGT without sodium

11 211

azide for 1 hr at room temperature. The stained protein was visualized by a mixture of

212

SuperSignal West Pico and SuperSignal West Femto chemiluminescence substrate from

213

Pierce/ThermoScientific (Rockford, IL) and X-ray film. The density of the specific signal

214

recorded by the X-ray film was quantified by scanning non-saturated exposures at 16-bit

215

resolution and measuring the optical density with the tools available in ImageJ.

216

Click-iT EdU. Cells were pulse-labeled with an appropriate amount of EdU from a 10

217

mM stock in DMSO diluted in a small volume of warm growth medium to make a final

218

concentration of 100 μM EdU. At the end of the labeling period, the wells were washed and

219

replaced with pre-warmed growth medium. At time of harvest, samples for DNA profile analysis

220

were harvested with trypsin while samples on glass coverslips were fixed, permeabilized, and

221

processed for immunofluorescence in a humidifying chamber. The samples were fixed with 4%

222

paraformaldehyde in PBS for 15 min, washed with FACS buffer (1% BSA in PBS), and

223

permeabilized with 0.5% Triton X-100 for 20 min at room temperature. The Click-iT reaction

224

cocktail was prepared by mixing Click-iT reaction buffer, CuSO4, fluorescently labeled azide,

225

and reaction buffer additive as indicated by the manufacturer. Samples were washed with FACS

226

buffer and allowed to incubate with the Click-iT reaction cocktail in a humidifying chamber at

227

room temperature for 30 min in the dark. The samples were washed and those on coverslips were

228

mounted on slides with Prolong Gold supplemented with DAPI. Samples for DNA profile

229

analysis were pelleted and resuspended in 500 μL of FACS buffer with 0.2 mg per mL RNase A

230

(from a 20 mg per mL) and 4 μg per mL PI (from a 1 mg per mL stock solution in water).

231

Cell synchronization. Mitotic cells were mechanically harvested from subconfluent cells

232

grown in 75 cm2 flasks and pelleted gently. The mitosis-enriched pellet was resuspended at a

233

concentration of 2x104 cells per mL in 2 mM HU in growth medium and plated in 60 mm dishes

12 234

for DNA profile analysis or on acid-treated glass coverslips in 35 mm dishes for

235

immunofluorescence. After 1 hour, plates were swirled to dislodge any dead S-phase cells and

236

the media replaced with a fresh solution of 2 mM HU in growth medium. Cells were held in HU

237

for 16 h to allow cells to cycle to the G1/S border. All samples were released from HU by

238

washing and replacing with growth medium. At various times post infection, cells were collected

239

for DNA profile analysis in order to determine cell cycle distribution in the sample. RESULTS

240

Cells infected during early G1 give rise to cells trapped in a mitosis-like state. A

241

subset of cells infected with an adenovirus deleted of the E1B-55K and E4orf3 genes develop

242

highly condensed chromatin resembling that of a mitotic cell (46). Because typically fewer than

243

10% of the infected cells develop the highly condensed chromatin, it seems likely that an

244

underlying process or condition limits the number of infected cells that condensed their

245

chromatin and resemble a mitotic cell. Potential processes include the death and loss of cells

246

following mitotic catastrophe, a small probability of entering mitosis, or the existence of a

247

limited subset of cells that are competent to enter mitosis. Since the outcome of an infection with

248

the E1B-55K-mutant virus is strongly determined by the stage of the cell cycle at the time of

249

infection (38, 47, 48), we performed two experiments to determine if entry into a mitotic-like

250

state was affected by the stage of the cell cycle at infection. Asynchronously dividing HeLa cells

251

were exposed to the thymidine analog 5-ethynyl-2´-deoxyuridine (EdU) for a 4-h interval in

252

order to identify cells that passed through S phase during this period of time. After returning the

253

cells to EdU-free medium, the cells were maintained for various intervals of time before being

254

infected. At the time of infection, a portion of the cells was collected and EdU-positive cells

255

were identified by the copper(I)-catalyzed reaction between a fluorescent azide and the alkyne in

13 256

EdU. This sample was analyzed by flow cytometry for DNA and EdU in order to determine the

257

stage of the cell cycle that EdU-positive cells had reached at infection. The labeling scheme and

258

associated flow cytometric analyses are shown in Fig. 1. After 72 h, another fraction of the

259

infected cells was stained for EdU and DNA and the number of mitotic-like nuclei were

260

enumerated by fluorescence microscopy. The results of scoring approximately 600 nuclei at each

261

time point show that cells with condensed DNA were seen most frequently among cells infected

262

during G2/M and early G1 (Table 1).

263

Although the EdU-labeling method permits analysis of unperturbed dividing cells, the

264

temporal resolution afforded by a 4 h labeling period was limited. For better resolution,

265

synchronously dividing HeLa cells were prepared as previously described (47, 49) and infected

266

with the E1B-55K/E4orf3 double-mutant virus at hourly intervals from 6 to 14 h following entry

267

into S-phase. A portion of the synchronized cells was collected and analyzed for DNA content by

268

flow cytometry in order to determine the stage of the cell cycle at the time of infection (Fig. 2A).

269

After 72 h of infection, the cells were fixed and the frequency of cells with mitotic-like nuclei

270

was determined (Fig. 2B). These results agree with the findings obtained with asynchronously

271

dividing cells (see Fig. 1) and indicate that cells infected with the E1B-55K/E4orf3 double-

272

mutant virus during a 4 h window in early G1 gave rise to the subset of cells with highly

273

condensed chromatin. These results suggest that the adenoviral E1B-55K and E4orf3 proteins

274

prevent cells infected early in G1 from becoming trapped in a mitotic-like state at late times of

275

infection.

276

Cells infected with the E1B-55K/E4orf3 double-mutant virus show evidence of

277

mitotic distress. Approximately four percent of asynchronously dividing HeLa cells contain

278

condensed chromatin typical of the later stages of mitosis (Fig. 3A, mock and Fig. 3B). The

14 279

mitotic spindle in these cells, visualized by staining for β-tubulin, occurs in a symmetrical

280

bipolar arrangement about the condensed chromatin in metaphase and anaphase cells (Fig. 3A

281

panel a). A fraction of the mitotic cells also contain highly phosphorylated histone H3 or

282

phospho-H3 (Fig. 3A panel b, and Fig. 3B). Phosphorylation of histone H3 initiates late in G2

283

phase, is completed in late prophase and is maintained through metaphase. Histone H3

284

phosphorylation precipitously decreases during anaphase as the cell exits mitosis (50). In

285

contrast to mock-infected cells, none of the cells infected with the wild-type or E4orf3-mutant

286

adenovirus and less than 0.5% of cells infected with the E1B-55K-mutant virus contained

287

condensed chromatin or phospho-H3 when evaluated at 72 hpi (Fig. 3B). This is consistent with

288

previous observations indicating that adenovirus-infected human epithelial cells cease

289

progression through the cell cycle (51). However, in these asynchronously infected cultures, a

290

significant number (>10%) of cells infected with the E1B-55K and E4orf3 double-mutant virus

291

contained chromatin characteristic of a mitotic cell (Fig. 3A, ΔE1B-55K/ΔE4orf3 and Fig. 3B).

292

Some of the double-mutant virus-infected cells also stained for phospho-H3, suggesting that

293

these cells progressed into mitosis. Although these cells superficially resembled mitotic cells,

294

many contained an asymmetric distribution of β-tubulin or even a multipolar spindle. The

295

abundance of aberrant mitotic spindles in these cells makes it seems likely that these cells will

296

fail to produce viable daughter cells. Expression of the small avian virus-derived protein apoptin

297

in human tumor cells has been reported to lead to a similar mixture of cells with apparently

298

normal, asymmetric, and multipolar spindles due to a block in the metaphase to anaphase

299

transition (33).

300 301

Cells infected with the double-mutant virus that appear to arrest in mitosis do so only at late times of infection (Fig. 3C). In some experiments, a transient increase in mitotic-like cells

15 302

occurred at 24 hpi. However, cells with highly condensed DNA do not begin to accumulate until

303

60 hpi. The number of cells apparently arrested in mitosis reached a maximum at 72 hpi. The

304

decrease in cells with mitotic-like nuclei after 72 hpi appears to be due to the death or selective

305

detachment of these cells from the substrate. Consequently, subsequent experiments evaluated

306

cells at 72 hpi. Taken together, these results suggest that the E1B-55K and E4orf3 proteins

307

prevent cells infected during early G1 from becoming trapped in a mitotic-like state at late times

308

of infection.

309

Adenoviral E1B-55K protein prevents entry into mitosis. Because significant

310

numbers of cells were observed to contain condensed DNA only after infection with the double-

311

mutant virus, it seems likely that the E1B-55K and E4orf3 proteins act independently to prevent

312

the infected cell from arresting in a mitotic-like state. Conceptually, these viral oncoproteins

313

could prevent the accumulation of mitotic-like cells by two non-exclusive mechanisms. First, the

314

viral proteins could prevent entry into mitosis. Second, the viral proteins could facilitate exit

315

from mitosis. Cherubini and associates observed that a fraction of E1B-55K-mutant virus-

316

infected cells accumulated highly condensed chromosomes after 12 h of exposure to colcemid

317

(52). By depolymerizing microtubules, colcemid prevents the completion of mitosis thus

318

trapping any cells that entered mitosis during the exposure to colcemid. Since mitotic-like cells

319

did not accumulate in colcemid-treated cells that were infected with the wild-type virus, we

320

reasoned that the E1B-55K protein might prevent the infected cell from entering mitosis. To test

321

this hypothesis, infected cells were treated with colcemid for 12 h prior to fixation and analysis

322

at various times after infection. The number of mitotic-like cells infected with the wild-type or

323

E4orf3-mutant virus did not increase, suggesting that cells infected with the wild-type or E4orf3-

324

mutant virus did not enter into mitosis (Fig. 4). As expected, cells with condensed DNA were

16 325

observed following infection with the double-mutant virus; the rate at which these mitotic-like

326

cells accumulated was the same in the presence (Fig. 4) or absence of colcemid (see Fig. 3C). In

327

contrast to the wild-type or E4orf3-mutant virus, colcemid trapped increasing numbers of

328

mitotic-like cells during the course of infection with the E1B-55K-mutant virus (Fig. 4). These

329

results, which recapitulate the findings reported by Cherubini (52), suggest that some E1B-55K-

330

mutant virus-infected cells enter and exit mitosis at late times of infection. This indicates that

331

another role for the E1B-55K protein is to prevent adenovirus-infected cells from entering into

332

mitosis.

333

E1B-55K prevents entry into mitosis through p53. The E1B-55K protein serves many

334

roles during an infection. The E1B-55K and E4orf6 proteins collaborate to form a novel SCF-

335

type E3 ubiquitin ligase with the scaffold protein Cul5, Elongins B and C, and the RING box

336

protein Rbx1 (12-14). This adenovirus-specific complex targets several cellular proteins for

337

degradation, many of which signal or respond to DNA-damage (see for example 53, 54-58).

338

Acting independently of E4orf6, the E1B-55K protein directs degradation of the cellular

339

transcription factor Daax, mislocalizes the DNA-damage responsive chromatin structure

340

regulator SPOC1, and can directly block p53-mediated transcription (11, 55, 59). In order to

341

identify the activity of the E1B-55K protein that prevents entry into mitosis, HeLa cells were

342

infected with the E1B-55K-mutant adenoviruses described in Table 2. At 60 hpi, the infected

343

cells were treated with colcemid to trap any cells entering mitosis and the cells with condensed

344

DNA were counted by fluorescence microscopy. As expected, cells infected with the three E1B-

345

55K-null adenoviruses as well as the virus bearing four nonsense mutations in the E1B-55K open

346

reading frames (H5pm4149) were trapped in a mitotic-like state by colcemid (Figs. 5A and 5B).

347

Among the viruses bearing missense mutations in the E1B-55K gene, only H5pm4109 was

17 348

associated with a significant increase in mitotic-like cells (Fig. 5B). This virus encodes an E1B-

349

55K protein that is unable to bind p53. All other E1B-55K variants analyzed in Fig. 5B appeared

350

to prevent entry into mitosis as well as the wild-type virus H5pg4100. Because over 20% of non-

351

infected cells were trapped in a mitotic-like state by this treatment (data not shown), the failure

352

to detect mitotic-like cells among virus-infected cells confirms that the cells were uniformly

353

infected. Because cells infected with the mutant virus H5pm4108 were not trapped in a mitotic-

354

like state by colcemid, we conclude that the ability to block entry into mitosis is independent of

355

the E1B-55K/E4orf6 protein complex since H5pm4108 expresses an E1B-55K protein that fails

356

to bind E4orf6 (see 60). These findings suggest that the key property of the E1B-55K protein

357

needed to prevent entry into mitosis is its ability to interact with p53.

358

In many HPV-transformed cells, expression of the integrated E6 gene of HPV directs the

359

continual degradation of p53 protein (61). Without the selective pressure to eliminate p53 gene

360

function, the TP53 gene in HeLa cells has remained intact (62). Wild-type p53 protein

361

accumulates to measurable levels in HeLa cells infected with adenovirus and to a high level in

362

cells infected with adenovirus mutants that fail to direct the degradation of p53. To examine the

363

possibility that p53 contributes to mitotic entry in the infected cell, p53-null H1299 cells were

364

infected with the E1B-55K/E4orf3 double-mutant virus and evaluated. In sharp contrast to HeLa

365

cells, no mitotic-like cells were observed among H1299 cells infected with this virus at any time

366

after infection (data not shown and see Fig. 5C).

367

To test directly a role for p53, H1299 cells were transfected with a p53-expression

368

plasmid and infected after 24 h. At 60 hpi, cells were exposed to colcemid or left untreated. At

369

72 hpi, cells were stained for p53 and DNA then evaluated by fluorescence microscopy.

370

Enforced expression of p53 is acutely toxic to H1299 cells (63). Consequently, transfection of

18 371

the p53-expression plasmid reduced the number of evaluable cells and induced apoptosis in the

372

mock-infected cells after three days (data not shown). It seems likely that the increased number

373

of p53-positive, mock-infected H1299 cells with condensed DNA was due to pyknosis or

374

chromatin condensation from apoptosis rather than mitosis (Fig. 5C, mock). Apoptosis is

375

inhibited in the adenovirus-infected cells because all of the viruses studied here express the anti-

376

apoptotic E1B-19K protein (64). Accordingly, in the absence of colcemid, neither pyknotic nor

377

mitotic-like nuclei were observed among E1B-55K-mutant virus-infected H1299 cells (Fig. 5C).

378

However, colcemid trapped a significant number of p53-positive H1299 cells in a mitotic-like

379

state. Similarly, mitotic-like nuclei were observed only in p53-positive H1299 cells infected with

380

the E1B-55K/E4orf3 double-mutant virus. As noted for HeLa cells, colcemid was not required to

381

trap the mitotic-like double-mutant virus-infected H1299 cells. At least for the HeLa and H1299

382

cell lines, these results are consistent with an unexpected role for p53; if the infected cells with

383

condensed nuclei are indeed mitotic, p53 appears to be necessary for the adenovirus-infected cell

384

to enter mitosis. Furthermore, the E1B-55K protein blocks this activity of p53.

385

E4orf3 targets cyclin B1. Colcemid is required to trap HeLa cells and p53-positive

386

H1299 cells infected with the E1B-55K-mutant virus in a mitotic-like state. By contrast, p53-

387

positive cells infected with the E1B-55K/E4orf3 double-mutant virus are trapped in a mitotic-

388

state without colcemid. If the absence of the E1B-55K protein allows infected p53-positive cells

389

to enter mitosis, this result suggests that the presence of E4orf3 protein may facilitate exit from

390

mitosis. Degradation of the major mitotic cyclin, cyclin B1 is a critical event that precipitates

391

exit from mitosis (65). The complete degradation of cyclin B1 is required for cells to proceed

392

with cytokinesis (66). Previous studies showed an increase in cyclin B1 in wild-type adenovirus-

393

infected WI-38 and A549 cells as well as an S-phase-dependent increase in cyclin B1 in E1B-

19 394

55K-mutant virus-infected cells (67). We therefore compared the nature and abundance of cyclin

395

B1 among HeLa cells infected with wild-type and mutant viruses by immunoblotting. The level

396

of cyclin B1 was indeed higher in virus-infected cells compared to mock-infected cells (Fig. 6)

397

although it seems unlikely that the modest increase in cyclin B1 level in double-mutant virus-

398

infected cells could force an apparent mitotic arrest. Adenovirus-infected cells also contained an

399

additional cyclin B1-related protein of slightly greater electrophoretic mobility than the form in

400

mock-infected cells (Fig. 6A). This product was recognized by four different cyclin B1-specific

401

antibody preparations. Both ~50-kDa forms of cyclin B1 appeared equally abundant when

402

queried by a phosphoserine-126-specific antibody (data not shown). The origin of the two forms

403

of protein of approximately 50-kDa remains unclear. A 35-kDa protein recognized by cyclin B1

404

antibodies was detected in lysates from cells infected with the E1B-55K/E4orf3 double-mutant

405

virus and to a lesser extent, with the E4orf3-mutant virus. This smaller form of cyclin B1 appears

406

to correspond to a cleavage product found during mitotic catastrophe termed cyclin B1Δ (34). It

407

was suggested that cyclin B1Δ acts as a dominant-negative inhibitor of cyclin B1 function and

408

sustains the mitotic block in cells that would otherwise exit mitosis (34). Both the elevated levels

409

of cyclin B1 and the presence of a dominant-negative inhibitor could retard E1B-55K/E4orf3

410

double-mutant virus-infected cells in mitosis. In addition, changes in the localization of cyclin

411

B1 in the infected cells point to another mechanism by which E4orf3 could facilitate exit from

412

mitosis.

413

Because E4orf3 disrupts cell signaling pathways by mislocalizing host proteins (68, 69)

414

we explored the possibility that E4orf3 mislocalizes cyclin B1 in the infected cell. The different

415

localizations of cyclin B1 among mock-infected cells were consistent with fluctuations in the

416

level and movement of cyclin B1 during cell cycle progression. Cells in G2 contained high levels

20 417

of cyclin B1 that was found in a diffuse or speckled pattern in the cytoplasm (Fig. 7A, panel a).

418

Early in mitosis, cells contained high levels of cyclin B1 that was largely coincident with

419

chromatin (Fig. 7A, panel b). Finally, a subset of cells with condensed chromatin was judge to be

420

in late mitosis because of the absence of cyclin B1 staining (Fig. 7A, panel c). A subset of cells

421

infected with the wild-type and E1B-55K-mutant virus contained nuclear cyclin B1; in many of

422

these cells, cyclin B1 was found in large aggregates throughout the nucleus (Fig. 7B panels a-d).

423

A fraction of cells infected with E4orf3-mutant viruses also contained nuclear cyclin B1. In

424

contrast to infected cells containing the E4orf3 protein, cyclin B1 was diffusely distributed in the

425

nucleus of cells infected with the E4orf3-mutant viruses (Fig. 7B, panels e-h). The distribution of

426

cyclin B1 in these infected cells more closely resembled the patterns observed in mitotic or G2

427

mock-infected cells. These results, which are quantified in Fig. 7C, show that the E4orf3 protein

428

alters the distribution of cyclin B1 in the cell nucleus during an adenoviral infection. E4orf3 may

429

functionally inactivate cyclin B1 in order to facilitate exit from mitosis.

430

E4orf3 overcomes metaphase arrest imposed by a non-degradable cyclin B1.

431

Degradation of cyclin B1 during mitosis requires key residues (R42xxL45xxI/V48xN50) in the

432

destruction box (70). Expression of cyclin B1 variants with mutations in the destruction box

433

force cells to accumulate in a mitotic-like state (66, 71). We expressed E4orf3 and the wild-type

434

or degradation-resistant cyclin B1 (CycB1 R42A) by transfection to determine if E4orf3 can

435

overcome the metaphase arrest imposed by CycB1 R42A (71). As expected, enforced expression

436

of wild-type cyclin B1 increased the fraction of cells with condensed DNA (Fig. 8A). Although

437

expression of the E4orf6/7 cDNA reduced the frequency of these mitotic-like cells, the

438

difference was not statistically significant (p=0.09). E4orf3 decreased the frequency of these

439

mitotic-like cells to statistically significant, although modestly reduced levels (7%, p=0.02).

21 440

However, the effect of E4orf3 was pronounced in cells expressing the degradation-resistant

441

R42A cyclin B1. More than half of the cells expressing the R42A variant contained condensed

442

DNA after two days. This number was not affected by E4orf6/7 (p=0.32). By contrast, E4orf3

443

significantly (p=0.01) and substantially reduced the fraction of mitotic-like cells expressing the

444

R42A variant (Fig. 8A). These results indicate that expression of E4orf3 overcomes the mitotic-

445

like state caused by elevated levels of cyclin B1 without forcing the degradation of cyclin B1.

446

Expression of E4orf3 by transfection did not promote the aggregation of cyclin B1 seen

447

in virus-infected cells (see Fig. 7C). We observed no differences in the localization of the cyclin

448

B1 fusion protein among cells expressing E4orf6/7 (Fig. 8B), E4orf3 (Fig. 8C) or no E4

449

construct (data not shown). Occasional aggregates of cyclin B1 were noted irrespective of the co-

450

transfected viral gene such as in Fig. 8B, panel a. However, cyclin B1 appeared to be largely

451

excluded from the nucleus in most of the cells expressing both E4orf3 and the cyclin B1 fusion

452

protein. Since so few productively transfected cells contained detectable levels of endogenous

453

cyclin B1, it was not possible to determine if E4orf3 affected the endogenous protein in a similar

454

manner. Because cells expressing E4orf3 and cyclin B1 R42A did not have significant levels of

455

the cyclin B1 in the nucleus after 24 h, 36 h, and 72 h of transfection (data not shown), it seems

456

likely that E4orf3 precludes entry of cyclin B1 into the nucleus or promotes nuclear export of

457

cyclin B1. DISCUSSION

458

Cells infected by the E1B-55K/E4orf3 double-mutant virus during early G1 are

459

predisposed to arrest in a mitotic-like state. The dependence of this phenomenon on the stage of

460

the cell cycle at the time of infection is unusual but not unexpected. We previously demonstrated

461

that cells infected during S-phase by the E1B-55K-deleted virus support a more productive

22 462

infection (47, 48) and are more rapidly killed than G1-infected cells (51, 72). The findings

463

reported here reinforce the notion that an infection initiated during G1 is restrictive for E1B-

464

55K-mutant adenoviruses.

465

The importance of the E1B-55K and E4orf3 proteins in preventing entry and arrest in

466

mitosis is counterintuitive given that these proteins disable DNA-damage checkpoints. In

467

particular, a dysfunctional G2 checkpoint can allow a cell to enter mitosis inappropriately (73).

468

The G2 checkpoint prevents mitotic entry by blocking activation of the mitotic kinase Cdk1,

469

which depends on activation of the Cdc25C phosphatase and inactivation of the Wee1 and Myt1

470

kinases. Cdc25C and Wee1 are regulated by the ATM (6) and ATR kinases (74). Agents that

471

disable the ATM and ATR pathways thus can force G2 cells to enter mitosis prematurely.

472

Because both E1B-55K and E4orf3 proteins inactivate ATM (75) and ATR (76), the ability of

473

these two adenoviral proteins to prevent cells from entering mitosis or arresting in mitosis is

474

surprising.

475

Insight into adenoviral control of mitotic entry was provided by Cherubini and associates

476

who showed that primary human cells infected with the E1B-55K-mutant virus dl1520

477

accumulated highly condensed chromosomes after 12 h of exposure to colcemid (52). Here we

478

show that additional adenoviruses bearing large deletions in the E1B-55K gene behaved

479

similarly (Fig. 5A). However, among ten E1B-55K-mutant viruses bearing missense mutations

480

(Table 2), all but H5pm4109 prevented cells from entering mitosis (Fig. 5B). We interpret these

481

observations to mean that many E1B-55K properties, including the ability to bind the E4orf6,

482

Daxx and Mre11 proteins, the ability to interact with SUMO-modified proteins, and C-terminal

483

phosphorylation are not critical for the E1B-55K protein to block to entry into mitosis. The

484

H5pm4109 E1B-55K protein contains a histidine in place of alanine at position 260. The

23 485

identical protein expressed by the related virus ONYX-053 is unable to bind both p53 and

486

E4orf6 (60). Because the T255A protein expressed by H5pm4108 fails to bind E4orf6 (60) but

487

prevents entry into mitosis, we conclude that the E1B-55K protein must be able to bind p53 in

488

order to prevent entry into mitosis. Surprisingly, neither the E1B-55K/Eorf3 double-mutant virus

489

nor the combination of colcemid and infection with the E1B-55K single-mutant virus forced the

490

p53-null H1299 cells into mitosis unless p53 was expressed by transfection (Fig. 5C). In addition

491

to the HeLa cells studied in this report, H460 cells, MCF10A cells and hTERT-immortalized

492

retinal pigmented epithelial cells also contain a wild-type p53 gene. A significant number of

493

these cells arrested in mitosis after infection with the double-mutant virus whereas the p53-null

494

H358 and PC3 cell lines failed to show a statistically significant increase in arrested cells after

495

infection (data not shown).

496

The requirement for p53 to promote mitotic entry in the adenovirus-infected cell is at

497

odds with the well-described ability of p53 to enforce cell cycle arrest. For example, p53

498

precludes entry into S-phase by imposing transcriptional and translational blocks to cell cycle

499

progression in response to serum-starvation (77) and prevents inappropriate entry into mitosis by

500

suppressing expression of the chromosome alignment protein Kif23 (17). Perhaps p53 is altered

501

in the adenovirus-infected cell and this altered form of p53 permits cells to slip past the G2

502

checkpoint despite DNA-damage or chromosomal abnormalities. There is ample precedence for

503

the corruption of normal cellular functions by adenovirus proteins. For example, the E1B-55K

504

proteins of different adenoviruses partner with the E4orf6 protein to reprogram ubiquitin-protein

505

ligases of the Skp1/Cul1/F-box (SCF) family (53, 54). SCF complexes orchestrate progression

506

through the cell cycle and activate checkpoint signaling (recently reviewed in 78). It may be

507

significant that the SCF complex associated with the nuclear interaction partner of ALK keeps

24 508

levels of cyclin B1 low and prevents early mitotic entry (79). Furthermore, the physical

509

interaction between the E1B-55K protein and p53 converts p53 from a transcriptional activator

510

to a potent repressor of transcription (11, 80).

511

The apparent need for an interaction between the E1B-55K protein and p53 in order to

512

prevent mitotic arrest may underlie observations suggesting that p53 enables the wild-type virus

513

to elicit greater cytopathic effects than the E1B-55K-mutant virus (81). It was later noted that

514

replication of the E1B-55K-null virus dl1520 was enhanced by expression of the gain-of-

515

function R248W p53 variant that could no longer bind DNA in a site-specific manner (82). It

516

would be of interest to determine if the interaction between the E1B-55K protein and the putative

517

form of p53 responsible for preventing entry into mitosis subverts the transcriptional activity of

518

p53 in a manner that phenocopies gain-of-function p53 variants. Some of these variants promote

519

progression into mitosis in response to genotoxic stress (83). Many gain-of-function mutations in

520

p53 map to the DNA-binding domain (84). Coincidentally, Tip60-dependent acetylation of p53

521

within the DNA-binding domain can determine how p53 governs cell fate in response to DNA-

522

damage (85). Recent evidence shows that Tip60 is targeted for degradation by the E1B-55K

523

protein (54). Although the altered function of p53 is seen irrespective of E1B-55K status, other

524

viral proteins may mimic the action of Tip60, either directly or through the action of redirected

525

cellular proteins.

526

It must be emphasized that the interaction between the E1B-55K protein and p53 cannot

527

simply ablate p53 function, otherwise p53-null cells infected with the E1B-55K/E4orf3 double-

528

mutant virus should have arrested in the mitotic-like state. If a modified form of p53 exists

529

during infection, perhaps the E1B-55K protein allows this form of p53 to reinforce the G2/M

530

checkpoint during the replicative stages of the adenoviral lifecycle.

25 531

The notion of a protein inhibiting the function of another while simultaneously preserving

532

some of the target’s function is not novel. The licensing factor Cdt1 is present during G1 and S

533

phase. During DNA synthesis Cdt1 is expelled from the nucleus, degraded, and inactivated to

534

prevent re-replication. Geminin was initially identified as a mammalian factor that contributed to

535

the inactivation of Cdt1 (86). However, it was later shown that Geminin binds Cdt1 to gain

536

nuclear localization (87) and preserve a portion of Cdt1 during late mitosis (88). Perhaps like

537

Geminin, the E1B-55K protein is able to inactivate most p53 function while retaining a low level

538

of p53 in order to strengthen the G2/M checkpoint.

539

The E1B-55K protein, through its interaction with p53 during an adenovirus infection,

540

acts to prevent entry into mitosis. Because sustained expression of E1B-55K alone does not

541

preclude cell division, this novel viral “checkpoint” must form only in the adenovirus-infected

542

cell. Adenovirus-infected cells that escape this viral G2 checkpoint and enter mitosis may be

543

prone to arrest because of other adenoviral proteins such as the E4orf4 protein. The E4orf4

544

protein inactivates APC/C by reprogramming the activity of protein phosphatase 2A, thereby

545

inducing G2/M arrest (89, 90). Interestingly, when expressed alone, E4orf4 promotes p53-

546

independent, caspase-independent cell death in tumor cells (91-93) while inhibiting apoptosis in

547

normal cells (94). Since prolonged arrest in a mitotic-like state or mitotic catastrophe is often

548

followed by apoptosis or senescence, it would be advantageous for adenovirus to express an

549

activity that facilitates exit from mitosis. Because colcemid fails to trap E4orf3 single-mutant

550

virus-infected cells in a mitotic-like state (Fig. 4), we conclude that E4orf3 does not prevent

551

entry into mitosis. However, because the E4orf3 gene is required to prevent the E1B-55K-mutant

552

virus from arresting in a mitotic-like state (Fig. 3) we suggest that E4orf3 provides an activity

553

that facilitates exit from mitosis.

26 554

Progression through mitosis beyond anaphase requires satisfaction of the spindle

555

assembly checkpoint and activation of the APC/C. Two coactivators of APC/C, Cdc20 and

556

Cdh1, dictate substrate specificity and activity. Cdc20 is sequestered by mitotic checkpoint

557

proteins while unattached kinetochores or perturbations in the mitotic spindle persist. APC/CCdc20

558

initiates anaphase by targeting the cohesins for degradation while APC/CCdh1 promotes the

559

irreversible exit from mitosis into G1 (reviewed in 78). A key target of both APC/CCdc20 and

560

APC/CCdh1 is the mitotic cyclin B1, whose loss leads to the rapid decline in Cdk1 activity (22).

561

Because cyclin B1 accumulates to high levels in adenovirus-infected cells (Fig. 6), E4orf3 does

562

not promote mitotic exit by simply removing cyclin B1. However, the E4orf3 protein may

563

functionally inactivate cyclin B1 by mislocalizing this protein within the infected cell (Fig. 7).

564

Cells infected with E4orf3-mutant viruses also contain a cyclin B-related protein (Fig. 6) similar

565

in size to a cleaved form of cyclin B1 that sustains the mitotic block (34). In the context of an

566

E1B-55K-deletion, E4orf3 may prevent accumulation of this inhibitory product.

567

E4orf3 inactivates many cellular proteins by altering their localization or directing them

568

to the aggresome (95, 96). For example, mislocalization of Nbs1 by the E4orf3 protein is

569

sufficient to prevent ATR activation, thereby crippling the DNA-damage response (76, 97). The

570

activity of cyclin B1 in association with the major mitotic kinase Cdk1 is exquisitely controlled

571

by its intracellular localization, with roles both in the cytoplasm and the nucleus (22). In the

572

infected cell, E4orf3 may promote aggregation of cyclin B1 (Fig. 7). When expressed by

573

transfection, E4orf3 may diminish nuclear levels of cyclin B1 (Fig. 8). E4orf3 may functionally

574

inactivate cyclin B1 to facilitate exit from mitosis. In this manner, E4orf3 may replace a cellular

575

process that was disabled by other adenovirus proteins in order to prevent the untoward

576

consequences of mitotic arrest.

27 577

A remaining question is, why are cells infected in early G1 predisposed to arrest in

578

mitosis after infection with the double-mutant virus? Cellular chromatin in the early G1 cell is

579

distinguished from that in other phases of the cell cycle by the need to reacquire specific

580

proteins. At late stages of mitosis, many proteins, including those that signal DNA-damage are

581

displaced from the condensing chromatin. Consequently G1 cells are able to respond to DNA-

582

damage only after these proteins re-associate with DNA (98, 99). The reacquisition of chromatin

583

proteins can also be regulated by proteins such as the histone variant H2A.Z, which suppresses

584

transcription during mitosis (100). The histone acetyl transferase Tip60 promotes H2A.Z-

585

mediated association of proteins at sites of DNA-damage. Perhaps the ability of the E1B-55K

586

protein to target Tip60 for destruction (54) renders cellular chromatin less able to signal DNA-

587

damage and thus unsuited for further progression through the cell cycle. During late mitosis and

588

early G1 chromatin is also reloaded with replication-critical licensing factors that were displaced

589

or degraded in S phase (101, 102). Early events during the adenovirus infectious cycle may

590

perturb these processes that occur during G1.

591

Studies using somatic cell nuclear transfer to generate cloned animal offspring

592

demonstrate how the unique nature of cells in early G1 can exert consequences over a very long

593

time. Donor cell nuclei derived from serum-starved G0 cells or early G1 cells were transferred

594

into enucleated bovine ova to create embryos. Although embryos derived from G0 nuclei showed

595

greater survival through the blastocyst stage, animals derived from G1 nuclei showed improved

596

birth weights and greater post-natal survival (103). Both the transfer of an early G1 nucleus and

597

the infection of a cell in early G1 have consequences evident much later in time.

598 599

While the properties of early G1 cells that predisposed them to mitotic arrest following infection with adenovirus are unknown, these cells have the potential to enter mitosis despite

28 600

suspension of normal cell cycle progression. Surprisingly, entry into mitosis by the virus-infected

601

cell requires p53, a protein better known for its ability to halt cell cycle progression. The

602

E1B-55K and E4orf3 proteins function independently of one another to prevent arrest in mitosis.

603

Although adenovirus directs the degradation of p53, it appears that a portion of the p53 protein is

604

repurposed by the E1B-55K protein to reinforce the G2/M checkpoint. Through the

605

reorganization of cyclin B1, E4orf3 protein may serve as a failsafe to facilitate exit from mitosis,

606

presumably circumventing cell death associated with prolonged mitotic arrest. ACKNOWLEDGEMENTS

607

Cell culture reagents were provided by the Cell and Viral Vector Core Laboratory,

608

supported by the Comprehensive Cancer Center of Wake Forest University NCI CCSG

609

P30CA012197 grant. Flow cytometry was performed in the Flow Cytometry Core Laboratory,

610

also supported by the Comprehensive Cancer Center of Wake Forest University NCI CCSG

611

P30CA012197 grant. R.L.T. was supported by the Training Program in Immunology and

612

Pathogenesis 5 T32 AI007401 from the National Institutes of Health. This research was

613

supported by Public Health Service grant R01 CA127621 (to D.A.O) from the National Cancer

614

Institute and Deutsche Forschungsgemeinschaft grant Do 343/7-1 (to T.D.)

615

The authors wish to acknowledge the helpful discussion of members of the Parks, Lyles

616

and Barton laboratories of Wake Forest University. We thank Guangchao Sui for kindly

617

providing a p53-expression construct. The content of this report is solely the responsibility of the

618

authors and does not necessarily represent the official views of the respective funding agencies.

29 REFERENCES 619

1.

620 621

Shepard RN, Ornelles DA. 2004. Diverse roles for E4orf3 at late times of infection revealed in an E1B 55-kilodalton protein mutant background. J. Virol. 78:9924-9935.

2.

Nichols GJ, Schaack J, Ornelles DA. 2009. Widespread phosphorylation of histone

622

H2AX by species C adenovirus infection requires viral DNA replication. J. Virol. 83:5987-

623

5998.

624

3.

625 626

Gen. Virol. 93:2076-2097. 4.

627 628

5.

Hill R, Lee PW. 2010. The DNA-dependent protein kinase (DNA-PK): More than just a case of making ends meet? Cell Cycle 9:3460-3469.

6.

631 632

Bologna S, Ferrari S. 2013. It takes two to tango: Ubiquitin and SUMO in the DNA damage response. Front Genet 4:106.

629 630

Turnell AS, Grand RJ. 2012. DNA viruses and the cellular DNA-damage response. J.

Shiloh Y, Ziv Y. 2013. The ATM protein kinase: regulating the cellular response to genotoxic stress, and more. Nat. Rev. Mol. Cell Biol. 14:197-210.

7.

633

Nam EA, Cortez D. 2011. ATR signalling: more than meeting at the fork. Biochem. J. 436:527-536.

634

8.

Brady CA, Attardi LD. 2010. p53 at a glance. J. Cell Sci. 123:2527-2532.

635

9.

Yew PR, Berk AJ. 1992. Inhibition of p53 transactivation required for transformation by

636 637

adenovirus early 1B protein. Nature 357:82-85. 10.

Teodoro JG, Branton PE. 1997. Regulation of p53-dependent apoptosis, transcriptional

638

repression, and cell transformation by phosphorylation of the 55-kilodalton E1B protein of

639

human adenovirus type 5. J. Virol. 71:3620-3627.

640

11.

Martin ME, Berk AJ. 1998. Adenovirus E1B 55K represses p53 activation in vitro. J.

30 641 642

Virol. 72:3146-3154. 12.

Querido E, Blanchette P, Yan Q, Kamura T, Morrison M, Boivin D, Kaelin WG,

643

Conaway RC, Conaway JW, Branton PE. 2001. Degradation of p53 by adenovirus

644

E4orf6 and E1B55K proteins occurs via a novel mechanism involving a Cullin-containing

645

complex. Genes Dev. 15:3104-3117.

646

13.

Harada JN, Shevchenko A, Pallas DC, Berk AJ. 2002. Analysis of the adenovirus E1B-

647

55K-anchored proteome reveals its link to ubiquitination machinery. J. Virol. 76:9194-

648

9206.

649

14.

Blanchette P, Cheng CY, Yan Q, Ketner G, Ornelles DA, Dobner T, Conaway RC,

650

Conaway JW, Branton PE. 2004. Both BC-box motifs of adenovirus protein E4orf6 are

651

required to efficiently assemble an E3 ligase complex that degrades p53. Mol. Cell. Biol.

652

24:9619-9629.

653

15.

654 655

of p53 target genes by a small viral protein. Nature 466:1076-1081. 16.

656 657

Soria C, Estermann FE, Espantman KC, O'Shea CC. 2010. Heterochromatin silencing

Schwartz D, Rotter V. 1998. p53-dependent cell cycle control: response to genotoxic stress. Semin. Cancer Biol. 8:325-336.

17.

Fischer M, Grundke I, Sohr S, Quaas M, Hoffmann S, Knörck A, Gumhold C, Rother

658

K. 2013. p53 and cell cycle dependent transcription of kinesin family member 23 (KIF23)

659

is controlled via a CHR promoter element bound by DREAM and MMB complexes. PLoS

660

One 8:e63187.

661

18.

662 663

Sanhaji M, Louwen F, Zimmer B, Kreis NN, Roth S, Yuan J. 2013. Polo-like kinase 1 inhibitors, mitotic stress and the tumor suppressor p53. Cell Cycle 12:1340-1351.

19.

Ingemarsdotter C, Keller D, Beard P. 2010. The DNA damage response to non-

31 664

replicating adeno-associated virus: Centriole overduplication and mitotic catastrophe

665

independent of the spindle checkpoint. Virology 400:271-286.

666

20.

667 668

Mol. Cell. Biol. 30:22-32. 21.

669 670

22.

23.

24.

Gavet O, Pines J. 2010. Activation of cyclin B1-Cdk1 synchronizes events in the nucleus and the cytoplasm at mitosis. J. Cell Biol. 189:247-259.

25.

677 678

Bruinsma W, Raaijmakers JA, Medema RH. 2012. Switching Polo-like kinase-1 on and off in time and space. Trends Biochem. Sci. 37:534-542.

675 676

Takizawa CG, Morgan DO. 2000. Control of mitosis by changes in the subcellular location of cyclin-B1-Cdk1 and Cdc25C. Curr. Opin. Cell Biol. 12:658-665.

673 674

Pines J, Rieder CL. 2001. Re-staging mitosis: a contemporary view of mitotic progression. Nat. Cell Biol. 3:E3-6.

671 672

Chin CF, Yeong FM. 2010. Safeguarding entry into mitosis: the antephase checkpoint.

Santos SD, Wollman R, Meyer T, Ferrell JE, Jr. 2012. Spatial positive feedback at the onset of mitosis. Cell 149:1500-1513.

26.

Dephoure N, Zhou C, Villen J, Beausoleil SA, Bakalarski CE, Elledge SJ, Gygi SP.

679

2008. A quantitative atlas of mitotic phosphorylation. Proc. Natl. Acad. Sci. U. S. A.

680

105:10762-10767.

681

27.

682 683

Heald R, McKeon F. 1990. Mutations of phosphorylation sites in lamin A that prevent nuclear lamina disassembly in mitosis. Cell 61:579-589.

28.

684

Hirano T. 2012. Condensins: universal organizers of chromosomes with diverse functions. Genes Dev. 26:1659-1678.

685

29.

Li M, Zhang P. 2009. The function of APC/CCdh1 in cell cycle and beyond. Cell Div 4:2.

686

30.

Mehta GD, Rizvi SM, Ghosh SK. 2012. Cohesin: a guardian of genome integrity.

32 687 688

Biochim. Biophys. Acta 1823:1324-1342. 31.

689 690

Castedo M, Perfettini JL, Roumier T, Andreau K, Medema R, Kroemer G. 2004. Cell death by mitotic catastrophe: a molecular definition. Oncogene 23:2825-2837.

32.

Galluzzi L, Vitale I, Abrams JM, Alnemri ES, Baehrecke EH, Blagosklonny MV,

691

Dawson TM, Dawson VL, El-Deiry WS, Fulda S, Gottlieb E, Green DR, Hengartner

692

MO, Kepp O, Knight RA, Kumar S, Lipton SA, Lu X, Madeo F, Malorni W, Mehlen

693

P, Nuñez G, Peter ME, Piacentini M, Rubinsztein DC, Shi Y, Simon HU,

694

Vandenabeele P, White E, Yuan J, Zhivotovsky B, Melino G, Kroemer G. 2012.

695

Molecular definitions of cell death subroutines: recommendations of the Nomenclature

696

Committee on Cell Death 2012. Cell Death Differ. 19:107-120.

697

33.

Lanz HL, Zimmerman RM, Brouwer J, Noteborn MH, Backendorf C. 2013. Mitotic

698

catastrophe triggered in human cancer cells by the viral protein apoptin. Cell Death Dis.

699

4:e487.

700

34.

701 702

Chan YW, Chen Y, Poon RY. 2009. Generation of an indestructible cyclin B1 by caspase-6-dependent cleavage during mitotic catastrophe. Oncogene 28:170-183.

35.

Li S, Szymborski A, Miron MJ, Marcellus R, Binda O, Lavoie JN, Branton PE. 2009.

703

The adenovirus E4orf4 protein induces growth arrest and mitotic catastrophe in H1299

704

human lung carcinoma cells. Oncogene 28:390-400.

705

36.

Teodoro JG, Heilman DW, Parker AE, Green MR. 2004. The viral protein Apoptin

706

associates with the anaphase-promoting complex to induce G2/M arrest and apoptosis in

707

the absence of p53. Genes Dev. 18:1952-1957.

708 709

37.

Jones N, Shenk T. 1979. Isolation of adenovirus type 5 host range deletion mutants defective for transformation of rat embryo cells. Cell 17:683-689.

33 710

38.

711 712

Shepard RN, Ornelles DA. 2003. E4orf3 is necessary for enhanced S-phase replication of cell cycle-restricted subgroup C adenoviruses. J. Virol. 77:8593-8595.

39.

Barker DD, Berk AJ. 1987. Adenovirus proteins from both E1B reading frames are

713

required for transformation of rodent cells by viral infection and DNA transfection.

714

Virology 156:107-121.

715

40.

Sarnow P, Hearing P, Anderson CW, Reich N, Levine AJ. 1982. Identification and

716

characterization of an immunologically conserved adenovirus early region 11,000 Mr

717

protein and its association with the nuclear matrix. J. Mol. Biol. 162:565-583.

718

41.

Pilder S, Moore M, Logan J, Shenk T. 1986. The adenovirus E1B-55K transforming

719

polypeptide modulates transport or cytoplasmic stabilization of viral and host cell mRNAs.

720

Mol. Cell. Biol. 6:470-476.

721

42.

722 723

required for efficient shutoff of host protein synthesis. J. Virol. 50:202-212. 43.

724 725

Babiss LE, Ginsberg HS. 1984. Adenovirus type 5 early region 1b gene product is

Groitl P, Dobner T. 2007. Construction of adenovirus type 5 early region 1 and 4 virus mutants. Methods Mol. Med. 130:29-39.

44.

Marton MJ, Baim SB, Ornelles DA, Shenk T. 1990. The adenovirus E4 17-kilodalton

726

protein complexes with the cellular transcription factor E2F, altering its DNA-binding

727

properties and stimulating E1A-independent accumulation of E2 mRNA. J. Virol. 64:2345-

728

2359.

729

45.

730 731 732

Nevels M, Tauber B, Kremmer E, Spruss T, Wolf H, Dobner T. 1999. Transforming potential of the adenovirus type 5 E4orf3 protein. J. Virol. 73:1591-1600.

46.

Turner RL, Wilkinson JC, Ornelles DA. 2014. E1B and E4 oncoproteins of adenovirus antagonize the effect of apoptosis inducing factor. Virology 456-457:205-219.

34 733

47.

Goodrum FD, Ornelles DA. 1997. The early region 1B 55-kilodalton oncoprotein of

734

adenovirus relieves growth restrictions imposed on viral replication by the cell cycle. J.

735

Virol. 71:548-561.

736

48.

737 738

proteins in cell cycle-independent adenovirus replication. J. Virol. 73:7474-7488. 49.

739 740

50.

Hans F, Dimitrov S. 2001. Histone H3 phosphorylation and cell division. Oncogene 20:3021-3027.

51.

743 744

Ornelles DA, Broughton-Shepard RN, Goodrum FD. 2007. Analysis of adenovirus infections in synchronized cells. Methods Mol. Med. 131:83-101.

741 742

Goodrum FD, Ornelles DA. 1999. Roles for the E4 orf6, orf3, and E1B 55-kilodalton

Goodrum FD, Ornelles DA. 1998. p53 status does not determine outcome of E1B 55kilodalton mutant adenovirus lytic infection. J. Virol. 72:9479-9490.

52.

Cherubini G, Petouchoff T, Grossi M, Piersanti S, Cundari E, Saggio I. 2006.

745

E1B55K-deleted adenovirus (ONYX-015) overrides G1/S and G2/M checkpoints and

746

causes mitotic catastrophe and endoreduplication in p53-proficient normal cells. Cell Cycle

747

5:2244-2252.

748

53.

Cheng CY, Gilson T, Wimmer P, Schreiner S, Ketner G, Dobner T, Branton PE,

749

Blanchette P. 2013. Role of E1B55K in E4orf6/E1B55K E3 ligase complexes formed by

750

different human adenovirus serotypes. J. Virol. 87:6232-6245.

751

54.

Gupta A, Jha S, Engel DA, Ornelles DA, Dutta A. 2012. Tip60 degradation by

752

adenovirus relieves transcriptional repression of viral transcriptional activator EIA.

753

Oncogene.

754 755

55.

Schreiner S, Wimmer P, Groitl P, Chen SY, Blanchette P, Branton PE, Dobner T. 2011. Adenovirus type 5 early region 1B 55K oncoprotein-dependent degradation of

35 756

cellular factor Daxx is required for efficient transformation of primary rodent cells. J.

757

Virol. 85:8752-8765.

758

56.

Blackford AN, Patel RN, Forrester NA, Theil K, Groitl P, Stewart GS, Taylor AM,

759

Morgan IM, Dobner T, Grand RJ, Turnell AS. 2010. Adenovirus 12 E4orf6 inhibits

760

ATR activation by promoting TOPBP1 degradation. Proc. Natl. Acad. Sci. U. S. A.

761

107:12251-12256.

762

57.

Schwartz RA, Lakdawala SS, Eshleman HD, Russell MR, Carson CT, Weitzman MD.

763

2008. Distinct requirements of adenovirus E1b55K protein for degradation of cellular

764

substrates. J. Virol. 82:9043-9055.

765

58.

Baker A, Rohleder KJ, Hanakahi LA, Ketner G. 2007. Adenovirus E4 34k and E1b 55k

766

oncoproteins target host DNA ligase IV for proteasomal degradation. J. Virol. 81:7034-

767

7040.

768

59.

Schreiner S, Kinkley S, Burck C, Mund A, Wimmer P, Schubert T, Groitl P, Will H,

769

Dobner T. 2013. SPOC1-mediated antiviral host cell response is antagonized early in

770

human adenovirus type 5 infection. PLoS Pathog. 9:e1003775.

771

60.

772 773

acid substitution mutants of adenovirus type 5 E1B-55K protein. J. Virol. 75:4297-4307. 61.

774 775

Shen Y, Kitzes G, Nye JA, Fattaey A, Hermiston T. 2001. Analyses of single-amino-

Doorbar J, Quint W, Banks L, Bravo IG, Stoler M, Broker TR, Stanley MA. 2012. The biology and life-cycle of human papillomaviruses. Vaccine 30 Suppl 5:F55-70.

62.

Athanassiou M, Hu Y, Jing L, Houle B, Zarbl H, Mikheev AM. 1999. Stabilization and

776

reactivation of the p53 tumor suppressor protein in nontumorigenic revertants of HeLa

777

cervical cancer cells. Cell Growth Differ. 10:729-737.

778

63.

Fries KL, Miller WE, Raab-Traub N. 1996. Epstein-Barr virus latent membrane protein

36 779

1 blocks p53-mediated apoptosis through the induction of the A20 gene. J. Virol. 70:8653-

780

8659.

781

64.

782 783

apoptosis. Symp. Soc. Exp. Biol. 52:241-251. 65.

784 785

Holland AJ, Cleveland DW. 2009. Boveri revisited: chromosomal instability, aneuploidy and tumorigenesis. Nat. Rev. Mol. Cell Biol. 10:478-487.

66.

786 787

Degenhardt K, Perez D, White E. 2000. Pathways used by adenovirus E1B 19K to inhibit

Wolf F, Wandke C, Isenberg N, Geley S. 2006. Dose-dependent effects of stable cyclin B1 on progression through mitosis in human cells. EMBO J. 25:2802-2813.

67.

Zheng X, Rao XM, Gomez-Gutierrez JG, Hao H, McMasters KM, Zhou HS. 2008.

788

Adenovirus E1B55K region is required to enhance cyclin E expression for efficient viral

789

DNA replication. J. Virol. 82:3415-3427.

790

68.

Ou HD, Kwiatkowski W, Deerinck TJ, Noske A, Blain KY, Land HS, Soria C, Powers

791

CJ, May AP, Shu X, Tsien RY, Fitzpatrick JA, Long JA, Ellisman MH, Choe S,

792

O'Shea CC. 2012. A structural basis for the assembly and functions of a viral polymer that

793

inactivates multiple tumor suppressors. Cell 151:304-319.

794

69.

Patsalo V, Yondola MA, Luan B, Shoshani I, Kisker C, Green DF, Raleigh DP,

795

Hearing P. 2012. Biophysical and functional analyses suggest that adenovirus E4-ORF3

796

protein requires higher-order multimerization to function against promyelocytic leukemia

797

protein nuclear bodies. J. Biol. Chem. 287:22573-22583.

798

70.

King RW, Glotzer M, Kirschner MW. 1996. Mutagenic analysis of the destruction signal

799

of mitotic cyclins and structural characterization of ubiquitinated intermediates. Mol. Biol.

800

Cell 7:1343-1357.

801

71.

Hagting A, Den Elzen N, Vodermaier HC, Waizenegger IC, Peters JM, Pines J. 2002.

37 802

Human securin proteolysis is controlled by the spindle checkpoint and reveals when the

803

APC/C switches from activation by Cdc20 to Cdh1. J. Cell Biol. 157:1125-1137.

804

72.

805 806

oncolytic potential of the E1B-55K deletion mutant adenovirus. J. Virol. 83:2406-2416. 73.

807 808

Thomas MA, Broughton RS, Goodrum FD, Ornelles DA. 2009. E4orf1 limits the

Vakifahmetoglu H, Olsson M, Zhivotovsky B. 2008. Death through a tragedy: mitotic catastrophe. Cell Death Differ. 15:1153-1162.

74.

Sorensen CS, Syljuasen RG. 2012. Safeguarding genome integrity: the checkpoint

809

kinases ATR, CHK1 and WEE1 restrain CDK activity during normal DNA replication.

810

Nucleic Acids Res. 40:477-486.

811

75.

Carson CT, Schwartz RA, Stracker TH, Lilley CE, Lee DV, Weitzman MD. 2003. The

812

Mre11 complex is required for ATM activation and the G2/M checkpoint. EMBO J.

813

22:6610-6620.

814

76.

Carson CT, Orazio NI, Lee DV, Suh J, Bekker-Jensen S, Araujo FD, Lakdawala SS,

815

Lilley CE, Bartek J, Lukas J, Weitzman MD. 2009. Mislocalization of the MRN

816

complex prevents ATR signaling during adenovirus infection. EMBO J. 28:652-662.

817

77.

Loayza-Puch F, Drost J, Rooijers K, Lopes R, Elkon R, Agami R. 2013. p53 induces

818

transcriptional and translational programs to suppress cell proliferation and growth.

819

Genome Biol. 14:R32.

820

78.

Bassermann F, Eichner R, Pagano M. 2014. The ubiquitin proteasome system -

821

implications for cell cycle control and the targeted treatment of cancer. Biochim. Biophys.

822

Acta 1843:150-162.

823 824

79.

Bassermann F, von Klitzing C, Münch S, Bai RY, Kawaguchi H, Morris SW, Peschel C, Duyster J. 2005. NIPA defines an SCF-type mammalian E3 ligase that regulates

38 825 826

mitotic entry. Cell 122:45-57. 80.

827 828

repression domain to p53. Genes Dev. 8:190-202. 81.

829 830

Yew PR, Liu X, Berk AJ. 1994. Adenovirus E1B oncoprotein tethers a transcriptional

Hall AR, Dix BR, O'Carroll SJ, Braithwaite AW. 1998. p53-dependent cell death/apoptosis is required for a productive adenovirus infection. Nat. Med. 4:1068-1072.

82.

Hann B, Balmain A. 2003. Replication of an E1B 55-kilodalton protein-deficient

831

adenovirus (ONYX-015) is restored by gain-of-function rather than loss-of-function p53

832

mutants. J. Virol. 77:11588-11595.

833

83.

Acin S, Li Z, Mejia O, Roop DR, El-Naggar AK, Caulin C. 2011. Gain-of-function

834

mutant p53 but not p53 deletion promotes head and neck cancer progression in response to

835

oncogenic K-ras. J. Pathol. 225:479-489.

836

84.

837 838

driving oncogene in breast cancer. Carcinogenesis 33:2007-2017. 85.

839 840

Walerych D, Napoli M, Collavin L, Del Sal G. 2012. The rebel angel: mutant p53 as the

Tang Y, Luo J, Zhang W, Gu W. 2006. Tip60-dependent acetylation of p53 modulates the decision between cell-cycle arrest and apoptosis. Mol. Cell 24:827-839.

86.

Lee C, Hong B, Choi JM, Kim Y, Watanabe S, Ishimi Y, Enomoto T, Tada S, Kim Y,

841

Cho Y. 2004. Structural basis for inhibition of the replication licensing factor Cdt1 by

842

geminin. Nature 430:913-917.

843

87.

Dimaki M, Xouri G, Symeonidou IE, Sirinian C, Nishitani H, Taraviras S, Lygerou Z.

844

2013. Cell-cycle dependent subcellular translocation of the human DNA licensing inhibitor

845

Geminin. J. Biol. Chem.

846 847

88.

Ballabeni A, Zamponi R, Moore JK, Helin K, Kirschner MW. 2013. Geminin deploys multiple mechanisms to regulate Cdt1 before cell division thus ensuring the proper

39 848 849

execution of DNA replication. Proc. Natl. Acad. Sci. U. S. A. 110:E2848-2853. 89.

Kornitzer D, Sharf R, Kleinberger T. 2001. Adenovirus E4orf4 protein induces PP2A-

850

dependent growth arrest in Saccharomyces cerevisiae and interacts with the anaphase-

851

promoting complex/cyclosome. J. Cell Biol. 154:331-344.

852

90.

Roopchand DE, Lee JM, Shahinian S, Paquette D, Bussey H, Branton PE. 2001.

853

Toxicity of human adenovirus E4orf4 protein in Saccharomyces cerevisiae results from

854

interactions with the Cdc55 regulatory B subunit of PP2A. Oncogene 20:5279-5290.

855

91.

856 857

Shtrichman R, Kleinberger T. 1998. Adenovirus type 5 E4 open reading frame 4 protein induces apoptosis in transformed cells. J. Virol. 72:2975-2982.

92.

Marcellus RC, Lavoie JN, Boivin D, Shore GC, Ketner G, Branton PE. 1998. The

858

early region 4 orf4 protein of human adenovirus type 5 induces p53-independent cell death

859

by apoptosis. J. Virol. 72:7144-7153.

860

93.

Lavoie JN, Nguyen M, Marcellus RC, Branton PE, Shore GC. 1998. E4orf4, a novel

861

adenovirus death factor that induces p53-independent apoptosis by a pathway that is not

862

inhibited by zVAD-fmk. J. Cell Biol. 140:637-645.

863

94.

Pechkovsky A, Lahav M, Bitman E, Salzberg A, Kleinberger T. 2013. E4orf4 induces

864

PP2A- and Src-dependent cell death in Drosophila melanogaster and at the same time

865

inhibits classic apoptosis pathways. Proc. Natl. Acad. Sci. U. S. A. 110:E1724-1733.

866

95.

Araujo FD, Stracker TH, Carson CT, Lee DV, Weitzman MD. 2005. Adenovirus type

867

5 E4orf3 protein targets the Mre11 complex to cytoplasmic aggresomes. J. Virol.

868

79:11382-11391.

869 870

96.

Liu Y, Shevchenko A, Shevchenko A, Berk AJ. 2005. Adenovirus exploits the cellular aggresome response to accelerate inactivation of the MRN complex. J. Virol. 79:14004-

40 871 872

14016. 97.

873 874

adenovirus E4 ORF3 protein is required for viral replication. J. Virol. 79:6207-6215. 98.

875 876

Evans JD, Hearing P. 2005. Relocalization of the Mre11-Rad50-Nbs1 complex by the

Nelson G, Buhmann M, von Zglinicki T. 2009. DNA damage foci in mitosis are devoid of 53BP1. Cell Cycle 8:3379-3383.

99.

Giunta S, Jackson SP. 2011. Give me a break, but not in mitosis: the mitotic DNA

877

damage response marks DNA double-strand breaks with early signaling events. Cell Cycle

878

10:1215-1221.

879

100. Kelly TK, Miranda TB, Liang G, Berman BP, Lin JC, Tanay A, Jones PA. 2010.

880

H2A.Z maintenance during mitosis reveals nucleosome shifting on mitotically silenced

881

genes. Mol. Cell 39:901-911.

882

101. Tsuyama T, Tada S, Watanabe S, Seki M, Enomoto T. 2005. Licensing for DNA

883

replication requires a strict sequential assembly of Cdc6 and Cdt1 onto chromatin in

884

Xenopus egg extracts. Nucleic Acids Res. 33:765-775.

885 886 887

102. Symeonidou IE, Taraviras S, Lygerou Z. 2012. Control over DNA replication in time and space. FEBS Lett. 586:2803-2812. 103. Goto Y, Hirayama M, Takeda K, Tukamoto N, Sakata O, Kaeriyama H, Geshi M.

888

2013. Effect of synchronization of donor cells in early G1-phase using shake-off method

889

on developmental potential of somatic cell nuclear transfer embryos in cattle. Anim Sci J.

890

104. Kindsmuller K, Schreiner S, Leinenkugel F, Groitl P, Kremmer E, Dobner T. 2009. A

891

49-kilodalton isoform of the adenovirus type 5 early region 1B 55-kilodalton protein is

892

sufficient to support virus replication. J. Virol. 83:9045-9056.

893

105. Härtl B, Zeller T, Blanchette P, Kremmer E, Dobner T. 2008. Adenovirus type 5 early

41 894

region 1B 55-kDa oncoprotein can promote cell transformation by a mechanism

895

independent from blocking p53-activated transcription. Oncogene 27:3673-3684.

896

106. Teodoro JG, Halliday T, Whalen SG, Takayesu D, Graham FL, Branton PE. 1994.

897

Phosphorylation at the carboxy terminus of the 55-kilodalton adenovirus type 5 E1B

898

protein regulates transforming activity. J. Virol. 68:776-786.

899

107. Wimmer P, Blanchette P, Schreiner S, Ching W, Groitl P, Berscheminski J, Branton

900

PE, Will H, Dobner T. 2013. Cross-talk between phosphorylation and SUMOylation

901

regulates transforming activities of an adenoviral oncoprotein. Oncogene 32:1626-1637.

902 903

108. Koyuncu OO, Dobner T. 2009. Arginine methylation of human adenovirus type 5 L4 100-kilodalton protein is required for efficient virus production. J. Virol. 83:4778-4790.

42 FIGURE LEGENDS 904

FIGURE 1. Click-iT EdU labeling protocol and DNA profile analysis of EdU-labeled cells.

905

(A) HeLa cells were labeled with Click-iT EdU for six 4-h periods (indicated by arrows) over the

906

course of 24 h prior to infection. At time 0 h, EdU-labeled cultures were infected with the E1B-

907

55K/E4orf3 double-mutant virus at an MOI of 10 and processed for immunofluorescence at 72

908

hpi or were collected for flow cytometric analysis. Shading indicates the phase of the cell cycle

909

the EdU-labeled cells are expected to have reached at infection (S phase, gray; G2/M, black; G1,

910

white). (B) Samples of cells labeled with Click-iT EdU were collected at the time of infection (0

911

h) and stained with propidium iodide for DNA content and for EdU as described in the Materials

912

and Methods. The shaded density profile represents the DNA profile for the entire population of

913

cells. The unshaded density profile shows the DNA profile of EdU-positive cells.

914

FIGURE 2. Early G1 cells give rise to mitotic-like nuclei after infection with the E1B-

915

55K/E4orf3 double-mutant virus. Synchronously dividing HeLa cells were obtained by mitotic

916

shake and hydroxyurea selection as described in the Materials and Methods. Synchronously

917

dividing cells were either harvested for DNA profile analysis or infected at the indicated time

918

after entering S phase with the E1B-55K/E4orf3 double-mutant virus at an MOI of 10. (A) Cell

919

cycle distribution by DNA profile analysis. (B) At 72 hpi, cells were stained for DNA with DAPI

920

and evaluated by fluorescence microscopy to determine the frequency of cells with condensed

921

DNA. The stage of the cell cycle at the time of infection is indicated below. A representative

922

experiment is shown of three that were performed with overlapping times of infection after

923

entering S phase.

924

FIGURE 3. HeLa cells infected with the E1B-55K/E4orf3 double-mutant virus show

43 925

evidence of mitotic distress. HeLa cells were mock-infected or infected at an MOI of 10 with

926

the indicated viruses and stained at 72 h post infection with DAPI to visualize DNA (blue),

927

antibodies to phospho-H3 as a marker of early mitosis (red), and antibodies to β-tubulin to

928

visualize mitotic spindles (green). (A) Representative images of (a-b) mock, (c) wild-type, and

929

(d-e) E1B-55K/E4orf3-mutant viral infections are shown. (B) The frequency of cells containing

930

mitotic-like condensed DNA, staining for phospho-H3 and asymmetrically stained mitotic

931

spindles was quantified for at least 500 cells for each viral infection. A representative experiment

932

of three is shown. Error bars indicate the 95% exact binomial confidence interval for the

933

representative experiment. (C) HeLa cells were infected with the E1B 55K/E4orf3 double-

934

mutant virus at an MOI of 10. Cells were stained for DNA at the indicated times post infection

935

and the frequency of mitotic-like cells was determined for approximately 500 cells at each time

936

point. The broken symbols for times beyond 72 hpi indicate imprecision due to the probable loss

937

of cells because of death or detachment. The experiment shown is representative of three

938

independent experiments with similar outcomes. Error bars indicate the 95% exact binomial

939

confidence interval for the representative experiment.

940

FIGURE 4. Colcemid traps cells infected with the E1B-55K-mutant virus in a mitotic-like

941

state. HeLa cells were infected at an MOI of 10 with the indicated viruses and treated with

942

colcemid for 12 h at the indicated times post-infection before being fixed and stained with DAPI

943

to visualize DNA. The frequency of mitotic-like cells was determined for approximately 500

944

cells for each virus at each time point.

945

FIGURE 5. The ability to inhibit entry into a mitotic-like state maps to the p53-binding

946

ability of the E1B-55K protein. (A) HeLa cells were infected at an MOI of 10 with the

947

indicated E1B 55K-null viruses and treated with colcemid or vehicle control 12 h prior to

44 948

staining with DAPI at 72 hpi to visualize DNA. The frequency of mitotic-like cells was

949

determined by fluorescence microscopy. A representative experiment of three independent

950

experiments is shown. For each of the E1B-55K-null viruses, colcemid significantly increased

951

the fraction of mitotic-like cells (p-value < 10-6 by Fisher’s exact test). Error bars indicate the

952

95% exact binomial confidence interval for the representative experiment. (B) HeLa cells were

953

infected at an MOI of 10 with the indicated viruses bearing point mutations in the E1B-55K gene

954

(described in Table 2) and treated with colcemid 12 h prior to staining at 72 hpi with DAPI to

955

visualize DNA. The frequency of mitotic-like cells is shown for each mutant infection. The

956

proportion of mitotic-like nuclei was non-randomly distributed among the 11 virus-infected

957

samples exclusive of the E1B-55K-null virus H5pm4149 (p < 0.0001, Chi-squared test). The

958

Chi-squared test was repeated after systematically excluding each sample. The p-value was non-

959

significant (p = 0.56) only when the virus H5pm4109 was excluded from the analysis. This

960

analysis was again repeated by excluding individual samples in the collection that also excluded

961

H5pm4109. The Chi-squared test reported non-significant p-values for each subset lacking

962

H5pm4109 and one other sample. This indicates that among the non-null viruses, only

963

H5pm4109-infected cells exhibited a significant change in the number of condensed nuclei in the

964

presence of colcemid. Results from a representative experiment of three independent

965

experiments with similar outcomes are shown. (C) p53-null H1299 cells were transfected with a

966

plasmid to express p53 24 h before being mock-infected or infected at an MOI of 10 with either

967

the E1B-55K single-mutant or the E1B-55K/E4orf3 double-mutant virus. The cells were either

968

left untreated or treated with 0.2 μg per ml of colcemid 12 h prior to immunostaining for p53

969

and visualizing DNA with DAPI staining at 72 hpi. Cells were classified as either p53-positive or

970

negative and the frequency of mitotic-like cells was determined. As noted in the text, condensed

45 971

nuclei in mock-infected cells resembled pycnotic nuclei of apoptotic cells. The number of p53-

972

positive virus-infected cells with condensed DNA was increased significantly over p53-negative

973

cells (p = 2×10-6 by Fisher’s exact test).

974

FIGURE 6. Cyclin B1 levels are elevated during adenoviral infections. HeLa cells were

975

mock-infected or infected with the indicated viruses at an MOI of 10. Cellular lysates were

976

collected in the presence of protease and phosphatase inhibitors at 72 hpi. Material from identical

977

numbers of infected cells were separated by SDS-PAGE, transferred to a nitrocellulose

978

membrane and immunoblotted for (A) cyclin B1 and (B) β-actin. An overexposed β-actin blot is

979

presented here to permit visualization of the weaker signals. Non-saturated exposures were used

980

for quantitative analyses. The position of a cyclin B1-related product of 35 kDa is indicated by

981

the arrowhead. (C) The optical density of the signal for the intact cyclin B1 products was

982

quantified, normalized to β-actin, and then normalized to the value measured from mock-infected

983

cells in three independent experiments. The mean and standard deviation are plotted on a log-

984

scale. Application of the t-test to log-transformed values shows that levels of cyclin B1 in were

985

significantly greater than mock-infected cells in dl309- and dl1520-infected cells (p80%) cells expressing both

1015

fusion protein and E4orf3 as indicated in panels a and b.

1016

47

TABLES TABLE 1. Edu-positive and mitotic-like double-mutant virus-infected cells EdU-positive cells (%)b

Designationc

Mitotic-like Edu-positive cells (%)

a

Labeling Period

G1

S

G2/M

-2 to +2

27

52

21

mid-S

25.6

-6 to -2

6

15

79

G2/M

69.6

-10 to -6

35

7

58

M, early-G1

50.0

-14 to -10

68

5

27

mid G1

20.0

-18 to -14

61

25

14

late G1

0.0

-22 to -18

56

25

20

G1, early-S

5.8

a

Hours exposed to EdU before being infected at an MOI of 10 at 0 h.

b

Percentage of EdU-positive cells with DNA content characteristic of the indicated stage of the

cell cycle. c

Predominant stage of the cell cycle for EdU-positive cells at 0 h.

48 TABLE 2. E1B-55K mutant viruses and key characteristics Virus

E1B-55K mutation

Likely defect

Reference

dl110

deletion, frame shift

null

(42)

dl1520

stop codon, deletion

null

(39)

dl338

deletion

null

(41)

4149

four stop codons

null for all 55K-related

(104)

4100

none

none

(43)

4108

T255A

E4orf6-binding

(Identical mutation to virus described in 60)

4109

H260A

p53-binding, E4orf6binding

(Identical mutation to virus described in 60)

4127

C454S/C456S

Mre11 binding

(105)

4174

S490A/S491A/T495A

Non-phospho C-term

(10, 106, 107)

4185

P70T/S73A

Unknown (P70T/S73A)

(108)

4197

E472A

Daax interaction 1

(55)

4198

K185A/K187A

Daax interaction 2

(55)

4216

GVVI233-236AAAS

SUMO interaction

Not published

4217

V339A/V341A/I342S

SUMO interaction

Not published

4227

S490D/S491D/T494D

Phosphomimetic C-term

(57, 107)

Adenovirus replaces mitotic checkpoint controls.

Infection with adenovirus triggers the cellular DNA damage response, elements of which include cell death and cell cycle arrest. Early adenoviral prot...
1MB Sizes 0 Downloads 6 Views