Download Open the publication - UEF Electronic Publications

Document related concepts

Gene therapy of the human retina wikipedia , lookup

Canine parvovirus wikipedia , lookup

Henipavirus wikipedia , lookup

Vectors in gene therapy wikipedia , lookup

Index of HIV/AIDS-related articles wikipedia , lookup

Transcript
Janne Ruotsalainen
Interferon Response as a Challenge
and Possibility for Developing
Alphavirus Based Oncolytic
Virotherapy
Publications of the University of Eastern Finland
Dissertations in Health Sciences
JANNE RUOTSALAINEN
Interferon Response as a Challenge and
Possibility for Developing Alphavirus Based
Oncolytic Virotherapy
To be presented by permission of the Faculty of Health Sciences, University of Eastern
Finland for public examination in Tietoteknia auditorium (TTA), Kuopio, on Friday,
September 26th 2014, at 1 p.m.
Publications of the University of Eastern Finland
Dissertations in Health Sciences
Number 247
Department of Biotechnology and Molecular Medicine, A.I. Virtanen Institute for Molecular
Sciences, Faculty of Health Sciences, University of Eastern Finland
Kuopio
2014
Grano
Kuopio, 2014
Series Editors:
Professor Veli-Matti Kosma, M.D., Ph.D.
Institute of Clinical Medicine, Pathology
Faculty of Health Sciences
Professor Hannele Turunen, Ph.D.
Department of Nursing Science
Faculty of Health Sciences
Professor Olli Gröhn, Ph.D.
A.I. Virtanen Institute for Molecular Sciences
Faculty of Health Sciences
Professor Kai Kaarniranta, M.D., Ph.D.
Institute of Clinical Medicine, Ophthalmology
Faculty of Health Sciences
Lecturer Veli-Pekka Ranta, Ph.D. (pharmacy)
School of Pharmacy
Faculty of Health Sciences
Distributor:
University of Eastern Finland
Kuopio Campus Library
P.O.Box 1627
FI-70211 Kuopio, Finland
http://www.uef.fi/kirjasto
ISBN (print): 978-952-61-1554-2
ISBN (pdf): 978-952-61-1555-9
ISSN (print): 1798-5706
ISSN (pdf): 1798-5714
ISSN-L: 1798-5706
III
Author’s address:
Department of Biotechnology and Molecular Medicine
A.I. Virtanen Institute for Molecular Sciences
Doctoral Program of Molecular Medicine
University of Eastern Finland
P.O. Box 1627
FI-70211 KUOPIO
FINLAND
[email protected]
Supervisors:
Professor Ari Hinkkanen, Ph.D.
Department of Biotechnology and Molecular Medicine
A.I. Virtanen Institute for Molecular Sciences
University of Eastern Finland
KUOPIO
FINLAND
Professor Kari Airenne, Ph.D.
Department of Biotechnology and Molecular Medicine
A.I. Virtanen Institute for Molecular Sciences
University of Eastern Finland
KUOPIO
FINLAND
Reviewers:
Professor Ilkka Julkunen, Ph.D.
Department of Virology
University of Turku
TURKU
FINLAND
Docent Vincenzo Cerullo, Ph.D.
Faculty of Pharmacy
Division of Biopharmaceutics and Pharmacokinetics
University of Helsinki
HELSINKI
FINLAND
Opponent:
Cristian Smerdou, Ph.D.
Division of Gene Therapy
Center for Applied Medical Research (CIMA)
University of Navarra
PAMPLONA
SPAIN
IV
V
Ruotsalainen, Janne
Interferon Response as a Challenge and Possibility for Developing Alphavirus Based Oncolytic Virotherapy
University of Eastern Finland, Faculty of Health Sciences
Publications of the University of Eastern Finland. Dissertations in Health Sciences 247. 2014. 104 p.
ISBN (print): 978-952-61-1554-2
ISBN (pdf): 978-952-61-1555-9
ISSN (print): 1798-5706
ISSN (pdf): 1798-5714
ISSN-L: 1798-5706
ABSTRACT
Certain malignancies such as glioblastoma multiforme are still incurable and in others like
colorectal carcinoma, the prognosis greatly depends on the time of diagnosis. Oncolytic
virotherapy is a novel approach to address these difficult forms of cancer. In addition to
direct lysis viruses can induce innate and adaptive immune responses against the tumor or
its vasculature. However, the infection can be prematurely halted by neutralizing
antibodies or components of the innate immune system, especially the type I interferon
response. Semliki Forest virus based oncolytic virus VA7 has previously been proven to be
a promising candidate for the treatment of several different forms of cancer showing
significant therapeutic power especially against melanoma xenografts. However, only local
virus injections were powerful against A549 lung adenocarcinoma xenografts, and tumor
resistance to virus was associated with the type I interferon response. Furthermore, rat
BT4C gliomas were resistant to VA7 in vivo and the treated animals developed neutralizing
antibodies against the virus. The initial aim of this thesis work (I-III) was to investigate
whether peripherally administered VA7 could home into intracranial gliomas, when the
tumors per se are virus susceptible. U87 gliomas were discovered to be highly sensitive to
VA7 infection in vitro and a single intravenous virus injection resulted in the tumor
infection and a cure in 16/17 of treated mice (I). Next the role of adaptive immune responses
to virotherapy was examined by treating intracranial mouse GL261 gliomas with VA7.
Neither intravenous nor local virus injections improved survival even when adaptive
immunity was suppressed with cyclophosphamide or rapamycin. Infected GL261 tumor
tissue could not be detected and surprisingly, cyclophosphamide administration increased
infection of the healthy brain parenchyma. GL261 cells could secrete and respond to type I
interferon upon VA7 infection and host-derived interferons likely had an additional role in
the resistance as pre-infected GL261 tumors developed with normal kinetics (II). The
contributions of innate and adaptive immune responses to VA7 therapy were studied in
more detail by utilizing two CT26 mouse colon carcinoma cell clones found to differ
dramatically in both their type I IFN responses and their immunogenicities. Intradermal
CT26LacZ tumors were immunogenic but they could be eradicated both from Balb/c and
SCID mice and VA7 injection resulted in a progressive perivascular tumor infection. When
CT26WT and CT26LacZ cells were mixed, as few as 10 % of the resistant CT26WT cells
were sufficient to compromise the therapy. Comparative RNA-Seq transcriptome analysis
revealed that in contrast to CT26LacZ, the CT26WT cells constitutively expressed several
interferon stimulated genes (III). In conclusion, VA7 demonstrated potential in targeting of
gliomas and colorectal carcinomas but the treatment efficacy depended primarily on the
tumor type I interferon responsiveness. Thus, careful pre-screening will be necessary to
indentify the susceptible candidates for alphavirus based oncolytic virotherapy.
National Library of Medicine Classification: QZ 380, WI 529, QW 940, QW 168.5.A7, QU 560
Medical Subject Headings: Oncolytic Virotherapy/methods; Oncolytic Viruses; Colorectal
Neoplasms/therapy; Glioma/therapy; Immunotherapy; Gene Therapy; Alphavirus/genetics; Semliki forest
virus; Interferon Type I/immunology; Mice
VI
VII
Ruotsalainen, Janne
Interferonivaste haasteena ja mahdollisuutena Alfaviruspohjaisen onkolyyttisen viroterapian kehittämisessä
Itä-Suomen yliopisto, terveystieteiden tiedekunta
Itä-Suomen yliopiston julkaisuja. Terveystieteiden tiedekunnan väitöskirjat 247. 2014. 104 s.
ISBN (print): 978-952-61-1554-2
ISBN (pdf): 978-952-61-1555-9
ISSN (print): 1798-5706
ISSN (pdf): 1798-5714
ISSN-L: 1798-5706
TIIVISTELMÄ
Jotkut syövät kuten glioblastoma multiforme ovat parantumattomia, ja toisissa kuten
paksu- ja peräsuolen syövissä ennuste riippuu diagnoosin varhaisuudesta. Onkolyyttinen
viroterapia on uusi lähestymistapa näiden vakavien syöpätyyppien hoidossa. Onkolyyttiset
virukset tuhoavat suoraan syöpäkudosta, minkä lisäksi ne voivat herättää potilaan
luontaisen ja hankitun immuniteetin kasvainta tai sen verisuonistoa vastaan. Neutraloivat
vasta-aineet tai luontainen immuniteetti (etenkin tyypin I interferonivaste) voivat kuitenkin
keskeyttää virus-infektion ennenaikaisesti. Semliki Forest virukseen pohjautuva
onkolyyttinen virus VA7 on osoittautunut aiemmissa tutkimuksissa lupaavaksi
hoitomuodoksi monissa syöpämalleissa ollen erityisen tehokas ihmisen melanooma
ksenotransplantaatteja vastaan. Ainoastaan paikalliset virushoidot toimivat kuitenkin A549
keuhkoadenokarsinooma ksenotransplantaattimallissa, jossa resistenssi systeemisesti
annetulle virukselle yhdistettiin tyypin I interferonivasteeseen. Rotan BT4C glioomamalli
oli myös resistentti ja viroterapiaa saaneiden eläinten verestä löytyi virusvasta-aineita.
Tämän väitöskirjan (I-III) ensimmäisen osatyön tavoitteena oli selvittää hakeutuuko VA7
kallon sisäisiin glioomiin, jos kasvaimet sinänsä ovat virukselle herkkiä. U87 glioomien
havaittiin olevan hyvin herkkiä VA7 infektiolle in vitro ja yksi virusannos laskimoon johti
kasvaimen infektoitumiseen ja 16/17 hiiren paranemiseen (I). Seuraavaksi hankitun
immuniteetin vaikutuksia viroterapiaan tutkittiin hoitamalla hiiren kallon sisäisiä GL261
glioomia. Laskimonsisäiset tai paikalliset VA7-injektiot eivät parantaneet selviytymistä,
eikä immuniteetin vaimentaminen syklofosfamidilla tai rapamysiinillä parantanut
hoitotulosta. VA7 ei infektoitunut GL261 kasvainkudosta, mutta syklofosfamidi tehosti
infektiota terveessä aivokudoksessa. Infektoitujen GL261 solujen havaittiin erittävän tyypin
I interferonia ja olevan myös herkkiä interferonille. Hiiren tuottamilla interferoneilla oli
luultavasti lisävaikutus sillä etukäteen infektoidut GL261 kasvaimet kehittyivät myös
normaalisti (II). Viimeisessä osatyössä selvitettiin tarkemmin synnynnäisen ja hankitun
immuniteetin vaikutuksia VA7 terapiaan käyttämällä kahta CT26 hiiren paksu- ja
peräsuolisyövän
solukloonia,
joiden
havaittiin
eroavan
merkittävästi
immunogeenisyydeltään ja tyypin I interferonivasteeltaan. Nahkansisäiset CT26LacZ
kasvaimet olivat immunogeenisiä, mutta ne kyettiin hävittämään sekä Balb/c että SCID
hiiristä ja kasvaininfektion huomattiin keskittyvän verisuonien ympärille. 10 % osuus
resistenttejä CT26WT soluja riitti heikentämään hoitotehoa merkittävästi, ja CT26WT
solujen havaittiin ilmentävän perustasolla monia interferoni-stimuloituvia geenejä.
Yhteenvetona VA7 osoittautui lupaavaksi hoitomuodoksi sekä gliooma- että paksu- ja
peräsuolen
syöpämalleissa,
mutta
hoitoteho
riippui
kasvainten
tyypin
I
interferonivasteesta. Havainnot puoltavat seulontamenetelmien kehittämistä niiden
kasvaimien löytämiseksi, jotka voisivat olla herkkiä alfaviruspohjaiselle onkolyyttiselle
viroterapialle.
Luokitus: QZ 380, WI 529, QW 940, QW 168.5.A7, QU 560
Yleinen Suomalainen asiasanasto: syöpätaudit, glioomat, paksusuolisyöpä, peräsuolisyöpä, syöpähoidot,
viroterapia, immuunihoito, Semliki Forest –virus
VIII
IX
Acknowledgements
The studies described in this thesis were carried out in the Department of Biotechnology
and Molecular Medicine, A. I. Virtanen Institute for Molecular Sciences, Faculty of Health
Sciences, University of Eastern Finland.
First of all, I want to sincerely thank my main supervisor Professor Ari Hinkkanen for
providing me with the opportunity to work in his laboratory on such an exciting research
topic. I have been very lucky to have you as a supervisor. Already at the very earliest
stages, you have demonstrated through your personal example the necessary elements of
research: curiosity, dedication, patience and genuine excitement in the face of scientific
discovery, big or small. I have learned that the only route to success in research goes
through a series of disappointments and your great sense of humor has made also those
easier to cope with. I want to thank also my second supervisor Kari Airenne for the
inspiring scientific talks we have had and for the many memorable national and
international conference trips we have shared. I want to express my warm gratitude to the
pre-examiners of the thesis work Professor Ilkka Julkunen and Docent Vincenzo Cerullo for
the careful inspection of the work. In addition, I am truly thankful to Ewen MacDonald for
the professional linguistic revision.
I express my deepest gratitude to the many people who have been directly involved in this
work: Miika Martikainen, Markus Vähä-Koskela, Jari Heikkilä, Minna Niittykoski, Minna
Kaikkonen, Seppo Ylä-Herttuala, John Bell, Tuulia Huhtala, Tytti Aaltonen, Ivana Kholova,
Olli Gröhn, Johanna Närväinen, Timo Liimatainen, Chantal Lemay, Julie Cox, Naomi De
Silva, Agnieszka Kus, Theresa Falls, Jean-Simon Diallo and Fabrice Le Boeuf. Especially I
want to thank Miika and Markus. Miika: thank you for putting up with me and my
“humor” all these years without completely losing it. I will be forever grateful not only for
your invaluable scientific contribution but also for your friendship. Markus: I want to thank
you for the crucial impact you had on this work and for the dozens of mind-boggling
scientific talks we have had. You are a great person and a great scientist and it has been a
privilege to work with you. Irrespective of whether or not the work has been published, I
also want acknowledge the contributions of co-authors and colleagues Erkko Ylösmäki, Ale
Närvänen, Diana Schenkwein, Emilia Makkonen, Outi Rautsi, Heini Belt, Kirsi Vuorinen,
Maartje Van Geenen, Marianna Héllen, Ville Ruottinen, Jari Pokka, Bijay Dhungel, Jakub
Kolodziejski, Michal Burmistrz, Justin Zoppe, Seppo Rönkkö, Leena-Sisko Johansson,
Kristiina Järvinen, Jukka Seppälä, Karoliina Autio, Marjukka Anttila, Matti Waris, Akseli
Hemminki, Ari Kauppinen, Anssi-Pekka Karttunen, Riikka Pellinen and Jarkko Ketolainen.
I want to thank Helena Pernu, Hanne Tanskanen, Jouko Mäkäräinen, Seija Sahrio, Jari
Nissinen, Laila Kaskela, Mirka Tikkanen, Anneli Miettinen, Anne Martikainen, Joonas
Malinen and the whole staff of LAC for the administrative and technical advice and
assistance without which this work would have not been possible. I wish to thank all the
friends and colleagues from the former Ark Therapeutics where I did my Master’s thesis for
the fun and stimulating working environment, which encouraged me to continue to pursue
X
a scientific career. Thank you Vesa, Taina, Miia, Tytteli, Tiina, Pyry, Hanna, Mikko, AnnMarie, Tommi, Jere, Haritha, Thomas, Roseanne, Anssi, Tarja, the late Riikka, Siiri, Juha and
all the rest. I have continued to bother many of you since those times but have never been
turned down. My special thanks go to Ann-Marie, who had an important role at a crucial
moment when I still had doubts about which path to follow. I want to thank my high
school biology teacher, Esa Hämäläinen, whose inspiring way of teaching was the first
spark to engage myself with the mysterious world of biosciences. I owe not only my
warmest gratitude but also sincerest apologies to my new and long-term friends and
relatives alike; Jarkko, Marko, Aleksi, Jussi, Rebekka, Antti, Susanna, Weronika, Magda,
Stefanie, Patrick, Sören, Anja, Pyare, Iina, Mikko, Miia, Ville, Heli, Otto, Tuula, Pentti, Jyrki,
Tanja, Make and all the others who I have been neglecting especially during the past year.
Thank you for not walking out on me. Without friends, life would be miserable.
My deepest gratitude goes to my family members, the late Tarja Ruotsalainen, Tauno
Lösönen, Juhani Ruotsalainen, Päivi Ruotsalainen, Joonas Lösönen, Jari Ruotsalainen, Katja
Ruotsalainen, Miia Tolvanen and Jari Ollikainen. Without your support, I could have not
gone through this and I could not imagine myself or my life without your existence and
influence. I dedicate greatest gratitude and respect to my late mother Tarja, who was not
only a great personality and a loving mother but also encouraged me to keep pushing the
boundaries and educate myself to the highest degree. I also want to thank Katja for being
the big sister when I needed one and for setting me on the right track at the critical
moment. I would have been even more unbearable as a philosopher than I currently am as
a bioscientist.
Finally, I want to dedicate my heartfelt gratitude to my wife, medical doctor Jenny Schmidt
for sticking with me during this process. Without your love, support, encouragement and
trust I could have not kept going. I apologize for how little time I have had for you during
the last year. I hope time will prove that all of this was worth it.
Kuopio, August 2014
Janne Ruotsalainen
This work was supported by Doctoral Program in Molecular Medicine, Orion Foundation,
Matti and Vappu Maukonen fund, Northern Savo Regional Fund, Academy of Finland,
EVO fund, Cancer Society of Finland and the Cancer Center of the University of Eastern
Finland
XI
List of the original publications
This dissertation is based on the following original publications:
I
II
III
Heikkilä, J., Vähä-Koskela, M., Ruotsalainen, J., Martikainen, M., Stanford, M.
McCart, J.,Bell, J., Hinkkanen, A. Intravenously Administered Alphavirus Vector
VA7 Eradicates Orthotopic Human Glioma Xenografts in Nude Mice. PLoS One.
6;5(1):e8603, 2010
Ruotsalainen, J.*, Martikainen, M.*, Niittykoski, M., Huhtala, T., Aaltonen, T.,
Heikkilä, J., Bell, J., Vähä-Koskela, M., Hinkkanen, A. Interferon- sensitivity of
tumor cells correlates with poor response to VA7 virotherapy in mouse glioma
models. Molecular Therapy. 20(8):1529-39, 2012. Published online 2012 March 20. doi:
10.1038/mt.2012.53.
Ruotsalainen, J., Kaikkonen, M., Niittykoski, M., Martikainen, M., Lemay, C.,
Cox, J., De Silva, N., Kus, A., Falls, T., Diallo, J-S, Le Boeuf, F., Bell, J., YläHerttuala, S., Hinkkanen, A., Vähä-Koskela, M., Clonal Variation in Interferon
Response Determines the Outcome of Oncolytic Virotherapy in Mouse CT26
Colon Carcinoma Model, Gene Therapy, in press
* Equal contribution
Also unpublished data is presented
The publications were adapted with the permission of the copyright owners.
XII
XIII
Contents
1 Introduction .................................................................................................................................................. 1
2 Literature review ........................................................................................................................................... 3
2.1 Cancer ....................................................................................................................................................... 3
2.1.1 Hallmarks of cancer ....................................................................................................................... 3
2.1.2 Oncogenesis, tumor microenvironment and cancer stem cell theory......................................... 3
2.1.3 Cancer immunoediting: Sum of immunosculpting and immunosuppression .......................... 5
2.1.4 Cancer in Finland and modern cancer therapies ......................................................................... 7
2.2 Interferons in cell autonomous, innate and adaptive immunity ........................................................... 9
2.3 Alphaviruses........................................................................................................................................... 15
2.3.1 Taxonomy, natural life cycle and
pathogenesis of Alphaviruses and Semliki Forest virus ..................................................................... 15
2.3.2 Alphavirus replication cycle ........................................................................................................ 17
2.3.3 Alphaviruses and adaptive immunity ....................................................................................... 19
2.3.4 Alphaviruses and innate- and cell autonomous immunity........................................................ 19
2.4 Oncolytic virotherapy ............................................................................................................................ 23
2.4.1 From the early 20th century to the present day .......................................................................... 23
2.4.2 Oncolytic virotherapy overlaps with gene therapy and immunotherapy of cancer
– emerging concept of oncolytic immunovirotherapy ....................................................................... 28
2.4.3 Combination therapy to improve OV efficacy ............................................................................ 37
2.4.4 Targeting of Oncolytic Viruses..................................................................................................... 39
2.4.5 Naturally targeted oncolytic viruses............................................................................................ 43
2.4.6 VA7 oncolytic virus....................................................................................................................... 46
2.4.7 Armed oncolytic viruses............................................................................................................... 47
2.4.8 Safety and ethics of oncolytic virotherapy .................................................................................. 48
3 Aims of the study ........................................................................................................................................ 51
4 Materials ans methods ................................................................................................................................ 53
4.1 Cell lines and mouse strains (I-III) ........................................................................................................ 53
4.2 VA7-EGFP production (I-III) ................................................................................................................. 54
4.3 Determination of Semliki Forest virus infectious titers (I-III).............................................................. 56
4.4 Preparing mice and cells for tumor transplantations (I-III) ................................................................. 56
4.5 SDS-PAGE and western blotting (III).................................................................................................... 57
4.6 SFV detection from tissues by immunohistochemistry (I-III) ............................................................. 58
5 Results .......................................................................................................................................................... 59
5.1 Peripherally administered VA7 homes into orthotopic U87-Fluc glioma xenografts
and destroys them (I) ................................................................................................................................... 59
5.2 The VA7 resistance of GL261 mouse gliomas reflects their robust type I IFN signaling (II)............. 60
5.3 Subclone of resistant CT26 colon carcinoma is highly sensitive to VA7
due to weak type I IFN response (III) ........................................................................................................ 64
6 Discussion .................................................................................................................................................... 69
7 Summary and conclusions.......................................................................................................................... 83
References ....................................................................................................................................................... 85
APPENDIX: Original publications I-III
XIV
XV
Abbreviations
AAV
Ab
arbovirus
Ad
ADCC
APC
BBB
BHK
CAR
CEA
CHIKV
CIK
CSC
CT
CTL
DAI
dsRNA
ECM
EGF
EEEV
EGFP
EGFR
EMT
GBM
GM-CSF
CML
CPA
HCC
HDAC
HGF
HPV
HSV
i.c.
adeno-associated virus
antibody
arthropod-borne virus
adenovirus
antibody-dependent
cellular cytotoxicity
antigen presenting cell
blood brain barrier
baby hamster kidney cells
coxsackie virus and
adenovirus receptor
carcinoembryonic antigen
chikungunya virus
cytokine induced killer cells
cancer stem cells
computed tomography
cytotoxic T lymphocytes
DNA-dependent activator of
IFN-regulatory factors
double strand RNA
extracellular matrix
epidermal growth factor
Eastern equine encephalitis
virus
enhanced green fluorescent
protein
epidermal growth factor
receptor
epithelial-to-mesenchymal
transition
glioblastoma multiforme
granulocyte-macrophage
colony-stimulating factor
chronic myeloid leukemia
cyclophosphamide
hepatocellular carcinoma
histone deacetylase
hepatocyte growth factor
human papilloma virus
Herpes simplex virus
intracranial
ICD
IFNAR
IL
i.p.
i.t.
i.v.
IFN
IL
IPA
IRG
ISGF3
LAMR
Luc
mAb
MCA
MDSC
MET
miRNA
MLV
MMP
MOG
MRI
Nab
NDV
NIS
NK-cell
NKT-cell
NLS
nsP
SFV
TGF
TNF
VEEV
OV
pDC
immunogenic cell death
Interferon-alpha/beta receptor
interleukin
intraperitoneal
intratumoral
intravenous
interferon
interleukin
Ingenuity pathway analysis
immune-related GTPase
Interferon-stimulated gene
factor 3
high-affinity laminin receptor
luciferase
monoclonal antibody
3-methylcholanthrene
myeloid derived suppressor
cell
mesenchymal-to-epithelial
transition
microRNA
murine leukemia virus
matrix metalloproteinase
myelin oligodendrocyte
glycoprotein
magnetic resonance imaging
neutralizing antibody
Newcastle disease virus
sodium iodine symporter
natural killer cell
natural killer T-cell
nuclear localization signal
nonstructural protein
Semliki Forest virus
transforming growth factor
tumor necrosis factor
Venezuelan equine
encephalitis virus
oncolytic virotherapy
plasmacytoid dendritic cell
XVI
PET
PFU
p.i.
poly I:C
PSA
p.t.i.
RdRp
RLR
RNAi
SCID-X1
SPECT
Tc
Th
TIMP
Treg
SAV
SFV
SMI
ssRNA
STAT
SINV
VEGF
VPA
VSV
WEEV
positron emission
tomography
plaque forming units
post infection
Polyinosinic:polycytidylic
acid
prostate specific antigen
post tumor induction
RNA-dependent RNA
polymerase
RIG-I like receptor
RNA interference
X-linked severe-combined
immunodeficiency
Single-photon emission
computed tomography
cytotoxic T-cell
T-helper cell
tissue inhibitor of
metalloproteinases
T-regulatory cell
salmonid alphavirus
Semliki Forest virus
small molecule inhibitor
single strand RNA
Signal Transducer and
Activator of Transcription
Sindbis virus
vascular endothelial growth
factor
valproic acid
vesicular stomatitis virus
Western equine encephalitis
virus
1 Introduction
Due to improved diagnostics and novel therapeutics, the prognoses of many malignancies
including breast cancer, prostate cancer, testicular cancer and childhood leukemia have
improved dramatically in western countries such as Finland (De Angelis et al. 2014;
Pukkala, Sankila, Rautalahti 2011). For example, in Finland in the period 2007-2009, the 5year survival from the most common cancers i.e. prostate (men) and breast (women) were
93% and 89% respectively. Thus, although the overall incidence of cancer has increased, the
mortality has decreased. In fact, due to longevity, more than one third of the Finnish
population will suffer cancer at some point in their lives (Pukkala, Sankila, Rautalahti 2011).
However, the prognoses of certain cancers such as malignant gliomas, lung cancer and
pancreatic cancer still remain poor, while prognoses of others such as colon carcinoma rely
heavily on the time of diagnosis: locally confined disease can be resected but advanced and
metastatic disease is virtually incurable (Brenner, Kloor, Pox 2014). Oncolytic virotherapy
(OV) presents a new and promising approach to tackle these devastating diseases. Due to
genetic manipulation or their natural tendencies oncolytic viruses replicate preferentially in
neoplastic tissue or in the tumor vasculature, and after initial cell lysis, potent antitumor
immune responses can be evoked. However, both innate and adaptive immune responses
can also be targeted against the virus being thus detrimental to the therapy (Russell, Peng,
Bell 2012). Recently, Talimogene laherparepvec (T-VEC) was employed in the treatment of
advanced melanomas, and passed as a first oncolytic virus in history a phase III clinical
trial, representing a significant milestone in this rapidly developing field (Andtbacka et al.
2013; Pol et al. 2013).
VA7 is an oncolytic alphavirus based on the avirulent Semliki Forest virus (SFV) strain that
even before the present work (I-III) had displayed potential in orthotopic xenograft models
of human osteosarcoma (Ketola et al. 2008) and lung adenocarcinoma (Maatta et al. 2008),
and considerable therapeutic power against human melanoma xenografts after systemic
delivery (Vaha-Koskela et al. 2006).
This dissertation aimed to further explore the possibilities of targeting tumors occurring in
difficult body locations such as the brain using an immunocompromised glioma xenograft
model (I) and in different immunocompetent mouse glioma models with or without
concomitant immunosuppression (II). The specific roles of innate and adaptive immune
responses were then further dissected utilizing two mouse colon carcinoma clones with
opposing patterns of cell autonomous antiviral signaling and immunogenicity (III).
2
3
2 Literature review
2.1 CANCER
2.1.1 Hallmarks of Cancer
Tissue homeostasis and regeneration are strictly regulated processes, where the
proliferation of cells is tightly controlled by growth suppressive and apoptotic
environmental signals, and where possible mutations, i.e. mistakes in the genome copying
process, are rapidly repaired. Cancer is a disease with a single cell origin. During a long,
multistep Darwinian selection process, the pre-cancerous cell acquires several biological
properties via genetic and epigenetic changes. These properties, or hallmarks, enable the
survival and development of the group of dividing cells into multicellular organ called
tumor. The tumor contains not only malignant cancer cells but also several types of
microenvironmental cells comprising the tumor stroma (Weinberg 2014). According to the
six classical hallmarks, malignant cancer cells succeed in sustaining proliferative signaling,
evading growth suppressors, activating invasion and metastasis, enabling replicative
immortality, inducing angiogenesis and resisting cell death. Two additional, strongly
emerging hallmarks of cancer are reprogramming of energy metabolism and evading
immune destruction. Two characteristics that enable this development are genomic
instability that provides the genetic diversity required for selection and cancer evolution,
and inflammation that promotes several hallmark capabilities (Hanahan and Weinberg
2011). Cancer is not a communicable disease apart from two cancers affecting canine
species: devil facial tumor disease (DFTD) and canine transmissible venereal tumor (CTVT)
and one transmissible sarcoma affecting Syrian hamsters (Murchison et al. 2010).
2.1.2 Oncogenesis, tumor microenvironment and cancer stem cell theory
The primary oncogenesis is caused by exposure to mutagens (sometimes classified as
initiator and promoter mutagens) followed by a failure in the DNA-damage sensing or
repair/apoptosis mechanisms. Occasionally, hereditary factors (e.g. BRCA gene mutations
in breast cancer) may contribute to the initial failure in the DNA-damage sensing and
repair mechanisms, while in other cases this machinery may be overdriven by massive and
repetitive carcinogen exposure as is the case in smoking related lung cancer. Persistent viral
infections can also induce cancer due to the virally encoded oncogenes and/or as a result of
infection related chronic inflammation. Viruses oncogenic for humans and other vertebrates
include members of papilloma/polyoma viruses, herpes viruses, hepatitis viruses and
retroviruses, while some adenoviruses are known to be tumorigenic in hamsters (Weinberg
2014). The microenvironment of cancer cells is undergoing constant spatio/temporal
change. During malignant development, the microenvironment changes dramatically from
the primary location to the invasive front and then again from the hostile environment of
the blood circulation to the completely new microenvironments faced by distant
4
metastases. Migration to new anatomical loci demands a major adaptation from the cancer
cells. Cancer cells are thus under constant selective pressure and the process from initial
oncogenic mutations to widespread deadly metastases typically spans several years.
Indeed, in certain malignancies such as breast cancer and melanoma dormant
micrometastases are found in distant target organs already in the early phase of the disease,
and after therapeutic destruction of the primary tumor, the metastases might erupt decades
later (Burrell et al. 2013; Hanahan and Weinberg 2011; Weinberg 2014). In other cancers,
such as pancreatic ductal adenocarcinoma, metastasis seems to occur late in the disease
progress (Hanahan and Weinberg 2011). This suggests that tissue colonization, i.e. the
development of micrometastases into macrometastases, which is the last and typically fatal
phase of tumorigenesis, is often the bottleneck in the complex course of the disease. Even
the metastatic preference of different neoplasms for certain target organs seems to depend
on the susceptibility of different environments to support metastatic outgrowth. The
direction of blood flow from the primary anatomical location plays likely also a part in
cases, when, for example, liver metastases sprout from primary colon carcinomas. The
evolutionary process creates major genetic and epigenetic clonal heterogeneity not only
between the primary and metastatic tumors, but also within the primary tumor itself. In
fact, the metastasis can be a two-way process and cancer cells migrate also from the
metastases back to the primary tumor thus further contributing to the total heterogeneity of
the disease (Burrell et al. 2013; Hanahan and Weinberg 2011; Weinberg 2014).
Tumors are complex organs colonized in their microenvironment not only by neoplastic
cells with varying sets of mutations and stages of differentiation (including cancer stem
cells, CSCs), but also with several types of stromal cells. These stromal cells include local
and bone marrow (myeloid) derived stem/progenitor cells, fibroblasts, vascular pericytes
and endothelial cells and several types of inflammatory immune cells such as macrophages,
neutrophils, mast cells, cytotoxic T-lymphocytes (CTLs), regulatory T-cells (T-regs), B-cells,
myeloid derived suppressor cells (MDSC) and macrophages. The last two immune cell
classes are closely related and often M2 biased and thus tuned for tissue repair, in contrast
to the cytotoxic M1 phenotype (Hanahan and Weinberg 2011; Mantovani et al. 2009).
CSCs are a subpopulation of neoplastic cells within tumors that express typical stem cell
markers, have an enhanced capability to induce tumors in immunocompromised mice and
are inherently resistant to chemo- and radiotherapy, possibly due to their slow pace of
proliferation. There is recent data suggesting that CSCs are actually cells that have
undergone epithelial-to-mesenchymal transition (EMT), epigenetic change in gene
regulatory program induced by inflammatory cytokines and growth factors TNF, IL-6,
TGF, HGF and EGF. The EMT program has long been recognized as being crucial for
cancer cell invasion and metastasis but it is also a normal physiological program in wound
healing and embryogenesis initiated by the sensing of hypoxia (Hanahan and Weinberg
2011; Holzel, Bovier, Tuting 2013). In addition to EMT induction, TGF has also many other
important roles in tumor progression, for example by exerting immunosuppressive
functions (Pickup, Novitskiy, Moses 2013). Epithelial carcinoma cells acquire a
mesenchymal (fibroblastic) phenotype in the invasive front of the tumors by activation of
EMT gene regulatory program, process characterized by loss of adhesion molecules such as
E-cadherin. Cancer cells then invade the blood vessels via proteolytic destruction of
5
extracellular matrix (ECM) and, once in the circulation, form aggregates with platelets.
These aggregates physically clog capillary vessels anchoring the cancer cell to its new
metastatic location. Cancer cells can reach the circulation also via the lymphatic system. It is
noteworthy, that platelets have a second important function in the metastasis cascade as
they protect the cancer cells from the immune system during the circulation (Weinberg
2014). In their new metastatic location, the mesenchymal cancer cells can undergo the
reverse program, mesenchymal-to-epithelial transition, and stay dormant in this epithelial
state for an extended time. Successful tissue colonization is often followed by reactivation
of the EMT program followed by further metastatic spawning of new malignant cells. In the
mechanistic cancer invasion and metastasis model, the EMT program induced in the cancer
cells by inflammation and hypoxia in the tumor microenvironment serves a dual function,
first by enabling the cancer cell mobility, and secondly, by creating therapy resistant CSCs.
These CSCs, by their very definition, are also capable of regenerating the tumor even after
extended time (possibly via transient epithelial phase) once the therapy has ended.
Interestingly, very recently the EMT program has been implicated in inflammation-induced
phenotypic plasticity and resistance not only against the classical cancer therapies but also
targeted therapies and even immunotherapy as a result of changes in the antigenic
landscape (Holzel, Bovier, Tuting 2013).
2.1.3 Cancer immunoediting: Sum of immunosculpting and immunosuppression
Cancer immunoediting is a process in which an individual’s immune cells (especially Tlymphocytes of adaptive immunity) protect him against tumorigenesis or sculpt the
antigenic profile of the tumors by removing cells, which present immunogenic neo-antigens
but are incapable of exerting immunosuppression (DuPage et al. 2012; Matsushita et al.
2012). Immunoediting has three phases: elimination (also called cancer
immunosurveillance), equilibrium and escape. In the elimination phase, some (occasionally
all) transformed cancer cells are recognized and eliminated by the various arms of the
host’s immune system. Natural killer (NK) cells can use their NKG2D receptors to identify
tumor stress ligands induced by carcinogens, viral infections or genotoxic stress. CD8+ and
CD4+ T-cells recognize tumor associated (TAA) and specific (TSA) antigens presented
on MHCI and MHCII molecules of dendritic cells, respectively. B-cells can augment the
antitumor immune response by secreting anti-tumor antibodies (Abs) against epitopes on
the cancer cells. Other, more recently discovered, immune cells possessing antitumor
functions include natural killer T-cells (NKT cells) and T cells. In addition to direct
tumor killing, these cells can also secrete cytotoxic and immunostimulatory cytokines.
Large amount of data has accumulated during the past few years strongly suggesting that
cancer immunosurveillance does exist in mice, and most likely, also in humans. For
example, mice with varying genetic defects either in their innate- (including type I and type
II IFNs) or adaptive immune responses are more prone to developing tumors both
spontaneously and after carcinogen or oncovirus challenge (Dunn, Koebel, Schreiber 2006;
Weinberg 2014). Furthermore, only a few of the tumors chemically induced in
immunocompromised mice grow when implanted into immunocompetent hosts, whereas
all of the tumors induced in immunocompetent mice can thrive in their own
6
immunocompetent strain, demonstrating that primary carcinogenesis does not always
produce immunogenic cancer cell clones, but when it does, these clones are indeed rapidly
eliminated by the host immune system (Shankaran et al. 2001; Weinberg 2014).
If some of the tumor cells survive the elimination, equilibrium phase follows where the
tumor does not grow in size but remains seemingly dormant via equilibrium of constant
cell proliferation and immunodestruction. The third and the final phase, the escape, begins
when immunosculpting has produced non-immunogenic tumor clones as a result of
mutations/silencing in the antigen expression (if this is redundant for the tumor), antigen
processing or the antigen presenting (MHC) machinery. Escape can take place also when
tumor clones with immunosuppressive capabilities emerge or the immune system of the
host becomes exhausted, the latter marked, for example, by expression of the inhibitory
CTLA4 receptor on T-cells (Hodi 2013; Mellman, Coukos, Dranoff 2011). Occasionally the
non-immunogenic or immunosuppressive tumor cell variants might be present already
early in disease progression, and since these variants are not eliminated, they also omit the
equilibrium phase and proceed directly to escape (Dunn, Koebel, Schreiber 2006).
Tumors and their stromal cells such as macrophages, regulatory-T cells (Tregs) and
myeloid derived suppressor cells (MDSCs) can produce different soluble or membrane
bound immunosuppressive factors in the tumor microenvironment including PD-L1, TGF,
RCAS1, IDO2, M-CSF, Galectin-1, VEGF-A, adenosine, prostaglandin E2, IL-10 and MICA/MIC-B. MIC-A and MIC-B are soluble NKG2D ligands that disturb both NK and effector
T-cell functions. STAT3 transcription factor is the crossing point for numerous oncogenic
and immunosuppressive signaling pathways and thus is constitutively activated in many
tumors (Mellman, Coukos, Dranoff 2011; Rabinovich, Gabrilovich, Sotomayor 2007; Salih,
Holdenrieder, Steinle 2008; Yu, Kortylewski, Pardoll 2007). Very recently, two independent
groups reported a new immunosuppressive ligand “V-domain Ig suppressor of T-cell
activation” (VISTA), which mimics PD-L1 in function (Le Mercier et al. 2014; Lines et al.
2014). However, in contrast to PD-L1 which is expressed not only by dendritic cells but also
by cancer cells, VISTA seems not to be expressed by cancer cells but only by T-reg and
myeloid cells in the tumor microenvironment.
The immunosuppressive factors expressed by cancer cells and immunosuppressive cells of
the tumor microenvironment can act by directly suppressing T-cell and dendritic cell
activation and function, or by activating and expanding immunosuppressive myeloidderived mesenchymal stem cells, T-regs and MDSCs. VEGF is a multifunctional
immunosuppressant and it has direct effects against T-cell development. In addition, it can
cause disturbances in vascular development resulting in aberrant blood vessels, thus
hindering the carefully controlled process of effector lymphocyte extravasation (Mellman,
Coukos, Dranoff 2011). Some of the immunosuppressive factors are produced in response
to hypoxia, a very common state of tumors with abnormal vasculature and
immunosuppressive in its own right. Hypoxia induces glycolysis resulting in the
generation of lactic acid, which lowers pH in the tumor microenvironment, thus leading to
suppression of lymphocyte function. Interestingly, tumors also often use glycolysis and
lactic acid fermentation even when oxygen is available perhaps partly for the purpose of
macromolecule synthesis, a phenomenon called the Warburg-effect. This deregulation of
7
cellular energetic is one of the emerging hallmarks of cancer (Choi et al. 2013; Hanahan and
Weinberg 2011). In addition to the Darwinian selection of antigen escape variants and
immunoevasion by direct expression of immunosuppressants, tumors can also avoid attack
by the immune system via inflammation induced transient dedifferentiation, a “wound
healing” response after therapeutic injury (Holzel, Bovier, Tuting 2013). Finally, some
cancers seem to bias the adaptive immunity towards the Th2 polarity representing allergy
and neutralizing Ab response type of immunity (mediated by IL-4) to avoid the tumoricidal
Th1 response of cellular immunity and opsonizing Abs mediated by IFN- and IL-12
(Abbas, Lichtman, Pallai 2007; Protti and De Monte 2012). The functional role of recently
discovered Th17 cells in cancer remains controversial, as these cells seem to possess both
protumorigenic and antitumorigenic properties (Ye, Livergood, Peng 2013). Th17 cells are
proinflammatory and important mediators of autoimmune diseases. Both
immunosuppressive Tregs and immunostimulative Th17 cells need TGF for their
differentiation from the naïve CD4+ T-cells, but certain environmental cues such as the
inflammatory cytokine IL-6 can skew the delicate balance between inflammation and
immune tolerance by triggering the differentiation of naïve CD4+ T-cells into Th17 cells
instead of Tregs (Barbi, Pardoll, Pan 2013). Virus-infected cells can secrete IL-6 (Dienz et al.
2012) and IL-6 could be possibly ectopically expressed from oncolytic viruses to induce
antitumor immunity.
Taken together, the immunosuppressive microenvironment of cancers is very complex and
might also vary significantly between different tumors. Some immune cell classes generally
considered to be tumor destructive, such as B-cells, NK-cells, NKT-cells, T cells, have
been also reported in certain circumstances to secrete IL-10 and thus exert
immunosuppressive functions (Mellman, Coukos, Dranoff 2011; Salih, Holdenrieder,
Steinle 2008).
2.1.4 Cancer in Finland and modern cancer therapies
As a result of advances in contemporary medicine, cancer is no longer an invariably lethal
disease. In fact, because of longevity, approximately one third of Finnish population will
suffer cancer at some point during their lifetime. Approximately one third of all cancers are
estimated to result from the use of tobacco products, which in addition to causing lung
cancers are involved in the etiology of carcinomas of mouth, pharynx, kidney, pancreas,
esophagus, cervix and uterus, and might contribute also to breast cancer (Pukkala, Sankila,
Rautalahti 2011). The prognoses of certain malignancies such as liver, gall bladder, lung
and pancreatic carcinomas and glioblastoma multiforme (GBM) remain still poor (Pol et al.
2013; Pukkala, Sankila, Rautalahti 2011).
GBM is the most common and most aggressive primary brain malignancy with a median
survival of 15 months and 5-year survival of less than 10% (Pol et al. 2013; Pukkala, Sankila,
Rautalahti 2011). The DNA-alkylating agent, temozolomide, has displayed some efficacy in
combination with standard surgery and chemo/radio therapy. However, patients cannot be
cured and they rarely live beyond two years post diagnosis (Maatta et al. 2009; Zemp et al.
8
2010). In other malignancies, such as colorectal carcinoma, the prognosis depends heavily
on the time of diagnosis: locally confined carcinoma in situ can be treated by resection,
whereas advanced invasive and metastatic disease is considered incurable. Oncolytic
virotherapy represents a novel possibility to treat these devastating diseases, and while no
virus medicine has been approved for western markets, T-VEC as the first one has recently
completed phase III trial for advanced melanoma (see chapter “2.4 Oncolytic Virotherapy”)
(Pol et al. 2013; Russell, Peng, Bell 2012). Gene therapy and immunotherapy of cancer are
two additional modern technologies being evaluated in clinical trials and these approaches
frequently overlap with oncolytic virotherapy (see chapter “2.4.2 Oncolytic virotherapy
overlaps with gene therapy and immunotherapy of cancer – emerging concept of oncolytic
immunovirotherapy”). After a long and complex process, the first gene therapeutic drug,
Glybera (alipogene tiparvovec), was recently approved for clinical use in Europe. Glybera is
an adeno-associated viral (AAV) vector engineered to express lipoprotein lipase in muscle
to be used in the treatment of the very rare condition of lipoprotein lipase deficiency (YlaHerttuala 2012).
Classical cancer therapies comprise treatments such as surgery, radiotherapy,
chemotherapeutics and anti-hormonal treatments (Pukkala, Sankila, Rautalahti 2011). Some
targeted cancer therapies are already widely in use e.g. small molecule inhibitors (SMIs)
and monoclonal antibodies against targets such as Her2/neu, epidermal growth factor
receptor (EGFR), vascular endothelial growth factor (VEGF), CD20, CD52 and CD33. The
antibodies antagonize oncogenic pathways and may also act via opsonizing the tumor cells
thus causing cellular death or removal by antibody-dependent cellular cytotoxity (ADCC).
According to preclinical data these antibodies might also stimulate adaptive immune
responses (Mellman, Coukos, Dranoff 2011). Furthermore, the anti-CTLA-4 antibody,
ipilimumab, was recently approved for the treatment of metastatic melanoma
demonstrating a marked and occasionally sustaining survival benefit for patients in the late
stages of the disease (Hodi 2013; Mellman, Coukos, Dranoff 2011). CTLA-4 is expressed on
the T-cell surface 2-3 days after activation and its natural role is to down-regulate T-cell
activation thus preventing autoimmune damage resulting from prolonged immune attack.
CTLA-4 competes with the activating CD28 receptor on Tc- and Th-cells for binding with
the B7 costimulatory ligand on antigen presenting cells (APCs) leading to T-cell
deactivation (Mak and Saunders 2011). Intriguingly, CTLA-4 ligation is also important for
Treg-mediated immunosuppression through a still poorly understood mechanism, and
ipilimumab is also hypothesized to block this interaction. A second class of antibodies on
the verge of clinical approval target the inhibitory PD-1 - PD-L1(L2) interaction between
activated T-cell and APCs (Mellman, Coukos, Dranoff 2011). PD-1 is functionally related to
CTLA-4 as they both mediate immune exhaustion after prolonged activation, but PD-L1
and PD-L2 can often be found not only on APC but also tumor membranes adding a layer
of specificity for cancer immunotherapy as compared to CTLA-4. Indeed, antibodies
against PD-1 have demonstrated efficacy in clinical trials against a variety of cancers with
less autoimmune side effects than ipilimumab (Mellman, Coukos, Dranoff 2011).
Interestingly, PD-L1 can also induce Treg differentiation together with TGF (Francisco,
Sage, Sharpe 2010).
9
A third class of clinically promising immunomodulatory antibodies against cancer are the
agonistic CD40 monoclonal antibodies, which stimulate APCs by binding to their CD40
receptors thus activating CTLs. In the absence of T-cells, these antibodies can activate also
cytotoxic myeloid cells (Vonderheide and Glennie 2013). Other cancer immunotherapies in
clinical use include IFN- and IL-2 treatments for melanoma and renal cell carcinomas, the
two cancer types considered to be the most immunogenic, and intravesical injection of
bacillus Calmette-Gue´rin used for the treatment of superficial bladder cancer.
Interestingly, those melanoma patients who respond best to the IFN therapy are often
naturally prone to autoimmune diseases suggesting that the immunity of these patients
might be Th1 biased. One type of immunotherapy, though originally not considered to be
one, is the allogeneic bone marrow transplantation and infusion of donor lymphocytes,
which are used in the treatment of leukemias and lymphomas where “graft-versusleukemia effect” has been found to have major contribution to the therapy. Finally, the first
active cancer immunotherapy Provenge (sipuleucel-T) involving ex vivo activated
peripheral blood mononuclear cells was approved in 2010 by Food and Drug
Administration for the treatment of advanced prostate cancer. Provenge’s mechanism of
action remains incompletely understood, and its therapeutic effects are modest, but the
approval was granted since the current therapeutic options for advanced prostate cancer
are very limited (Mellman, Coukos, Dranoff 2011).
In addition to monoclonal antibodies, the emergence of potent SMIs against a variety of
natural or mutated kinases, such as Imatinib (Gleevec) for inhibition of BCR-ABL1 mutant
fusion kinase in chronic myeloid leukemia (CML), has advanced significantly the treatment
of cancer. Unlike chemotherapeutics and radiation which target all dividing cells, these
kinase inhibitors are more cancer specific, thus increasing their safety. Nevertheless, as with
traditional therapies, cancer cell clones resistant to the inhibitors frequently develop as a
result of acquired mutations to the targeted kinases or the cancer cell’s ability to activate
alternative signaling routes (Zhang, Yang, Gray 2009).
2.2 INTERFERONS IN CELL AUTONOMOUS, INNATE AND ADAPTIVE
IMMUNITY
In order to enter a cell, hijack its macromolecular synthesis machinery and to launch a
productive infection, viruses first need to breach natural protective barriers such as the
skin, mucus, saliva, tears and gastric acid. Once inside the cell, the first branch of immunity
that the virus will encounter is cell autonomous immunity, most importantly the interferon
(IFN) system (Figure 1).
10
Figure 1. Type I IFN response in vertebrate immunity. Interferons (INFs) are evolutionally
conserved cytokines amongst vertebrates, which have broad antimicrobial, immunostimulatory
and growth-suppressive functions. Type I IFNs protect against viruses, bacteria and protozoa
and act as a link between innate-, adaptive- and cell-autonomous immunity (Abbas, Lichtman,
Pallai 2007; MacMicking 2012).
Interferon defense system is evolutionally conserved amongst all vertebrates. IFNs have
broad antimicrobial and growth-suppressive functions and can stimulate both the innate
and adaptive branches of immunity (MacMicking 2012). IFNs are classified according to
their receptors into type I (IFN-, IFN-, IFN-, IFN-, IFN-, IFN- and 13 different IFN-
subtypes in humans and 14 in mice), type II (IFN-) and type III (4 different IFN-
subtypes) interferons. While type I and type III IFNs have separate receptors, i.e. IFNAR
and IFNLR, respectively, they share the same downstream signaling pathway involving
Signal Transducer and Activator of Transcription STAT1/STAT2 heterodimer
phosphorylation by Jak1/Tyk2 kinases resulting in formation of the STAT1/STAT2/IRF9
transcription complex (Honda and Taniguchi 2006; MacMicking 2012; van Pesch et al.
2004). All IFN classes possess documented antiviral activity but to this end, the best
characterized are the type I interferons (mainly IFN-). The only type II IFN is IFN- and
it signals via IFNGR and STAT1/STAT2 heterodimer. IFN- is important mediator of
adaptive immune response, but it is also involved in virus clearance from the central
nervous system (CNS) at least during Sindbis and Vaccinia virus infections (Binder and
Griffin 2001; Kundig, Hengartner, Zinkernagel 1993; McFadden et al. 2009). Type I IFN
seems to be more important than type II IFN in the restriction of Herpes simplex virus
(HSV)-I CNS replication (Leib et al. 1999). While type I IFNs can be responded to and
secreted by most somatic cells, both the production and response to antiviral type III IFNs
seem to be limited to epithelial cells and tissues, especially the lungs. Nevertheless, the
similar signal transduction cascades downstream of type I and type III IFN receptors results
in induction of the same set of antiviral genes. The limited response to type III IFNs is
related to the fact that high levels of the type III IFN receptor are found only in epithelial
11
cells, and the potential mechanisms of epithelial cell and tissue type specific induction of
this IFN class are unknown (McFadden et al. 2009).
The engagement of pattern recognition receptor (PRR) with pathogen associated molecular
patterns (PAMPs) induces the expression and auto/paracrine signaling of type I IFNs. The
PRRs which are most important in viral infections can be roughly divided into two classes
according to the PAMP type that they detect and the cell types in which they are found:
Toll like receptors (TLRs) which detect microbial structural components and nucleic acids,
are mostly found in plasma membranes and endosomes of leukocytes whereas the cytosolic
pattern recognition receptors (e.g. RIG-I like receptors RLRs) detect foreign nucleic acids
and are found in most cell types. The expression patterns of some TLRs vary between
species, and whereas TLR9 detecting CpG DNA is expressed in B-cells and several types of
phagocytes in mice, it is expressed only in plasmacytoid dendritic cells (pDCs) and B-cells
in humans (Hochrein and Wagner 2004; Thompson et al. 2011). In addition to immune cells,
TLRs have been detected also in cancer cells (Sato et al. 2009). Furthermore, non-immune
cells can detect cytosolic DNA via receptors such as IFI16 or RNA polymerase III, the latter
curiously recognizing foreign DNA indirectly by synthesizing RIG-I substrates out of it.
However, signaling via these receptors remains incompletely understood and there might
well be cell type specific differences (Broz and Monack 2013).
RLRs belong to the H box helicase (Asp–Glu–x–Asp/His box) family (Goubau, Deddouche,
Reis e Sousa 2013). They are crucial for the detection of double-strand RNA (dsRNA) and
single-strand RNA (ssRNA) genomes as well as replication intermediates of RNA-viruses.
RLR family has three central members: MDA-5 (melanoma differentiation factor 5), RIG-I
(retinoic acid-inducible gene I) and LGP-2 (laboratory of genetics and physiology 2). RLRs
signal through MAVS (mitochondrial antiviral signaling, aka IPS-1, CARDIF or VISA)
activating IRF3/IRF7 (in homo- or heterodimers), IRF1 and NF-B leading to the expression
of type I IFNs and pro-inflammatory cytokines, including chemokines (Broz and Monack
2013; Honda and Taniguchi 2006). In addition to detecting viral RNA, RLRs themselves are
also induced by type I IFNs thus participating in a positive-feedback loop (Goubau,
Deddouche, Reis e Sousa 2013). Other PRR classes found in the animal kingdom, which are
important for innate immunity against pathogens such as viruses include C-type lectin
receptors (CLRs) and scavenger receptors found in phagocytes and some epithelial cells
(Jiang et al. 2006), and several soluble PRRs such as mannose binding lectins (Dommett,
Klein, Turner 2006). Furthermore, in addition to the classic RLRs new members of cytosolic
PRRs are constantly discovered (Broz and Monack 2013). Complement is not part of the
PRR system per se but not only does it recognize and destroy lectin or antibody opsonized –
microbes, it can also target foreign non-self structures unprotected by complement
regulatory proteins (Ricklin et al. 2010).
Type I IFNs are produced by and act on almost all nucleated cell types. They can inhibit
every stage of the virus lifecycle by inducing hundreds of interferon stimulated genes
(ISGs) (Goubau, Deddouche, Reis e Sousa 2013; Grabowska and Riemer 2012; Liu et al.
2012; Sadler and Williams 2008; Thomas et al. 2011). Type I interferons signal through
common IFNA-receptor (IFNAR) coupled with JAK/STAT, p38 MAPK and PI3K pathways.
However, type I IFNs differ to some extent in the physiological responses that they evoke
12
such as potency against different viruses through poorly understood mechanisms. All type
I IFNs cause identical conformational changes in their heterodimeric IFNAR1/2 receptor.
This leads to receptor autophosphorylation and activation of Tyk2 and Jak1, which are
constitutively attached to IFNAR1 and IFNAR2, respectively, and which further activate
STAT1 and STAT2 by phosphorylation. The phosphorylated STAT1 and STAT2
heterodimerise and become associated with IRF9 in the cytoplasm or nucleus. The resulting
complex Interferon-stimulated gene factor 3 (ISGF3) acts as a transcription factor binding to
interferon stimulated response elements (ISRE) found on promoters of many ISGs. ISGF3 is
the only factor binding to ISRE elements. Not only the classical STAT1/STAT2 heterodimer
but also the STAT5/CrkL dimer can be phosphorylated and then translocated to the nucleus
upon INFAR signaling, inducing the expression of certain ISGs. However, in contrast to
ISGF3, STAT5/CrkL complex binds to interferon gamma activated site (GAS) rather than
ISRE, thus resembling IFN- activated STAT1-homodimer. ISRE and GAS signaling are not
nevertheless biologically fully distinct, and many ISGs harbor both elements. Furthermore,
type I IFNs can also induce STAT1 homo-dimerization and signaling via GAS (HervasStubbs et al. 2011; MacMicking 2012; Platanias 2005). Finally, STAT1 has been reported to
possess some direct antiviral functions against SARS-CoV, independent of type I, II or III
IFN signaling (Frieman et al. 2010).
The anti-proliferative effects and ISG induction levels seem to correlate positively with
receptor binding affinities of different type I IFNs (Platanias 2005; Thomas et al. 2011). In
addition to the differences in the quantity also the quality of ISG induction may
occasionally vary between different type I IFNs, and IFN- but not IFN- was reported to
induce -R1/CXCL11 (Rani et al. 1996). Importantly, as a result of complex and poorly
understood cell physiological interactions, ISGs can also have different roles in different
cells and responses to different viruses can be dramatically different also in a given cell
type (McFadden et al. 2009). IFN- can be divided into early (IFN-4 and IFN-) and late
interferons (IFN-non4, i.e. rest of the IFN- classes) according to whether in addition to
PRR activation, priming by auto/paracrinic IFNAR signaling (and resulting IRF7
expression) is a prerequisite for their expression. IFN- is the master-regulator of type I IFN
response requiring only PRR activation and IRF3-homodimerization for its induction. IFN is transcribed approximately 2h p.i., which is even earlier than the induction of the other
early type I IFN, IFN-4, the transcription of which is detectable around 8h p.i.
(Lienenklaus et al. 2008). Paracrine signaling of type I IFN primes neighboring cells for an
imminent virus attack by induction of effector ISGs and IRF7. This process is fast, and the
first signs of phosphorylated STAT1 can be found already 5 minutes after IFN-
stimulation (Dempoya et al. 2012). In the case of PAMP detection in the primed cell, the
induced IRF7 is activated by phosphorylation and homo-dimerization, or by heterodimerization with IRF3. Phosphorylated IRF7 dimer is then translocated to the nucleus
where it induces the rapid transcription not only of early type I IFNs but also of all the late
type I IFNs hence resulting in positive-feedback signaling cascade (Figure 2) (Honda and
Taniguchi 2006). Interestingly, type I IFN independent STAT1 phosphorylation has also
been recently reported (Dempoya et al. 2012). However, STAT1 phosphorylation was not
detectable before 8 h after poly I:C induction of IFNAR-deficient U5A cells, a time point
which is 5 h later than that found in IFNAR-signaling competent A549 cells.
13
Figure 2. Paracrinic signaling of type I IFNs (modified with permission from Zemp et al. 2010).
1. Nucleic acids of invading microbes such viruses are detected by pathogen recognition
receptors (PRRs), such as cytosolic helicases Rig-I and MDA5 in case of SFV. 2. IRF3 is
phosphorylated in a process involving Mitochondrial antiviral-signaling protein (MAVS, not
shown) and TANK-binding kinase 1 (TBK1). 3. Phosphorylated IRF3 goes through
homodimerization (and heterodimerization with IRF7, not shown) and nuclear translocation
leading to induction and low level secretion of chemokines (not shown) and first early type I
IFN, IFN-. In addition, some interferon stimulated genes (ISGs) are induced already at this
stage, including Zinc-Finger Antiviral Protein (ZAP) and Viperin, both shown to have potent
antiviral functions against SFV. 4. IFN- is secreted and causes autocrine and paracrine
signaling via common heterodimeric type I IFN receptor IFNAR1/IFNAR2 the chains of which are
constitutively coupled to TYK2 and JAK1 kinases, respectively (not shown). 5. The kinases
phosphorylate STAT1 and STAT2 causing their heterodimerization and attachment with IRF9,
forming the ISGF3 complex. 6. ISGF3 translocates to nucleus and induces transcription of
multiple (potentially hundreds) ISGs including antiviral effectors such as OAS, PKR, Viperin,
Zap, Ifit1/3 or PRRs such as Rig-I. Some of these ISGs such as Mx1 and ISG20 (not shown) are
“true-ISGs” that are not induced by activated IRF3 (see point 3.). 7. In addition to antiviral
ISGs and PRRs, IRF7 is induced, priming the cell to react to imminent microbial attack. 8.
Nucleic acids of invading microbes are detected by upregulated PRRs in the primed cell 9.
leading to IRF7 activation by phosphorylation and IRF7 homodimer formation, which 10.
translocates to the nucleus to drive the expression of type I IFNs such as IFN- and several
classes of IFN- (Honda and Taniguchi 2006; Zemp et al. 2010).
The binding of type I IFN to its receptor initiates a complex and incompletely understood
signaling process which can occasionally be IFN type and cell type-specific. In addition to
STAT1/STAT2 it can involve also several other STATs (STAT 3 to 6) in homo- and
heterodimeric complexes. STAT4 and STAT6 mediated signaling seems to be cell type
specific taking place primarily in endothelial and lymphoid cells. Furthermore, in addition
14
to prototypic JAK/STAT signaling, type I IFN signal can be mediated via PI3K and p38
MAPK pathways. Type I IFN induced PI3K signaling can regulate also translation in
addition to transcription via mammalian target of rapamycin (mTOR). The cross-talk
between these different pathways (especially STAT and p38) downstream of IFNA-receptor
appears to be crucial for full scale cell autonomous antiviral response (Platanias 2005).
Interestingly, despite the shared receptor IFN- and IFN- induced ISGs do not completely
overlap even in one given cell line and 20 ISGs, including PKR and Hif-1, were reported to
be exclusively or preferentially induced by IFN- (Der et al. 1998). Recently, more light was
shed into the mechanisms of type I IFN specificities when IFN- but not IFN- was found
to be capable of signaling via single IFNAR1 receptor chain inducing separate set of genes
compared to IFN- signaling via full IFNAR1/2 receptor (Kaur and Platanias 2013).
Furthermore, while all nucleated cell types are thought to express IFNAR, the magnitude of
ISG induction can vary markedly between tissues (Briolat et al. 2014).
In some, presumably rare, cases the presence of type I IFN can actually increase virus
replication at least in vitro when IFN signaling, while inherently antiviral, happens to
stimulate the expression of the virus receptor. This has been reported to take place in
porcine reproductive and respiratory syndrome virus (pRRSV) infection of monocytes
(Duan, Nauwynck, Pensaert 1997) and porcine circovirus type 2 (pCV2) infection of porcine
cell lines (Meerts, Misinzo, Nauwynck 2005).
In addition to the different IFNs, TNF which is produced by leukocytes after TLR
engagement also possesses the ability to induce the pro-inflammatory transcription factor
NF-B promoting cellular antiviral state. Both TNF and IFNs also mediate inflammation and
prime antiviral adaptive immune responses. Despite the clear direct antiviral effects of TNF
reported already 1980s, it has been nevertheless much less studied in this context than the
IFNs. The third class of cytokines that has important antiviral functions is the interleukins,
which however, while also having some direct antiviral effects, mostly stimulate antiviral
adaptive immunity (McFadden et al. 2009). Finally, IL-6 can induce ISG gene expression via
ISRE elements similarly to type I IFNs, but it uses STAT3 instead of ISGF3 (Hubackova et
al. 2012). IL-6 is induced in many normal tissues as a result of PRR activation or an
inflammatory cytokine signal (including IFNs). This induction is orchestrated by NF-B
together with some other transcription factors (Hong, Angelo, Kurzrock 2007). The
transcription of IFN- and several other cytokines is controlled by the enhanceosomecomplex comprised of NF-B, phosphorylated IRF3 (and IRF7 especially in type I IFN
primed cells), CBP (or p300) and ATF2-JUN (Honda and Taniguchi 2006).
The direct effects of type I IFNs on tumors and cellular immunity are remarkably versatile
and they can include direct (immunogenic) oncolysis, inhibition of angiogenesis, MHC-I
induction on cancer cells and stimulation of NK, DC and T-cell responses (Hervas-Stubbs et
al. 2011). Nevertheless, the effects of type I IFNs on adaptive immunity are both time and
concentration dependent. Indeed, type I IFNs have at least seemingly immunosuppressive
functions in deletion of old memory T-cells and premature activation of naïve,
unproliferated T-cells that have not yet met their cognate, MHC-presented antigen (Welsh
et al. 2012). Furthermore, high type I IFN concentrations can impair IL-12 production by
DCs (Hervas-Stubbs et al. 2011). The existence of multiple different type I IFNs all signaling
15
via a single receptor probably highlights the evolutionary pressure to ensure that the
crucial functions of these cytokines are maintained in antiviral innate and adaptive
immunity. While type I IFNs seem to exhibit a certain level of overlap and redundancy,
many of their specific roles probably still remain undiscovered. Much of the diversity might
lie in the tissue-specific induction rather than in cell type-specific response. Indeed,
laboratory of Paul Hertzog demonstrated recently that IFN-, a previously uncharacterized
type I IFN, has a unique role in the protection of the female genital tract, and in striking
contrast to the other known type I IFNs, it is not induced by PRR activation but instead it is
constitutively produced by epithelial cells (Fung et al. 2013). Probably for historical reasons
the majority of the studies of type I IFN immunostimulatory functions have been conducted
using IFN- subtypes, which are also the ones mainly employed for interferon therapy of
different neoplasms (mostly melanoma and hematological malignancies) (Hervas-Stubbs et
al. 2011). Hence, it remains poorly defined whether the different type I IFN classes differ in
their immunoregulatory and anti-cancer efficacy. However, as type I IFN anti-proliferative
effects on cells seems to depend on the IFNAR binding affinity (Thomas et al. 2011) and
IFN- has been shown to possess higher receptor affinity and ISG induction power than
IFN--classes (van Boxel-Dezaire, Rani, Stark 2006), further comparative studies in the
context of cancer immunotherapy are highly warranted.
2.3 ALPHAVIRUSES
2.3.1 Taxonomy, natural life cycle and pathogenesis of Alphaviruses and Semliki Forest
virus
The genus Alphavirus belongs to family Togaviridae and is comprised of more than 40 small
(65-70nm), enveloped, icosahedrally shaped positive-sense ssRNA viruses (Jose, Snyder,
Kuhn 2009). Alphaviruses are arthropod-borne viruses (arboviruses) that use birds and
mammals as amplifying hosts and are typically spread by mosquito vectors (Murray,
Rosenthal, Pfaller 2013). In addition, aquatic salmonid alphavirus (SAV) species exist with
an unknown transmission vector, although some louse species have been proposed as the
vector (Metz et al. 2011). Some alphaviruses are important pathogens of humans and
domestic animals such as Old World Chikungunya virus (CHIKV) and Sindbis virus
(SINV) and New World viruses Venezuelan- (VEEV), Western- (WEEV), and Eastern
equine encephalitic virus (EEEV). The Old World viruses typically cause symptoms such as
fever, rash and arthritis, which especially in CHIKV infection can be coupled with long
lasting and disabilitating arthralgia. The New World alphaviruses EEEV, VEEV and WEEV
can cause deadly encephalitis in horses and occasionally also in man while no vaccines are
widely clinically available (Chattopadhyay et al. 2013; Jose, Snyder, Kuhn 2009). Live
attenuated, and formalin inactivated vaccines against VEEV exist for use in horses, U.S.
military workers and laboratory personnel at risk of infection, but the vaccines have limited
safety and immunogenicity in humans (Dupuy et al. 2011; Goldberg March 6, 2014; Paessler
and Weaver 2009).
16
Interestingly, even though Semliki Forest virus (SFV) belongs to Old World Alphaviruses
which are associated rather with arthritis than neurologic type of pathologies, even the
avirulent strain A7(74) and derived vector VA7 are neurotropic in mice. SFV is not
considered to be a human pathogen, but some neuropassaged strains such as L10 cause a
rapid and fatal CNS infection in mice and other rodents (Bradish, Allner, Maber 1971).
SFV4 is a less virulent SFV strain, which replicates efficiently in mouse brain if
administered by an intracranial (ic) injection but only occasionally when injected
intraperitoneally. This probably reflects the reduced neurotropism of SFV4 as compared for
example to the avirulent A7(74) (Fragkoudis et al. 2007). In addition, the avirulent SFV
strains cause lethal encephalitis in neonatal mice younger than 12 days while mice older
than 14 days are generally protected against the disease, reflecting a poorly understood
phenomenon of age-dependent neurovirulence. While decreased membrane biogenesis in
mature neurons has been postulated to underlie the reduced infectivity of A7(74)
(Fazakerley 2002), it would be important to study in this context the role of the
immunological maturation from Th2 to Th1 biased immunity occurring around day 14
postnatal (Lovett-Racke et al. 2000). A7(74) does not replicate in the CNS cells of SCID mice
that are suppressed in their adaptive immunity (Amor et al. 1996). However, IFN- the
prototypical Th1 cytokine (along with IL-12), is secreted not only by T-cells but also by NKcells, a class of immune cells present also in SCID mice. CNS infection of avirulent SFV
strains is known to cause not only a disruption of the blood brain barrier (BBB) but also a
transient demyelinating disease resembling multiple sclerosis. The demyelination might
result from direct CTL-mediated destruction of infected oligodendrocytes and/or a CTL
response triggered by molecular mimicry between SFV envelope proteins with myelin
oligodendrocyte glycoprotein (MOG) (Fazakerley 2002; Mokhtarian et al. 1999). Athymic
nude mice are protected against A7(74) induced demyelination, despite the reported
subclinical CNS persistence of the virus in these mice (Amor et al. 1996).
In adult mice, intraperitoneal injection of both the avirulent A7(74) and virulent SFV4
results in high-titer plasma viremia of 107-108 plaque forming units per milliliter (PFU/mL)
peaking at day 1. The neurotropic viruses are seeded from blood into perivascular foci in
the brains and spinal cord and are undetectable in the blood beyond day 4 as a result of
viral clearance by IgM antibodies (Fazakerley 2002; Tuittila et al. 2000). A7(74) infection
results in the appearance of perivascular virus antigen deposits in neurons and
oligodendrocytes (but not astrocytes) in which they can be detected until the virus is
immunologically cleared by d10 post infection (p.i.). However, SFV4 strain continues to
replicate causing pancerebral infection and death of the mice within 4-6 days p.i. SFV4
brain titers can reach up to 108 PFU/mL before the death of the mouse, being approximately
four log units higher than the respective peak titers of 104 PFU/mL with the avirulent
A7(74). The major virulence determinants between the virulent SFV4 and the avirulent
A7(74) have been traced to the nonstructural protein 3 (nsP3) gene (Fazakerley et al. 1993;
Tuittila et al. 2000). Human SFV infections are common in Africa, where in certain areas up
to 25% of the population (67 of the 267 studied cases) has been reported to be seropositive
(Adekolu-John and Fagbami 1983). An obvious diagnostic limitation in the neutralization
and hemagglutination inhibition assays used is the possibility of false positives as a
resulting from cross-reactive antibodies between two antigenically related viruses. Indeed,
17
while 52% (139/267) of the blood samples tested positive for anti-alphavirus antibodies in
Kainji Lake Basin, Nigeria, only in one sample the antibodies were monotypic for SFV
(Adekolu-John and Fagbami 1983). Similarly in another study, while 33% (41/123) of the
tested adult population in Angola possessed neutralizing antibodies against SFV, in a
separate test only 11% (7/63) of the samples containing anti-SFV antibodies had no
antibodies against Sindbis or Chikungunya (Kokernot et al. 1965). Relatively mild
symptoms such as headache, fever, myalgia and arthralgia have been associated with SFV
infection (Mathiot et al. 1990). Interestingly, direct intracranial (i.c.) injection of avirulent
A7(74) and neurodetargeted SFV4-miRT124 results in the brains almost exclusive infection
of oligodendrocytes in white matter tract of corpus callosum (Fazakerley et al. 2006;
Ylosmaki et al. 2013). The A7(74) oligodendrocyte infection can be prevented by
intraperitoneal (i.p.) injection of the pattern recognition receptor (PRR) agonist,
polyinosinic:polycytidylic acid (poly I:C) 3h pre-infection (Fazakerley et al. 2006).
The only Alphavirus found in Finland is Sindbis virus, which is spread by late summer
ornitophilic Culex and Culiseta mosquito vectors. The virus is the etiologic agent of Pogosta
disease causing rash and occasionally arthritis. Pogosta disease emerges typically in seven
year intervals for incompletely understood reasons, but this seasonality may be related to
population cycling of the postulated amplifying host; grouse (Sane et al. 2010).
2.3.2 Alphavirus replication cycle
SFV enters the cells via receptor-mediated endocytosis in clathrin-coated vesicles, but the
main receptor remains unknown (Figure 3). Interestingly, cell culture passaged SFV and
Sindbis use glycosaminoglycan heparan sulfate (HS) as an attachment receptor, and virus
cell culture adaptation to HS binding has been proposed to correlate with the level of
neurovirulence (Ryman and Klimstra 2008; Smit et al. 2002). Sindbis virus produced in
mosquito cells but not in mammalian cells, was found to infect dendritic cells using C-type
lectins as attachment receptors (Klimstra et al. 2003). It is worth speculating whether this
curious phenomenon is an evolutionary trait, as infection of skin resident DCs after a
mosquito bite is the first step in the natural life cycle of many alphaviruses (Klimstra et al.
2003) and DC tropism probably becomes redundant in the following secondary infection
cycles inside the host. Dendritic cell tropism has been reported also for VEE and this has
been exploited in cancer immunotherapy (MacDonald and Johnston 2000; Moran et al.
2007). SFV glycoproteins have also been reported to bind these receptors, and this has
obvious immunotherapeutic implications for SFV based oncolytic vectors (see chapter
“2.4.2 Oncolytic virotherapy overlaps with gene therapy and immunotherapy of cancer –
emerging concept of oncolytic immunovirotherapy”) (Froelich et al. 2011). Nevertheless,
Schulz and coworkers reported that SFV is incapable of directly infecting DCs but crosspriming is nevertheless induced when DCs phagocytose the infected cells (Schulz et al.
2005). Though speculative, the discrepancy results concerning the infectivity of DCs with
SFV might relate to the differences between cell-culture-adapted laboratory virus strains
propagated and maintained in different producer cell lines.
18
SFV E1 glycoprotein mediates viral envelope fusion with the endosomal membrane as a
result of acidic pH induced conformational change in the envelope glycoproteins, releasing
the viral nucleocapsid into the cytosol (Fazakerley 2002; Kielian, Marsh, Helenius 1986;
Wahlberg and Garoff 1992). The RNA genome is then released from the capsid via an
incompletely understood molecular mechanism that involves attachment to ribosomes
(Singh and Helenius 1992; Singh et al. 1997). Since the SFV genomic RNA has a positive
polarity, it can immediately function as messenger RNA, initiating the synthesis of viral
nonstructural proteins. Soon after infection, the cell becomes non-permissive to other SFV
particles by poorly understood mechanism termed superinfection exclusion. The
superinfection exclusion seems to take place at multiple levels of cell surface binding,
endosomal penetration and capsid uncoating (Singh et al. 1997). The well characterized
temporal processing of the replicase complex by nsP2 protease and the resulting block in
the synthesis of negative RNA strands have also been proposed to be responsible for this
obscure, but presumably an evolutionary important phenomenon (Jose, Snyder, Kuhn
2009). It is unclear how virus-specific the superinfection exclusion is, and intriguingly
recently SFV infection of DBT mouse glioma cells has been reported to inhibit Vaccinia
virus infection (Vaha-Koskela et al. 2013).
Figure 3. Alphavirus replication cycle. (Helenius et al. 1980; Huotari and Helenius 2011; Jose,
Snyder, Kuhn 2009; Schwartz and Albert 2010; Singh and Helenius 1992; Spuul et al. 2011)
19
2.3.3 Alphaviruses and adaptive immunity
Type I IFNs are crucial for early restriction of alphavirus replication in extraneural tissues
(Fragkoudis et al. 2007) and T-cell derived IFN- contributes to non-cytolytic clearance of
attenuated alphaviruses from a neuronal subpopulation (Binder and Griffin 2001; Keogh et
al. 2003). After the immediate innate responses, a rapid adaptive IgM response develops
followed by class-switching to the IgG humoral response which is primarily responsible for
clearing of the virus infection from all parts of the CNS (Binder and Griffin 2001; Fazakerley
2002). Thus, the importance of innate and adaptive humoral immune responses in
protection and clearance of alphavirus infections is well established. Following infection
with the avirulent SFV strain A7(74), SCID mice lacking T- and B-cells develop a chronic
peripheral and CNS viremia, leading to neurological symptoms and death even though the
virus does not seem to replicate in the CNS cells (Amor et al. 1996). SCID mice infected with
A7(74) derived oncolytic virus VA7 also developed neurological symptoms after successful
treatment of human melanoma xenografts (Vaha-Koskela et al. 2006). Peripheral infection
of attenuated alphaviruses is eliminated in nude mice; this is ascribed to the gradual
development of IgM antibodies. However, the A7(74) infection was reported to result in
subclinical but persistent CNS infection in nude mice (Amor et al. 1996). Interestingly, and
supporting the role of protective IgM response in nude mice, transfer of immune-sera from
nude mice disrupts the persistent blood viremia in SCID mice, while E2 specific nonneutralizing monoclonal antibody (mAb) engraftment or IgG rich immune-sera from fully
immunocompetent mice clears the viral infection both from the periphery and the CNS
(Amor et al. 1996; Fazakerley 2002). This model was recently refined by the laboratory of
Diane Griffin, who demonstrated using Sindbis as a model virus that infection of
immunocompetent mice resulted in T- and B-cell CNS infiltration by day 3 p.i. and,
interestingly, all infectious Sindbis virus was cleared from the CNS by the early IgM
response (Metcalf and Griffin 2011). However, a low level of viral RNA persisted for
months in the CNS together with constitutive IgG production which probably kept at bay
any recurrence of renewed viral infection. Intriguingly, exogenic antibodies can even rescue
mice deficient in type I or type II IFN responses from Sindbis infection (Byrnes, Durbin,
Griffin 2000) and antibodies have been reported to clear Sindbis infection even in neuronal
cultures (Ubol et al. 1995). Apart from the well known pivotal role Th-cells play in B-cell
class switching and thus also in production of IgG antibodies against alphaviruses and the
role T-cells have as an IFN- source, the innate and adaptive cellular immune responses
seem to be secondary to humoral immunity in alphavirus clearance (Griffin et al. 1997).
2.3.4 Alphaviruses and innate- and cell autonomous immunity
The exact mechanisms of cell autonomous anti-viral signaling in alphavirus infection
remain incompletely understood. However, type I IFN signaling is crucial for the host
survival and IFNAR-knockout mice succumb rapidly to peripheral SFV infection
(Fragkoudis et al. 2007; Muller et al. 1994). On the other hand, type II IFN plays an
important role in non-cytolytic CNS clearance at least in Sindbis virus infection (Binder and
Griffin 2001; Lee, Schultz, Griffin 2013). Nevertheless, ependymal cells, meningeal cells and
20
oligodendrocytes were reported to be infected as a result of impaired type I IFN response,
suggesting that separate antiviral cytokines may protect neurons and glial cells of CNS
(Fragkoudis et al. 2007; Ryman et al. 2000). Interestingly, neither SFV nor VSV infection was
lethal in IFNGR-knockout mice, while vaccinia infection killed IFNAR-knockout and
IFNGR-knockout mice equally efficiently (Muller et al. 1994).
Type I IFN knockout mice are also vulnerable to Sindbis infection and the virulence is
further enhanced in type II IFN receptor deficient animals (Gil et al. 2001). Surprisingly,
despite the clear role of STAT1 in mediating type I-III interferon responses, STAT1
knockout mice are resistant to Sindbis virus infection reflecting the importance of STAT1independent IFN signaling pathways (Gil et al. 2001). Infection with another alphavirus,
Chikungunya (CHIKV) nevertheless resulted in the rapid death of mice with STAT1knockout genotype (Schilte et al. 2010). Remarkably, two transcription factors Zfx and Mga
were recently discovered in in vitro and in vivo RNA interference (RNAi) screens to be
master regulators of ISG induction by type I IFN, with their knock-down affecting the
basal, IFN- and Sindbis-virus induced expression of several crucial components of the
type I IFN pathway: RIG-I, STAT1, STAT2, IRF7, IRF9, IFNAR1, IFNAR2, IKK and IFN-
(Varble et al. 2013).
SFV infection induces a robust secretion of interferons in vitro and in blood and tissues in
vivo (Bradish and Allner 1972; Bradish, Allner, Fitzgeorge 1975; Ylosmaki et al. 2013). Blood
interferon levels peak 36h p.i. and are absent by 5d p.i. (Bradish, Allner, Fitzgeorge 1975).
Virulent SFV strains induce higher blood interferon levels than the avirulent SFV strains
(Alsharifi et al. 2005; Fleming 1977). Similarly, virulent Sindbis virus induced higher type I
IFN serum levels than the avirulent counterpart (Trgovcich, Aronson, Johnston 1996).
Initiation of alphavirus infection clearly depends on the inoculation dose, and while i.p.
injection of 30 PFU of L10 was benign in mice, 1000 PFU almost uniformly caused lethal
infection if the animals had not first been primed by a high dose i.p. dsRNA injection
(Bradish and Titmuss 1981). Cytosolic RLRs are pivotal for the host detection of alphavirus
infections such as that by CHIKV (Werneke et al. 2011). In accordance, RIG-I was found to
be a powerful SINV inhibitor especially in conjunction with Zinc-Finger Antiviral protein
(ZAP) while MDA-5 had a lesser effect (Karki et al. 2012). It is noteworthy that viral RNA
can be detected by one or several members of the RLRs depending on the virus-species
(Goubau, Deddouche, Reis e Sousa 2013). In the recent study of Nikovov et al. both MDA5
and RIG-I were important for SFV genome detection and IFN- induction. Strikingly,
however, these pattern recognition receptors (especially RIG-I) could also sense host cell 5’ppp dsRNA secondary structures generated by viral RNA-dependent RNA-polymerase
(RdRp) through transcription of host cell RNA (Nikonov et al. 2013). Nikonov et al. further
proposed that this effect correlates with alphavirus virulence as RdRp from avirulent SFV
induced more IFN- than the polymerase from the virulent strain, but this was probably
due to the reduced JAK/STAT inhibitory power of the used nsP2-NLS mutant.
Furthermore, they showed that type I IFN induction resulted in the interruption of SFV
mutant replication due to destruction of its positive-sense RNAs. If the mechanism
postulated by Nikonov et al. is confirmed by others in the future, it may represent a
groundbreaking and completely novel evolutionary tract for cell autonomous pathogen
detection, where in addition to microbial structural components and genomes also their
21
action on host RNAs may be detected by PRRs. Whether the transcription of host RNA by
viral RdRp postulated by Nikonov et al. is only inadvertent and thus deleterious to the
virus or whether it plays a role in the viral lifecycle remains to be shown (Nikonov et al.
2013). One also has to appreciate the difficulty of controlling the IFN- induction
experiments, as even plasmid DNA transfections can induce a type I IFN response (Rautsi
et al. 2007). Thus, in the future, it will be important to verify the findings of Nikonov and
colleagues by alternative methods such as transfection of the viral polymerase proteins.
Like most viruses, SFV has developed mechanisms to avoid IFN response and these include
rapid shutoff of the host macromolecular synthesis in the infected cells. A portion of the
cytoplasmic nsP2 protein of SFV and other Old World alphaviruses is transported to
nucleus utilizing the nuclear localization signal (NLS) of the protein. In the nucleus, nsP2
induces ubiquitination and complete degradation of a catalytic subunit of RNA polymerase
II, Rpb1, thus interrupting host cell transcription. This occurs in 6 h post infection and thus
even before the reported inhibition of STAT1 phosphorylation. nsP2 is an extremely
versatile protein and in addition to its role in host cell transcription shutoff it functions as a
protease in the replicase complex and has two additional enzymatic activities, a helicase
and an RNA 5’ triphosphatase (Akhrymuk, Kulemzin, Frolova 2012). It will be interesting
to see, whether the RNA 5’ triphosphatase activity proves to be an additional means for
some alphaviruses to evade PRR detection of viral (or host) RNA-intermediates, as short,
blunt end and double stranded 5’ triphosphate RNA is the alleged substrate for RIG-I
(Schlee and Hartmann 2010). New World alphaviruses have acquired an alternative way to
suppress host cell transcription by blocking nuclear pores with a viral capsid protein.
Moreover, following SFV infection, PKR-mediated eIF2 phosphorylation, which is one of
the prototypical cell autonomous antiviral responses, preferentially blocks the translation of
the host cell mRNA, as translation of SFV structural proteins is not dependent on eIF2 due
to the complementary viral RNA secondary structures (Akhrymuk, Kulemzin, Frolova
2012; Garmashova et al. 2007). Amongst the alphaviruses at least CHIKV seems to have also
some more specific, yet uncharacterized, mechanisms for translational shutdown of
infected cells in addition to PKR induction (White et al. 2011).
Strikingly, neither of the two classical IFN- response pathways OAS/RNaseL or PKR, is
central for SFV-infected host survival based on in vivo knock-out experiments. While PKR
seems to be dispensable in vivo, it does possess an anti-SFV function as a cytosolic PRR
inducing type I IFNs (Barry et al. 2009). ZAP has been reported to act in concert with
several other ISGs to inhibit alphavirus replication, at least in vitro by blocking specifically
the translation of viral (but not host) RNA. In fact, viral translation inhibition seems to be a
common effector function also for many other ISGs (Schoggins et al. 2011). ZAP is induced
by type I and type II IFNs but also by the IRF3 homodimer which is activated by PRR
engagement directly after PAMP detection, and thus ZAP cannot be regarded as a “true
ISG” (Karki et al. 2012; Wang et al. 2010b). This seems to be the case also for another
important anti-alphavirus effector, Viperin (White et al. 2011) (Figure 2).
In addition to global transcriptional- and translational shutdown, different alphaviruses are
also specifically capable of inhibiting STAT1 and STAT2 phosphorylation by some still
incompletely understood mechanisms. This inhibition seems to be mediated either by the
22
nsP1 (Sindbis) or nsP2 genes (SFV) and it correlates with the virulence of SFV and other
alphaviruses (Breakwell et al. 2007; Cho et al. 2013; Deuber and Pavlovic 2007; Farmer et al.
2013; Fros et al. 2010; Simmons, Wollish, Heise 2010; Wollish et al. 2013; Yin et al. 2009).
Interestingly, the NLS of Semliki Forest virus nsP2 is required for the suppression of type I
IFN response, but the effect is independent of host cell gene expression shutoff (Breakwell
et al. 2007). Furthermore, the nsP2 gene has recently been reported to have yet another role
in the evasion of type I IFN response, at least for CHIKV, by blocking nuclear translocation
of STAT1 (Fros et al. 2013). Intriguingly, the prominent neurovirulence factor for SFV lies in
nsP3, but the functions of nsP3 in the replicase complex, or possibly as an independent
protein are mostly unknown (Tuittila et al. 2000). In view of the important roles of nsP1 and
nsP2 in the inhibition of IFN responses and in alphavirus neurovirulence, it will be
interesting to determine whether nsP3 serves a similar purpose.
EEEV is a highly virulent Alphavirus and it does not induce any marked systemic IFN-
production in infected mice. On the contrary, the virulence of VEEV is based on its ability
to replicate despite the presence of even high (80000 IU/ml) IFN- concentrations in the
serum (Ryman and Klimstra 2008). The systemic type I IFN is probably produced by pDCs
(Cao et al. 2008). Very recently, the virulence of VEEV was mapped to specific secondary
5’UTR motifs that functionally mimic 5cap 2'-O methylation of higher eukaryotic mRNAs
to prevent inhibition by IFIT1 ISG (Hyde et al. 2014). IFIT1 is a multifunctional ISG, with
one of its most prevalent functions being the inhibition of viral mRNA translation
(Diamond and Farzan 2013). As an interesting side note, some tumors can stimulate pDC
differentiation, which together with MDSCs, inhibits adaptive immune responses by
evoking T-cell tolerance (Rabinovich, Gabrilovich, Sotomayor 2007). It is perceivable that in
some cases this may, as a secondary effect, prime for rapid attack against oncolytic viruses
in cancer patients receiving virotherapy.
In summary, alphaviruses have developed tools to avoid IFN- induction and/or
signaling most likely due to high evolutionary pressure. These viral mechanisms include
universal shutdown of host cell transcription and translation and inhibition of STAT1
phosphorylation and traffic, and for some alphaviruses the ability to replicate “under the
radar” without triggering any marked type I IFN induction. Some alphaviruses are also
capable of specifically inhibiting host cell translation. The specific mechanisms for these
evasive maneuvers remain to be discovered as does the potential role of SFV
neurovirulence factor nsP3. MicroRNAs are a newly recognized field of anti-alphavirus
innate immunity. Antiviral microRNA expression has very recently been shown to regulate
the viral tropism and to increase pathogenicity of EEEV by limiting the virus replication in
myeloid lineage cells (Trobaugh et al. 2014).
23
2.4 ONCOLYTIC VIROTHERAPY
2.4.1 From the early 20th century to the present day
Oncolytic virotherapy involves targeting cancer, its microenvironment or
existing/developing vasculature with replicating (modified or natural) viruses, and is
distinct, though overlapping, with gene therapy and the immunotherapy of cancer. The
first reports that viruses may be able to eradicate cancer date back to early 20th century
(Vaha-Koskela, Heikkila, Hinkkanen 2007) and involved incidental cancer remissions after
either natural viral infections or vaccinations. Interestingly, the reported cases were almost
exclusively either lymphomas or leukemias. For instance, a patient suffering from chronic
leukemia had a remarkable decrease of white blood cells after having “flu-like” symptoms
in 1904. Other reported examples of partial and occasionally even complete responses
between the period 1950s-1980s included measles virus and vaccine induced remissions of
multiple different lymphomas and leukemias and vaccinia virus vaccination-induced
remission of leukemia. In addition to hematological malignancies, one of the early reports
described Newcastle disease virus (NDV) induced response of Chicken farmer’s gastric
cancer after occupational exposure to the virus (Liu, Galanis, Kirn 2007). Furthermore, a
patient with cervical cancer experienced a spontaneous remission after immunization with
rabies vaccine in 1912. There are several potential reasons why these fortunate responses to
vaccinations or natural viral infections were primarily restricted to leukemias and
lymphomas. Firstly, because of the hematologic origin of these diseases, leukemia and
lymphoma patients are inherently immunosuppressed, which could facilitate systemic
delivery of opportunistic viruses and attenuated vaccine strains. Secondly, leukemias and
lymphomas are frequent also among younger patients, who are the individuals most likely
to be vaccinated or to contract a primary virus infection (Liu, Galanis, Kirn 2007). Finally, at
least one of the reported responsive hematologic neoplasms, Burkitt’s lymphoma is caused
by Epstein-Barr virus of the Herpesviridae family, which encodes type I IFN antagonizing
gene LF2 (Wu et al. 2009). Hence, it is tempting to speculate that genes expressed from
oncoviruses rendered some of these tumors to become good targets for viral infections.
The reports of spontaneous cancer remissions after viral infections led to a series of patient
treatments with non-engineered viruses starting from 1950s. This was the “first generation”
of oncolytic viruses and the strains used were typically attenuated through serial in vitro
passages. Mumps, measles, West Nile virus, adenoviruses, reovirus, NDV and vaccinia
virus were tested with different tumors and different administration routes. Interestingly,
vaccinia induced tumor remissions were occasionally complete and long lasting, and could
occur at sites distant from those of viral administration. In general, the treatments were
well tolerated, with the most common side effects being pain at the virus injection site and
“flu-like” symptoms. Nevertheless, approximately 10% of the patients treated with the
West Nile virus were diagnosed with encephalitis (Liu, Galanis, Kirn 2007) and a case of
severe encephalitis was also reported in one out of four leukemia patients treated with
Bunyamwera (Moore 1954). Many other viruses were also tested in both animal models and
in humans during the 1950s and 1960s, but the experiments were often poorly controlled
24
and in general, the results remained modest. For example, one patient with an acute
leukemia received as many as six different viruses in succession, while in another study 13
different viruses or strains were provided to 57 patients with varying malignancies but no
therapeutic effect was observed (Moore 1954; Vaha-Koskela, Heikkila, Hinkkanen 2007).
In parallel with the advances in modern biotechnology, since the 1990s virotherapy has
been again actively developed to complement the traditional cancer treatment regimens.
The second generation of oncolytic viruses was engineered to carry mutations or deletions
in one or more viral genes essential for virus replication in the natural target tissue but
which were redundant in neoplastic tissue to create conditionally replicating viruses (See
chapter: “2.4.4 Targeting of Oncolytic Viruses”). The mutations were often introduced into
virulence genes responsible for inhibiting cell autonomous antiviral (and antiproliferative)
responses in healthy tissues, responses of which are often already suppressed in tumors.
While this approach increased the tumor selectivity of the viruses, on the one hand, and
improved their safety on the other, these improvements occasionally came with the cost of
reduced viral replication power (Thorne, Hermiston, Kirn 2005)).
There are several requirements for virus design that need to be addressed in the
construction of a good oncolytic viral agent. First, the virus must be able to efficiently
replicate in and destroy the targeted tumor tissue, but it is equally important that the virus
spares the normal, healthy tissue from extensive damage. The genetic stability of the viral
agent is desirable for both safety and commercial manufacturing reasons. It is also
preferable to use non-integrating viruses to avoid the risk of insertional
mutagenesis/oncogenesis. By using non-integrating viruses, one can also be more confident
that the virus will take part in the lytic cycle, which is usually considered an important part
of the anti-tumor effect of oncolytic viruses. Incorporating a safety mechanism (“off
switch”) to inactivate the virus if necessary should also be taken into consideration,
whenever the virus is not inherently responsive to antiviral agents. Finally, the oncolytic
virus needs to be suitable for high-titer manufacturing and purification processes following
the “Good Manufacturing Practices” required for clinical studies (Thorne, Hermiston, Kirn
2005).
The third generation conditionally replicating oncolytic viruses are deletion mutants
“armed” with exogenous genes encoding for example cytotoxic or immunostimulatory
proteins increasing the oncolytic efficacy or suicide genes which additionally increase
safety by allowing down-regulation of the infection if necessary (O'Shea 2005). Several viral
vectors have been used in clinical trials based on e.g. vaccinia virus, adenoviruses, HSV-I,
NDV and reovirus type 3 (Vaha-Koskela, Heikkila, Hinkkanen 2007). A list of the current
(5/2014) oncolytic virotherapy trials reported in clinicaltrials.gov can be found in Table 1.
The top 3 oncolytic viruses involved at the moment in either phase III or advanced phase II
trials are reovirus (Reolysin), HSV-I (Talimogene laherparepvec, T-VEC, previously known
OncoVEXGM-CSF) and Vaccinia virus (JX-594). T-VEC and JX-594 harbor immunostimulatory
granulocyte-macrophage colony-stimulating factor (GM-CSF) gene, while Reolysin is an
unmodified reovirus (Donnelly et al. 2013a; Heo et al. 2013; Miest and Cattaneo 2014). The
main immunostimulatory function of GM-CSF in the context of oncolytic
immunovirotherapy is thought to be the activation of professional antigen presenting cells
25
(Dranoff et al. 1993; Parato et al. 2012). While NK-cell activation has often been listed as one
of the effects of GM-CSF (Bartlett et al. 2013; Cerullo et al. 2010) there seems to be no
convincing literature to back this claim. Surprisingly, GM-CSF has been recently reported to
play a strong immunosuppressive role in some cancers (Bayne et al. 2012; Kohanbash et al.
2013).
26
Table 1. Ongoing oncolytic virotherapy clinical trials. In the table the clinical trial names, phase
and therapeutic viruses together with possible co-treatments are listed as reported “Active” or
“Recruiting” in clinicaltrials.gov 16.5.2014
27
Replication competent, oncolytic viruses may reach even metastasized tumors and thus be
used as a noninvasive means to target tumors in hard-to-reach body locations such as the
brain. For example, intravenous administration of replication-competent NDV led to tumor
eradication in several patients with multiresistant GBM, with prolonged survival of more
than 9 years and high quality of life (Csatary et al. 2004). Different HSV-I mutants,
adenoviruses and reovirus constructs have also been clinically tested for treatment of
malignant glioma. Additionally, phase I glioma therapy trials have recently been started for
modified polio and measles viruses. In a phase I clinical trial, single adenovirus (Ad) DNX2401 injection into GBM tumors as a first line therapy resulted in notable responses,
including complete regression of the disease, in many of the 24 treated patients (Pol et al.
2013). Equally impressive results were reported from a phase I GBM trials using local
administration of lentivirus Toca511 as the first line treatment in combination with prodrug
TocaFC, and in another study where poliovirus PVS-RIPO was used as the second line
treatment combined with resection. Remarkably, three of the seven patients who received
PVS-RIPO achieved a complete response. In general, administration of the viruses into the
brains of patients has been found to be both safe and well tolerated (Pol et al. 2013; Zemp et
al. 2010) (Table 1).
Several different GM-CSF expressing viruses have proven to be clinically effective against
different solid tumors, thus underlining the importance of stimulating the anti-tumor
immunity in order to achieve a good therapeutic efficacy. GM-CSF –encoding oncolytic
herpesvirus T-VEC has completed phase III clinical trial in patients with unresectable
metastatic melanoma, where the virus proved both safe and effective (response rate 26%).
As many as 11% of the patients have been reported to experience complete response
(disappearance of the disease), and the primary end point of durable response rate was
met. However, T-VEC failed to meet its secondary endpoint of improved overall survival,
even though a clear beneficial trend (p=0.051) compared to control group was seen
(Williams 4th of April 2014). T-VEC is the first oncolytic virus to have completed phase III
trial, giving hope to obtain the first approved oncolytic virus in Western World (Andtbacka
et al. 2013; Pol et al. 2013). The phase III melanoma trial design provoked some ethical
debate, as the control group received a subcutaneous injection of GM-CSF instead of
standard care. T-VEC has shown some efficacy also in phase I/II trial of stage III/IVA
squamous cell carcinoma of the head and neck (Sivendran et al. 2010). Curiously, one phase
III trial (ClinicalTrials.gov identifier: NCT01161498) for this malignancy was suddenly
terminated in early July 2011 “to permit significant modification of clinical trial design
mandated by the changing therapeutic landscape for patients with SCCHN” (Regan 29th of
July 2011).
Vaccinia virus JX-594 expressing GM-CSF has also completed phase I and phase II clinical
trials and demonstrated promising efficacy in the treatment of primary and metastatic liver
cancers (Breitbach et al. 2010). In a randomized phase II dose-finding trial, the overall
survival of the high-dose group compared to low-dose group was significantly better
(Donnelly et al. 2013a; Heo et al. 2013; Koski et al. 2009). A phase IIb trial is ongoing and
even though the official results remain to be published, according to preliminary
information, JX-594 failed to meet the primary endpoint of improved overall survival
(Sheridan 4th of September, 2013).
28
GM-CSF-encoding oncolytic adenoviruses Ad5-D24-GM-CSF and Ad5/3-D24-GM-CSF
have also been tested in patients in the treatment of several different solid tumors. These
viruses were reported to induce antitumoral immunity and to exert antitumoral effects
some of which were complete, but the therapy effects on overall survival were not reported.
A phase I/II clinical trial with Ad5/3-D24-GMCSF combined with low-dose metronomic
cyclophosphamide to reduce T-regs is ongoing (Cerullo et al. 2010; Koski et al. 2010).
Finally, after promising results in a phase II trial, an unmodified, oncolytic reovirus
Reolysin has entered phase III trial in patients with head and neck carcinoma for use in
combination with chemotherapy. In the phase II trial one complete response and six partial
responses were seen in 17 patients who received intravenous Reolysin in combination with
paclitaxel and carboplatin treatment. In addition, two patients experienced major clinical
responses (complete resolution of skin nodules). The safety profile was in general good but
one patient experienced sepsis, not a surprising toxicity considering the
immunosuppressive nature of the chemotherapeutic agents being combined with a
replicative virus (Breitbach et al. 2010; Karapanagiotou et al. 2012).
2.4.2 Oncolytic virotherapy overlaps with gene therapy and immunotherapy of cancer –
emerging concept of oncolytic immunovirotherapy
Gene therapy, i.e. (mostly virus vector mediated) transfer of therapeutic genes, has proven
very promising for the treatment of monogenic diseases but the results in cancer treatment
have remained modest (Brenner et al. 2013). Cancer is nevertheless the most frequently
chosen disease for gene therapy trials and 64.4% (1186/1843) of all gene therapy trials
completed or ongoing before 2012 concerned cancer (oncolytic virotherapy trials included)
(Ginn et al. 2013). The concepts of gene therapy of cancer, immunotherapy of cancer and
oncolytic virotherapy are overlapping, and while viral gene transfer vectors or replicons are
often used to deliver immunostimulatory genes or cancer antigens to tumors or immunecells (Lam et al. 2013), the same genes and antigens can often be incorporated into
replicative oncolytic viruses (Russell, Peng, Bell 2012). For example, Rommelfanger et al.
recently used the mouse B16 melanoma model and combined systemic VSV carrying
human melanoma antigen to adoptive cell transfer with encouraging results
(Rommelfanger et al. 2012). A completely novel paradigm was introduced, when instead of
selected TAA or TSA, a cDNA library prepared from healthy human prostate was cloned
into VSV and injected systemically into mice to successfully treat murine prostate tumors
(Kottke et al. 2011). The human prostate cDNA library worked significantly better than
mouse cDNA library demonstrating that “altered self” epitopes can be more immunogenic
as compared to self epitopes, and it will be interesting to determine in the future whether
this holds true when single, universal TAAs such as survivin or telomerase are used as
vaccines. Interestingly, the treated mice experienced no autoimmune symptoms and it was
postulated that this was because of lack of inflammatory signals in the prostate. However,
Kottke and coworkers used a subcutaneous prostate tumor model and thus it remains
unclear whether the primary organ inflammation caused by orthotopic tumors would be
sufficient to induce autoinflammatory disease in this therapeutic context.
29
Cell death can cause tolerance or be immunogenic depending on the type and location of
the dying cell and the factors it releases or expresses on its surface during the death process
(Griffith and Ferguson 2011). The effective maturation of dendritic cells resulting in
successful presentation of tumor antigens to T-cells in lymph-nodes requires PAMP
detection (or other maturation stimulus) upon engulfment of the dying cells, and a lack of
the maturation stimulus can promote tolerance (T-cell deletion, anergy or T-reg induction)
instead of antitumor immunity (Mellman, Coukos, Dranoff 2011). Oncolytic viruses kill
tumor cells directly through apoptosis and necrosis, but they have also been shown to
induce adaptive antitumor immune responses via immunogenic cell death (ICD) (Bartlett et
al. 2013; Boozari et al. 2010; Donnelly et al. 2013b; Miyamoto et al. 2012; Workenhe and
Mossman 2014). Different oncolytic viruses have been reported to cause all four major types
of cell death, apoptosis, necrosis, autophagy and pyroptosis, and often the death of a cell
population is a mixture of these processes (Bartlett et al. 2013; Fink and Cookson 2005).
Alphaviruses kill cells primarily by apoptosis evoked by nsP2 of Old World viruses and the
capsid protein of New World viruses, while also the envelope glycoproteins are proapoptotic (Barry et al. 2010; Jose, Snyder, Kuhn 2009). SFV-infected cancer cells have been
reported to die both via p53 independent apoptotic (Glasgow et al. 1998; Murphy et al.
2000; Murphy, Sheahan, Atkins 2001; Smyth et al. 2005) and non-apoptotic pathways
(Vaha-Koskela et al. 2006). Death by non-apoptotic pathways might result from the fact that
varying apoptotic signaling pathways become silenced in the course of tumorigenesis and
thus the death of these cells may resemble the necrotic mode of death of alphavirusinfected apoptosis-resistant mosquito cells (Jose, Snyder, Kuhn 2009).
The mode of death of oligodendrocytes (Fazakerley et al. 2006; Glasgow et al. 1997) and
that of adult neurons infected by neurovirulent alphavirus strains remain incompletely
understood as both necrosis and apoptosis have been documented, whereas immature
neurons that are susceptible also to avirulent strains die by apoptosis (Glasgow et al. 1997;
Griffin 2005; Nargi-Aizenman and Griffin 2001). SFV infected cells accumulate
autophagosomes and this effect is viral glycoprotein-dependent, but seems to be a
bystander phenomenon as the disruption of autophagy has no effect on the viral replication
(Eng et al. 2012). According to a very recent study SFV induced apoptosis of healthy
mammalian cell lines depends on the PRR signaling mediator MAVS (aka CARDIF, VISA,
IPS-1) that activates caspases 8 and 3, and interestingly is independent of the simultaneous
presence of IFN- or even IRF3/7 signaling (El Maadidi et al. 2014). As a side note, since all
dead cells are often called necrotic in pathological terminology, the inflammatory “necrotic
death process” resulting in the release of cellular contents into the extracellular space is
occasionally termed oncosis (Fink and Cookson 2005). Caspase-independent necrosis and
autophagy, and caspase-1-dependent pyroptosis are all different types of ICD, whereas
caspase-dependent apoptosis can be either immunogenic or tolerogenic. In addition to
oncolytic viruses, also some other cancer therapies such as anthracyclines, oxaliplatin,
cyclophosphamide (CPA), EGFR-specific antibodies and UV irradiation can promote ICD.
The common denominator for ICD is endoplasmic reticulum (ER) stress sensed by
unfolded protein response (UPR) causing the expression and release of danger (or damage)
associated molecular patterns (DAMPs) calreticulin, ATP and HMGB1. ER-stress results in
calreticulin traffic to the cell surface (PRR binding “eat me” signal for APCs), ATP release
30
(“find me” signal) and HMGB1 release, which in its fully reduced or disulphide-bond
possessing states (but not in the oxidized state) stimulates DC maturation, Th1 cell
expansion and works as a chemoattractant (Bartlett et al. 2013; Workenhe and Mossman
2014). Importantly, apoptosis induced by replicative SFV (but not by replicon) was reported
to involve an ER-stress response evoked by envelope glycoproteins (Barry et al. 2010).
During their replicative cycle oncolytic viruses trigger multiple pattern-sensing
mechanisms resulting in the secretion of different cytokines such as type I IFNs and leading
to ER-stress. Virus-induced inflammation, in turn, summons and activates antigenpresenting cells, which prime antiviral adaptive immunity. Since virus infection-induced
inflammation is highly immunogenic in nature and optimally occurs only in cancer cells,
oncolytic viruses may be able to break tolerance against weak tumor-associated (self)
antigens (TAA) and thus unmask those tumor cells otherwise hidden from immune
surveillance. The immunogenicity can be further boosted by deleting viral
immunosuppressive genes and adding stimulatory cytokines such as IL-12, IL-18, IL-2,
GM-CSF, CD40L, B7.1 and CCL5, or PRR agonists such as CpG DNA-elements or a variety
of heat shock proteins. Heat shock proteins are especially interesting immunostimulants as
not only do they act as direct DAMPs on cell surface or upon release, but they can also bind
tumor antigens thus enhancing their presentation (Bartlett et al. 2013). Indeed, depending
on the virus and tumor model, the oncolytic efficacy may rely almost exclusively on such
bystander immune responses against the tumor rather than lytic virus replication (Cheema
et al. 2013; Galivo et al. 2010; Prestwich et al. 2009; Wongthida et al. 2010). However, in
other occasions, the initial tumor debulking by the replicative virus is followed by
eradication of residual disease by T-cells of activated adaptive immune system (Huang,
Guo, Hwang 2012; Naik et al. 2012). While the initial debulking by lytic replication cycle
and/or inflammation induced innate immune destruction has been demonstrated in animal
models, the debulking function remains to be demonstrated clinically (Russell, Peng, Bell
2012). Occasionally, efficient tumor infection and destruction in preclinical models does not
seem to require any adaptive immune responses at all (Lun et al. 2005).
In a preclinical settings, oncolytic virotherapy has been successfully combined with
different kind of immunotherapies such as adoptive T-cell transfer (Rommelfanger et al.
2012), DC-vaccination (Woller et al. 2011) and agonistic or antagonistic monoclonal
antibodies against immune-checkpoint regulators (Dias et al. 2012; John et al. 2012;
Quetglas et al. 2012a; Zamarin et al. 2014). The multitude of different successful oncolytic
immunovirotherapy approaches suggests that live viruses are indeed powerful adjuvants,
and the infection-related inflammation can induce Th1 type of antitumoral immunity. Thus,
the viral infection can counteract the tolerogenic and anti-inflammatory Th2 type of
immunity, which is typical for the tumor microenvironment and resembles a physiological
wound healing process. Finally, classical chemotherapeutics and radiotherapy have been
successfully combined with OV. For instance, in a recent study, doxycycline treatment was
noted to augment the tumor-specific replication of oncolytic vaccinia virus. Doxycycline
also boosted anti-cancer immunity by enhancing and stabilizing the expression of the MIC
stress ligand on cancer cells making them more attractive targets for adoptive cell transfer
of cytokine induced killer cells (CIK) (Tang et al. 2013). Intriguingly, the best result was
obtained when CIK cells were used as a carrier for oncolytic vaccinia virus. Similarly, a
31
combination of oncolytic HSV-1 with an ICD inducer mitoxantrone was recently
successfully employed to break immunotolerance against TUBO breast cancer in syngeneic
mice (Workenhe et al. 2013).
Alphaviruses have been reported to replicate preferentially in tumor cells. This can result
from, firstly, from defects in the cancer cell type I IFN induction or secondly, and as
growing body of evidence suggests more importantly, from defects in the type I IFN
responsiveness of cancer cells. In addition to neoplastic cells, also most of the healthy
stromal cells can produce type I IFNs, and thus any defects cancer cells may have in their
type I IFN induction are probably not sufficient to make them vulnerable to viral infections
(Huang, Guo, Hwang 2012). Alphavirus mediated oncolysis can stimulate antitumor
immunity (Huang, Guo, Hwang 2012) and by using SFV replicons, this effect has been
shown to be greatly enhanced when immunostimulatory cytokines such as IFN- (Quetglas
et al. 2012b) or IL-12 were incorporated into the virus (Rodriguez-Madoz, Prieto, Smerdou
2005; Rodriguez-Madoz et al. 2009; Rodriguez-Madoz et al. 2014). It is uncertain whether
SFV can induce cross-priming and cytotoxic T-cell responses by directly infecting DCs. It
has been shown that a lentivirus vector pseudotyped with SFV glycoproteins gained
dendritic cell tropism by enabling transduction via c-type lectin receptor (Froelich et al.
2011), but Schulz and colleagues could not achieve efficient DC infection with the SFV
replicon (Schulz et al. 2005). Nevertheless, SFV was able to induce efficient TLR3 mediated
activation of DCs and cross-priming of cytotoxic T-cells after phagocytosis of viral dsRNAcontaining cells (Schulz et al. 2005). The controversial data regarding direct infection of DCs
by alphaviruses might stem from the fact that alphaviruses can change their tropism as a
result of adaptation to cell culture conditions (Smit et al. 2002). Furthermore, only Sindbis
viruses produced in mosquito cell lines but not in mammalian cells were reported to be
able to use c-type lectins as their receptor and infect DCs (Klimstra et al. 2003).
Interestingly, TLR5 stimulating bacterial Flagellin was recently found to increase both Th1
and Th2 adaptive responses when cloned into an alphavirus replicon (Knudsen et al. 2013).
SFV infection per se is also known to induce the global activation of B- and T-cells of mouse
in a type I IFN-dependent manner (Alsharifi et al. 2005) and thus SFV infection may not
only stimulate local immune responses in the tumor microenvironment but also have
broader immunostimulatory effects.
IL-12 gene transfer and expression from oncolytic viruses, vectors or replicons, is not a
straight-forward process as the cytokine is a heterodimer and it is systemically toxic to
animals. However, strategies have been developed to localize the expression to tumor beds
(Pan et al. 2012). Notably, SFV-IL-12 was reported to be a more potent immunotherapy
vector than adenovirus with the same cytokine insert in two separate syngeneic rodent
tumor models. The authors speculated that this presumably resulted firstly from the lower
immunogenicity of SFV as compared to adenovirus making multiple intratumoral
injections possible, secondly from SFV replicon induced apoptosis and thirdly due to a
more efficient IL-12 production from the SFV replicon (Rodriguez-Madoz et al. 2009).
Replicons and fully replicative viruses might be a better option for cancer immunotherapy
than the classical vectors, which are often based on adeno- or lentiviruses or DNA-plasmids
(Matrai, Chuah, VandenDriessche 2010; Volpers and Kochanek 2004) as the classical viral
vectors do not replicate their genome inside the cells, possibly resulting in poorer gene
32
expression and reduced vector induced cell death. In addition, unlike many other viruses
such as myxomavirus (Ogbomo et al. 2013), alphaviruses are not known to suppress
adaptive immune responses. Indeed, in a very recent study, the SFV-IL-12 vector proved to
be slightly more efficient than the same cytokine expressed from a plasmid against genetic
mouse hepatocellular carcinoma (HCC) model, even though higher and longer term IL-12
expression was obtained with the plasmid. The authors proposed that SFV was superior
because it induced direct tumor cell apoptosis and provoked an immunogenic type I IFN
response. Furthermore, short term expression was postulated to increase the safety when
highly toxic cytokines such as IL-12 were used (Rodriguez-Madoz et al. 2014). This study
also confirmed that SFV-IL-12 has efficacy also against antigenically heterogeneous forms
of cancer, a question that had remained unanswered after the only transient remissions
seen previously in a woodchuck HCC model (Rodriguez-Madoz et al. 2009). Recently, also
HSV-IL-12 showed a remarkable efficacy as compared to an unmodified virus in a
challenging mouse glioma stem cell model (GSC) demonstrating a multimodal anti-cancer
action consisting of direct killing of both GSCs and differentiated cancer cells,
antiangiogenesis and T-reg reduction (Cheema et al. 2013). While both the unmodified and
IL-12-encoding viruses were able to directly kill cancer cells, the antiangiogenic and T-cell
mediated bystander effects remained dependent on the ectopically expressed IL-12.
While SFV infection is immunogenic, one has to keep in mind that a productive viral
infection combined with the interferon signaling, or even the plain nsP2 protein of SFV
derived oncolytic viruses can dramatically change the host cell gene expression
(Akhrymuk, Kulemzin, Frolova 2012; Sadler and Williams 2008). While this might be
beneficial for the purposes of inflicting direct cell death, it inevitably also changes the
antigenic profile of the infected tumor cells compared to distant uninfected cancer cells,
possibly biasing the adaptive immune responses towards infection-related antigens. Granot
et al., recently presented an interesting immunotherapy approach to circumvent such
pitfalls and treated LacZ transduced CT26.CL25 tumors with a Sindbis replicon expressing
LacZ (Granot, Yamanashi, Meruelo 2014). While their replicon could not actually infect the
tumors, LacZ was expressed in the draining lymph nodes, and immunotherapy resulted in
an efficient tumor eradication and cross-protection against CT26WT “immune-escape”
variants via epitope spreading. Second obvious caveat to the use of replicative viruses for
induction of cytotoxic cellular and opsonizing humoral (i.e. Th1 biased) immune responses
relates to the fact that viral infections often trigger a Th2 neutralizing antibody response
(Abbas, Lichtman, Pallai 2007). This can possibly be overcome by adding exogenous Th1
cytokines or ligands such IL-12 (Cheema et al. 2013) or soluble CD40L (Pesonen et al. 2012)
into the therapeutic viruses. However, the identification of soluble CD40L in the serum of
cancer patients has been recently implicated in immunosuppression (Huang et al. 2012).
The beneficial and detrimental roles of innate, adaptive and cell autonomous immune
responses in the context of oncolytic virotherapy remain incompletely understood and may
depend on the tumor model, therapy virus, administration route and timing used. At least
the complement system (Ikeda et al. 2000; Wakimoto et al. 2002) and neutralizing
antibodies (Nabs)(Ilett et al. 2009) have been reported to inhibit oncolytic viruses. Notably,
virus neutralizing antibody responses can occasionally be very fast rather than slow and
adaptive. Preimmune IgM antibodies were reported to be able to neutralize oncolytic HSV-
33
1 and very recently also VSV, and the neutralization was further enhanced when
complement proteins were present (Ikeda et al. 1999; Tesfay et al. 2014). In addition to
direct neutralization effects, complement proteins, antibodies, different coagulation factors
and possibly also other serum proteins can also opsonize viruses by sequestrating them to
allow their destruction by phagocytotic macrophages in the spleen and Kupffer cells in the
liver (Russell, Peng, Bell 2012). The successful preclinical and clinical approaches to evading
the viral neutralization and sequestration have included the use of tumor homing cell
carriers such as dendritic cells, T-cells (Ilett et al. 2009), mesenchymal stem cells (GarciaCastro et al. 2010) and irradiated myeloma cells (Liu, Russell, Peng 2010), pharmacological
ablation or saturation of antibodies, complement and phagocytes (Haisma and Bellu 2011;
Ikeda et al. 1999; Ikeda et al. 2000; Koski et al. 2009; Magge et al. 2013; Shashkova et al. 2008;
Ziegler et al. 2002), shielding the therapeutic virus with polymers (Eto et al. 2008; Fisher
and Seymour 2010) and serotype switching (Sarkioja et al. 2008). Priming of T-cell
immunity is frequently observed in oncolytic virotherapy and often considered as essential
for the therapeutic efficacy (Naik et al. 2012), but complete and long lasting tumor
regressions after virotherapy have been documented also in athymic animal models (Lun et
al. 2005).
The distinct contributions of CD4+ T-helper cells and CD8+ cytotoxic T-cells for oncolytic
virotherapy remain equivocal and while both classes have been reported to be necessary
(Kottke et al. 2010), occasionally the CD4+ cell population seems to be redundant (Diaz et
al. 2007; Rommelfanger et al. 2012; Sobol et al. 2011). Neutrophils were reported to be the
main effector in OV when VSV was used in CT26LacZ mouse colorectal tumor model
(Breitbach et al. 2007) and neutrophils played an important role also in the efficacy together
with CD4+ and CD8+ T-cells when HSV-1 was combined with the immunogenic cell death
inducer, mitoxandrone (Workenhe et al. 2013). Most intriguingly, depletion of any of the
CD8+, CD4+ or NK-cell classes impaired the efficacy in a mouse melanoma model when
reovirus was combined with an intratumoral VEGF pulse to infect and destroy cancer
vasculature (Kottke et al. 2010). However, only CD8+ and NK-cells but not CD4+ cells were
recently reported to be necessary for VSV immunovirotherapy in a mouse melanoma model
where the tumor antigen expressed from VSV was combined with adoptive cell transfer
(Rommelfanger et al. 2012). Similarly, CD8+ T-cells, NK-cells and neutrophils but not CD4+
T-cells were necessary when an immunostimulatory agonistic anti-4-1BB antibody was
combined with vaccinia virotherapy in the murine AT-3 breast carcinoma model (John et al.
2012).
In a recent study, it was noted that in addition to the CD8+ T-cells that were induced
against TAAs via epitope spreading, also CD8+ cells against the oncolytic HSV-1 viral
epitopes were important for tumor control in a murine breast cancer model (Sobol et al.
2011). By, using tumor antigen transgenic and thus tolerized mice, adaptive antiviral
immune responses were reported to be therapeutically important together with adaptive
responses elicited against tumor antigen (Sobol et al. 2011). However, while the authors
showed that some level of antitumor effects could be achieved also in the tolerized mice,
they failed to provide direct evidence to support their claim for a beneficial in vivo role for
antiviral T-cells as they did not exclude the possibility of the presence of secondary tumor
antigens against which the mice were not tolerized. In line with these results, Gürlevic et al.
34
reported that while p53 and telomerase promoter targeting by oncolytic adenovirus
enhanced tumor specificity, dramatically lowered the off-target liver toxicity and inhibited
adaptive antiviral immune responses, surprisingly the non-targeted control viruses were
equally potent in eliciting antitumor immune-responses (Gurlevik et al. 2010). This report is
one of the few examining the effects of induced antiviral adaptive responses in oncolytic
virotherapy efficacy, and contradicts the concept that off-target replication of oncolytic
viruses would reduce the likelihood of on-target antitumor immune responses. Similarly,
on one occasion robust anti-SFV immune responses were reported to be beneficial instead
of detrimental for the efficacy of the therapy, and pre-immunization of mice with the virus
resulted in tumor inflammation and improved virotherapy efficacy in CT26 tumor model
(Smyth et al. 2005).
The role of IFN- (Naik et al. 2012; Quetglas et al. 2012b) and NK-cells (AlvarezBreckenridge et al. 2012b; Bhat and Rommelaere 2013; Cheema et al. 2013; Diaz et al. 2007)
is context dependent as they can potentially either augment (via an antitumor immune
response), hamper (via an antiviral immune response) or be irrelevant/counterbalancing to
the efficacy of oncolytic virotherapy. Type I IFNs are cytotoxic but their direct therapeutic
effect against cancer cells can be suppressed as a result of the acquired apoptosis resistance
common for malignancies. Intriguingly, Beug et al. demonstrated very recently that
oncolytic VSV or viral RNA/DNA mimetics induced durable cures both in an
immunocompetent and xenograft mouse tumor models mediated by IFN- response, but
the therapeutic efficacy required pharmacological suppression of inhibitor of apoptosis
(IAP) proteins (Beug et al. 2014). Not much is known about the role of NKT-cells, T cells,
CD4+ or CD8+ T-regs or MDSCs in the context of OV. Unexpectedly, in the study of Diaz
and colleagues, depletion of T-regs was reported to hamper VSV mediated oncolytic
virotherapy as it stimulated immunity against the therapeutic virus (Diaz et al. 2007).
Similar results were obtained when oncolytic virus infection was combined with DC
vaccination. T-reg depletion resulted in an antibody response against the virus and,
furthermore, to a compensatory induction of MDSCs (Woller et al. 2011). Both direct and
indirect evidence exist that macrophages/microglia are often detrimental to virotherapy,
presumably because they terminate the infection too early by host cell phagocytosis (Fulci
et al. 2006; Fulci et al. 2007; Lun et al. 2010; Spurrell et al. 2014), but also by secreting
antiviral IFN- (Liu et al. 2013). However, macrophages are beneficial at least in certain
antibody therapies where they, in addition to directly phagocytosing opsonized cancer
cells, also prime anti-tumor T-cell responses (Tseng et al. 2013). Type I and type III IFNs
have been recently proven to be crucial for the efficacy of oncolytic VSV against B16
melanomas (Wongthida et al. 2011). Oncolytic virotherapy has also been applied
successfully in combination with rapamycin, cyclophosphamide (CPA) and different
HDAC inhibitors to reduce tumor type I IFN signaling, and some of these combinations
simultaneously stimulated adaptive antitumor immune responses (see chapter “2.4.3
Combination therapy to improve OV efficacy”).
Despite the reported positive contributions of adaptive antiviral responses, there is
however a high amount of evidence suggesting that heterologous prime-boost tumor
vaccination strategies, where the adjuvant virus is changed or replaced with a plasmid
vector, while the virus-encoding tumor antigen is kept the same, work better than
35
homologous vaccinations with a single virus (Bartlett et al. 2013). This was the case in a
study where Vaccinia virus was combined with SFV, with or without the expression of
model tumor antigens, and in a second study where adenovirus was combined with
vaccinia in Syrian hamster model (Tysome et al. 2012; Zhang et al. 2010). Interestingly, in
the latter study, the order of viruses applied was crucial and injection of adenovirus before
vaccinia was a superior strategy over the reverse sequence, and resulted in T-celldependent eradication of the tumors. Importantly, these heterologous regimes did not
reduce the amounts of Nabs or total antibodies as compared to homologous prime-boost
regimen. In another investigation, an SFV-replicon encoding natural P1A tumor antigen
worked well together with adenovirus in heterologous prime-boost setting conferring
protection against murine P815 mastocytoma. SFV worked best when it was used for
priming. Interestingly, SFV-replicon proved to be more powerful in the induction of
antitumor immunity than adenovirus-vectors upon homologous prime-boost vaccination,
and while the quantity of anti-tumor CD8+ T-cells was higher after adenovirus vaccination,
the quality was better with SFV (Naslund et al. 2007). While in the homologous prime-boost
regimens the vector neutralization probably plays a role resulting in a reduced efficacy in
some models and with certain viruses, it seems clear that xenogenic viral antigens can also
easily become immunodominant in a homologous setting especially when weak tumorassociated self antigens are targeted. Indeed, Bridle and colleagues used the mouse B16
melanoma model with oncolytic adenovirus and VSV carrying natural TAAs and
demonstrated that VSV as a single agent biased CD4+ and CD8+ T-cell response against
viral antigens. However, the use of heterologous prime-boost strategy directed the response
towards the tumor antigen resulting in impressive treatment efficacy in the difficult
metastatic setting (Bridle et al. 2009). Interestingly, the investigators used dopachrome
tautomerase of human origin (hDCT) as the target antigen, further supporting the concept
of using “altered self” antigens in immunotherapy. However, the use of xenogenic antigens
to disrupt immunotolerance inherently carries the risk that the resulting peptides presented
on MHCs will be immunologically too different to induce efficient cross-protection. Thus
the cross-immunogenicity of each individual antigen needs to be vigorously validated
before advancing to clinical trials. Although heterologous prime-boost strategies seem to
work better than homologous approaches when they are compared side-by-side, sequential
infections with heterologous, natural pathogens have been also shown to be
immunosuppressive by impairing memory T-cell function (Welsh et al. 2012).
Xenogenic human tyrosinase-related protein-1 (hTRP-1) expressed from regular DNA
plasmid protected mice against B16 melanoma in homologous vaccination regime better
than the mouse equivalent in two separate studies (Leitner et al. 2003; Weber et al. 1998).
Intriguingly, in the latter study, also mTRP-1 conferred protection when it was expressed
from Sindbis replicon instead of a plasmid, a result that was also obtained independently
by another group using alphavirus VEE replicon (Goldberg et al. 2005). As a side note,
Leitner et al., utilized a “layered” technology where the Sindbis replicon was expressed
under a CMV promoter from a DNA plasmid, which was then delivered by gene gun. The
layered plasmid-replicon technology probably increases the safety of the vector, as the
emergence of fully replicative viruses becomes impossible when alphavirus structural
proteins are not needed in the production process. mTRP-1 vaccination was however
efficient when expressed from the plasmid in un-tolerized TRP-1 double knock-out mice,
36
elaborating that normal mice displayed tolerance against this tumor-associated self antigen
and alphaviral vectors were superior over conventional plasmids in unmasking the
immune response against it. Curiously, of the four plasmid/replicon combinations
expressing mTRP-1/hTRP1 only the conventional plasmid expressing hTRP1 induced an
autoimmune disease, vitiligo. This was not, however, due to the xenogenic nature of the
antigen, as vaccinia virus bearing the mTRP-1 insert, while protective against B16, induced
the same autoimmune symptoms. Leitner and colleagues used RNase L knock-out mice to
demonstrate that the better immunogenicity of the replicon compared to a plasmid was not
related to higher antigen expression but, the cell autonomous antiviral response induced by
the viral “danger signals”, such as dsRNA (Leitner et al. 2003). Virus derived PAMPs can
thus work as an adjuvant enhancing immune responses in a dual manner: on one hand by
directly engaging with PRRs of professional APCs (Mellman, Coukos, Dranoff 2011) and on
the other hand by inducing an immunogenic type I IFN response in situ. This response
causes not only ICD but also induction of MHC-I expression and stimulation of dendritic
cells, NK-cells and T-cells (Hervas-Stubbs et al. 2011). As the knockout of the prototypical
apoptotic effector RNase L had such a marked effect on the immunogenicity of Sindbis
vector (Leitner et al. 2003), this suggests that the most important of the many
immunostimulative effects mediated by type I IFN response after alphavirus infection
might be the induction of ICD. This hypothesis was later further supported by Leitner and
colleagues reporting that Sindbis vector expressing a melanoma antigen was less
immunogenic when transfected together with anti-apoptotic Bcl-X(L) even though the
expression of the antigen (and antibodies against it) was increased (Leitner et al. 2004).
Curiously, another study using the same gene gun delivery approach obtained opposite
results, and incorporation of Bcl-X(L) improved the immunogenicity of SFV replicon
vaccine against HPV-transduced TC-1 tumors (Kim et al. 2004). One interesting, though
purely speculative hypothesis for this discrepancy, is that the primary dermal target cells
for the TAA expression in the latter study could have been DCs, and thus the longer
survival of these cells enabled enhanced antigen presentation in the lymph nodes.
Subsequent to these studies, the importance of cell autonomous immunity and especially
type I IFNs for alphavirus mediated antitumor immune response has been corroborated by
Leitner and others (Leitner et al. 2006; Ljungberg et al. 2007). In type I IFN receptor
knockout mice the T-cell mediated efficacy of alphavirus replicon-based tumor vaccination
was abolished, simultaneously causing an increase in Th2 type antibodies against the TAAs
(Leitner et al. 2006). Finally, one recent study did not find any correlation between type I
IFN induction and immunogenicity of four different alphavirus replicons (Maruggi et al.
2013).
In addition to apoptosis, type I IFN signaling has been recently coupled with the induction
of autophagy in several cancer cell lines (Schmeisser, Bekisz, Zoon 2014). The oncolytic
measles virus was also found to induce selective autophagy to mitigate the cytosolic PRR
response in cancer cells (Xia et al. 2014). Intriguingly, in one report, the antitumor efficacy
of SFV replicon with ectopic HPV antigen expression depended on the route of
immunization, with fewer viruses being needed when i.v. or i.m. routes were used
compared to i.p. or s.c. injections (Daemen et al. 2004). It would have been informative to
add an intradermal site to the comparison, as it has occasionally been reported to be more
immunogenic than subcutaneous route (Bonnotte et al. 2003). However, in the phase I
37
clinical trial of CEA encoding vaccinia against metastatic adenocarcinoma both routes had
equal safety and immunogenicities (Conry et al. 1999). The site of initial inoculation is
presumably less important for replicative viruses assuming that the tumor antigen is stable
and remains in the progeny viruses. Alphaviruses have also been used with some
promising preclinical results in vaccination against TAAs of mastocytoma, mammary and
prostate tumors, as well as against the antigens of oncogenic viruses. These approaches
have been reviewed in detail elsewhere (Quetglas et al. 2010).
2.4.3 Combination therapy to improve OV efficacy
Humoral innate- (IFNs, complement, TNF) and adaptive (antibodies) immune responses
can seriously hamper the efficacy of virotherapy by clearing the therapy virus too rapidly
thus preventing successful infection of tumor tissue. Pre-existing or vehicle-induced
antibodies can prevent the virus from even reaching the tumor cells (Kirn, Martuza,
Zwiebel 2001). IFN- plays a crucial role in restricting the replication of SFV and other
alphaviruses (Fragkoudis et al. 2007; Grieder and Vogel 1999; Ryman et al. 2000) and thus
only tumors that are defective in this response can be efficiently targeted. In viral infection,
IFN- can be produced systemically by pDCs (Cao et al. 2008) and locally by all other
nucleated cell types such as, the CNS cells, tumor infiltrating leukocytes and even by the
cancer cells themselves. Some tumors have acquired mutations that render them incapable
of producing and/or responding to IFN-, thus enabling tumor specific virus replication,
but even in those cases, neutralizing antibodies (Nabs) remain a potential hurdle
(McFadden et al. 2009; Zemp et al. 2010). A large variety of different pharmaceuticals has
been tested in combination with oncolytic viruses with the intention of lowering the innateand adaptive immune responses against the virus and/or stimulating the immunity against
the tumors.
Cyclophosphamide (CPA) has been used in different immunocompetent tumor models in
combination with oncolytic HSV (Fulci et al. 2006; Ikeda et al. 1999; Ikeda et al. 2000;
Wakimoto et al. 2004), reovirus (Qiao et al. 2008), vaccinia (Lun et al. 2009) and VSV
(Willmon et al. 2011). The immunomodulatory effects of CPA are pleiotropic (Bartlett et al.
2013). CPA can aid virus replication by reducing neutralizing antibody levels (Ikeda et al.
1999; Qiao et al. 2008) and inhibiting innate immune responses such as antiviral cytokines,
complement and phagocytes (Fulci et al. 2006; Ikeda et al. 1999; Ikeda et al. 2000; Qiao et al.
2008; Wakimoto et al. 2004). Furthermore, low dose CPA can promote anti-tumor immunity
by depleting regulatory T-cells, stimulating memory T-cell proliferation and sensitizing
tumors to TRAIL-dependent CD8+ T-cell-mediated tumor destruction (Koski et al. 2010;
Schiavoni et al. 2000; van der Most et al. 2009). Proliferation of memory T-cells may be, at
least in part, a result of the induction of type I IFN, and upregulation of IFN- production
as a response to CPA treatment has also been reported by other groups (van der Most et al.
2009). Controversially, CPA has also been found to lower the production of antiviral
cytokines, including IFN- (Wakimoto et al. 2004). Thus, the cytokine milieu related
effects of CPA remain elusive and may depend on the dosage and timing. Importantly, the
oncolytic viruses themselves exert complex effects on immunity and occasionally the
introduction of CPA does not produce synergistic but antagonistic effects on the
38
therapeutic efficacy (Willmon et al. 2011). Remarkably, the anticancer immunotherapeutic
effects of CPA have been recently reported to rely on Gram-positive intestinal bacteria
which translocate into secondary lymphoid organs in response to CPA (Viaud et al. 2013).
Along with CPA, rapamycin is another immuno-regulator commonly used in combination
with OV. Rapamycin has been shown to suppress both tumor and host IFN- responses,
thus in optimal circumstances allowing more efficient oncolytic virus replication and tumor
killing (Zemp et al. 2010). Rapamycin has been utilized in combination with multiple
oncolytic viruses such as VSV (Alain et al. 2010), myxoma (Lun et al. 2010; Lun et al. 2007)
and vaccinia (Lun et al. 2010; Lun et al. 2009). Rapamycin was shown to have also direct
antineoplastic effects, and interestingly it has been tested also in glioblastoma clinical trials
(Cloughesy et al. 2008). Rapamycin is an inhibitor of mTORC1 that can stimulate type I IFN
production through phosphorylation of its two target proteins, 4E-BPs and S6K1/2. The
main mechanism of rapamycin synergy with OV is likely related to its capability to
suppress the systemic production of IFN- by pDCs (Cao et al. 2008), which then aids
virus replication in semi-IFN--sensitive tumors (Alain et al. 2010). Valproic acid (VPA)
belongs to the class of histone deacetylase (HDAC) inhibitors and along with CPA and
rapamycin, is also commonly exploited pharmaceutical in combination with OV. HDAC
inhibitors have been shown to act synergistically with OV at least by inhibiting IFNresponses and potentiating the virus-induced cancer cell apoptosis (Nguyen et al. 2008).
VPA was shown to suppress NK-cell mediated early clearance of HSV (AlvarezBreckenridge et al. 2012a). Surprisingly, it was demonstrated that VPA may also boost
antigen targeted immunity while simultaneously decreasing autoimmunity and the
adaptive responses against viral vector (Bridle et al. 2013).
Modulation of VEGF-A signaling in combination with replicative oncolytic VSV and
reoviruses has been reported to be a powerful therapeutic approach in B16, B16ova and
B16-VEGF mouse melanoma models, resulting in long term antitumoral effects even in
B16ova tumors where reovirus replication per se was poor (Kottke et al. 2010). The rationale
behind this approach was to induce a VEGF-A burst in the tumors by administering VEGFA (in the case of VEGF-A non-expressing B16 tumors) or by first administering and then
withdrawing VEGF-A inhibitor (in the case of VEGF-A expressing B16-VEGF tumors). Both
of these strategies caused a rebound in the VEGF-A levels and the induced VEGF-A burst
allowed tumor-associated endothelial cells transiently to support viral replication. This led
to CD4+, CD8+ and NK-cell-mediated immune attack on tumor vasculature resulting in
long-term regressions. Intriguingly, co-administration of VEGF-R/PDGF-R inhibitor
Sunitinib with OV resulted also in synergistic effects, which, however, were discovered to
result from Sunitinib off-target inhibition of both PKR and RNaseL (Jha et al. 2013).
Classical chemotherapeutic and radiotherapeutic approaches have been successfully
combined with OV to achieve synergy, for example by enhancing tumor-specific viral
replication or inducing enhanced antitumor immune responses. Many of these approaches
have been reviewed elsewhere (Ottolino-Perry et al. 2010). Finally, oncolytic viruses have
been successfully combined not only with different pharmaceuticals, but also with other
oncolytic viruses. At least vaccinia and VSV infections seem to act synergistically (Le Boeuf
et al. 2010), while vaccinia and SFV do not have this property (Vaha-Koskela et al. 2013).
39
2.4.4 Targeting of Oncolytic Viruses
The ultimate goal is of oncolytic virotherapy is to develop safe viruses, which after systemic
administration will specifically home to primary and metastatic tumors and destroy them
or their vasculature by lytic replication and/or stimulation of antitumor immunity. There
are multiple strategies involving a plethora of different constructs to achieve tumor
specificity. These strategies include deletion-, transcription-, receptor-, translation-, directed
evolution-, cell carrier- and microRNA -targeting of oncolytic viruses. From each of these a
few representative examples will be described.
Deletion targeting to create conditionally replicating viruses
Viral genes are constantly under evolutionary pressure in the changing environment, and
unlike higher organisms, viruses do not usually carry redundant genetic material inside
their spatially limited virions. Viral genes have evolved to suppress extrinsic and intrinsic
apoptotic pathways (for example, degradation of TNF-receptor or p53) (Benedict and Ware
2001; Querido et al. 2001), disrupt DNA-damage signaling (e.g. MRE11) (Martin and Berk
1999), to prevent innate- (Muster et al. 2004) and adaptive immune responses (Andersson et
al. 1985), to stimulate nucleotide pool synthesis (in non-dividing cells) (McCart et al. 2001),
to induce cellular proliferation by Rb inactivation or direct E2F activation (O'Connor and
Hearing 2000) and to activate growth factor receptors (McCart et al. 2001), translational
machinery (O'Shea et al. 2005) and histone acetyl transferases (HATs) (Howe et al. 1990).
These viral genes are often redundant in cancer cells, which have already acquired
mutations and epigenetic changes that deregulate cell survival, proliferation,
macromolecular synthesis and immune-evasion (O'Shea 2005). Hence, by deleting one or
more of the respective viral genes, one can suppress virus replication in normal healthy
cells without significantly hampering the replication in neoplastic cells. Creating
conditionally replicating alphaviruses by targeted deletions is not considered possible as
the viral genomes are small and most of the viral genes are indispensable for productive
infection.
Transcriptional targeting
A common approach to achieve tumor selective replication of a DNA virus is to place one
or more of the essential viral genes under the control of a tumor-specific promoter or a
tissue-specific promoter overexpressed in the tumor (Dorer and Nettelbeck 2009; Thorne,
Hermiston, Kirn 2005). The examples of tumor- and tissue-specific promoters to target
viruses such as HSV-1 and especially different adenovirus serotypes include albumin or
alpha-fetoprotein promoters for HCC, calponin promoter for sarcoma, probasin and
prostate specific antigen (PSA) promoters for prostate cancer and hypoxia response
elements for multiple different cancers. While transcriptional targeting of oncolytic agents
has shown promise in terms of specificity, the novel promoters are often weaker in
comparison to native viral promoters.
An excellent example of a multifaceted transcriptional targeting approach was reported by
Gürlevic and colleagues, who designed a highly cancer-specific adenovirus Ad-p53T by
40
exploiting two ubiquitous cancer cell characteristics: expression of telomerase and a
suppressed p53 signaling (Gurlevik et al. 2010). As p53 is expressed mostly in noncancerous cells, a viral repressor was placed under the control of the p53 promoter and
furthermore, the autologous adenoviral early genes promoter was replaced with the human
telomerase promoter. As opposed to the unmodified control virus, the engineered construct
exhibited no high virus dosage related hepatotoxicity, and importantly, the promoter
modifications did not compromise the oncolytic efficacy in a syngeneic murine cancer
model. However, as p53 was retrovirally knocked down in the targeted CMT64 small cell
lung cancer cells, the specificity of Ad-p53T remains to be shown in natural tumors. In
addition, radiation inducible promoters such as the survivin-promoter have been
incorporated into oncolytic viruses to achieve tumor specific replication and transgene
expression in conjunction with radiotherapy (Nandi et al. 2008).
As eukaryotic cells do not express RdRp, transcriptional targeting is generally not
applicable to RNA-viruses such as Semliki Forest virus. An elegant way to circumvent this
limitation was described by Guan et al., who engineered a hybrid vector comprising a
helper-dependent adenovirus encapsulated SFV replicon driven by alpha-fetoprotein
promoter and the IL-12 transgene expressed from SFV subgenomic promoter. The construct
was HCC-specific and demonstrated high in vivo efficacy without causing any liver toxicity
(Guan et al. 2006).
Targeting virus entry
In addition to manipulating viral replication, it is possible to achieve retargeting already
one step earlier by modifying the viral cell surface attachment. The viral surface proteins of
one serotype can be replaced by those of another or even with proteins from a completely
different virus, a well established process termed pseudotyping.
Oncolytic viruses which work well in selected tumor cell lines, may fail to target the actual
tumors in clinical trials, because of lack of expression of viral receptors and co-receptors in
the tumors. For instance, adenovirus serotype 5 (Ad5), which has been commonly used in
gene therapy and oncolytic virotherapy for GBM, uses its fiber protein knob domain to
bind Coxsackie virus and adenovirus receptor (CAR) on the cell surface. However, CAR
expression is in general low in tumors and especially in GBM, and this seriously impairs
the oncolytic use of the vector (Kim et al. 2003; Nandi and Lesniak 2009). Several successful
approaches have been undertaken to retarget the adenovirus tropism towards more
prominent epicellular structures in gliomas and other tumors. These include pseudotyping
Ad5 with Ad3 fiber knob to create Ad5/3 chimeras (Zheng et al. 2007), incorporation of the
RGD motif into the knob to target integrins (Dmitriev et al. 1998) and oncogenic FGFR1
receptor targeting by chemical conjugation of FGF2-ligand to the adenovirus capsid (Wang
et al. 2005). One has to bear in mind that chemical conjugation as well as chemical shielding
against immune-clearance using substances such as polyethylene glycol (PEG)
(Arudchandran et al. 1999; Eto et al. 2008) are probably more suitable for viral vector and
replicon mediated uses than for oncolytic virotherapy applications, as these modifications
do not pass after the first infection cycle to the budding progeny virions. Stable, genetic
retargeting to a commonly overexpressed receptor EGFR has been achieved both for an
41
oncolytic adenovirus (van Beusechem et al. 2003) and more recently for an oncolytic HSV
(Uchida et al. 2013).
Recently, as a good example of a rationale genetic engineering approach, measles fusion
protein and single chain antibodies have been exploited to create a novel and, in theory,
universal virus retargeting method. In this approach, the single chain antibody engineered
to virus surface first binds to its specific epitope on cancer cell surface, and then the measles
fusion protein mediates fusion of viral and cellular membranes leading to virus entry. The
measles system has been utilized to retarget a variety viral vectors and replicative viruses
to cancer-associated molecules such as EGFR, folate receptor, prostate membrane-specific
antigen (Ayala-Breton et al. 2012) and CD133 CSC marker (Bach et al. 2013). The single
chain antibody has also been replaced successfully with designed ankyrin repeat protein, in
an effort to retarget lentiviral vector to HER2/neu (Munch et al. 2011).
With respect to alphavirus vectors, the laboratory of Daniel Meruelo has shown the Sindbis
virus replicon to be amenable to retargeting to cancer cells, both by direct genetic
engineering of the envelope towards LH/CG receptors of choriocarcinoma cells (Sawai and
Meruelo 1998) and ligation of IgG-binding domain of protein A to achieve targeting
towards a variety of cell surface molecules when combined with specific monoclonal
antibodies (Iijima et al. 1999; Ohno et al. 1997). The impediment to applying similar
strategies to SFV is its unknown receptor.
Translational targeting
One other, though a relatively uncommon approach to achieve tumor-selective virus
replication is the translational stabilization of viral mRNAs in a cancer-specific manner.
Expression of COX2, a proto-oncogene and an inflammatory mediator has been associated
with poor prognosis in many cancer types (Lee, Myung, Song 2013). COX2 is activated by
cytokines and growth factors via the Ras/p-MAPK pathway and at least part of the
activation can be attributed to stabilization by 3’UTR of its mRNA. Tumor specific
stabilization of an oncolytic adenovirus mRNA was accomplished by ligating the 3’UTR of
COX2 to the adenovirus early gene, E1A mRNA (Ahmed et al. 2003).
Tumor targeting at the level of mRNA stabilization has been achieved also with poliovirus.
Poliovirus has an inherent ability to target many tumors including GBM that are rich in its
receptor Necl-5 (Goetz and Gromeier 2010; Pol et al. 2013). However, the inherent
neuropathogenicity of the virus would normally limit its use in oncolytic virotherapy. As
CNS specific translation of poliovirus non-capped mRNA depends on a cell type-specific
IRES element, its replacement with complementary sequence from human rhinovirus type
2 resulted in abrogation of neurovirulence while replication in malignant cells was retained
(Goetz and Gromeier 2010). According to the preliminary results from an ongoing phase I
clinical trial, poliovirus PVS-RIPO has proven to be very promising in the treatment of
incurable GBM (Pol et al. 2013). While potentially amenable, translational targeting of
oncolytic alphaviruses has not been tested. Some alphaviruses such as EEEV, VEEV and
WEEV have a high replication power, but their therapeutic use is limited because of their
neurovirulence (Jose, Snyder, Kuhn 2009).
42
Targeting with microRNA technology
Different RNAi pathways, especially those involving miroRNAs (miRNAs), are known to
be dysregulated in cancer. Some miRNAs are oncogenic and thus overexpressed in
malignancies, while others have tumor suppressor functions and are thus silenced (Baer,
Claus, Plass 2013). So far, work has concentrated on incorporating target sequences of tissue
specific miRNAs into the genomes of oncolytic viruses. This approach was recently
successfully adapted to prevent adenovirus hepatotoxicity by addition of miR122 target
sites (Ylosmaki et al. 2008; Ylosmaki et al. 2013) and the neuropathogenicity of the virulent
SFV4 strain by including miR124 targets in the viral genome (Ylosmaki et al. 2013). In
addition, miRNA detargeting of alphaviral replicon has been reported (Kamrud et al. 2010).
SFV and other alphavirus vectors might be especially suitable for miRNA based targeting
since RNAi is an integral part of antiviral defense against SFV infections in mosquito cells
(Siu et al. 2011). Remarkably, high tumor specificity was recently reported by Sugio and
colleagues for telomerase promoter-driven adenovirus harboring different target sequences
for miRNAs that were under expressed in cancer cells without compromising oncolytic
efficacy (Sugio et al. 2011). In this approach, liver detargeting was also accomplished by
addition of miR-122a target sequence to the virus genome.
Targeting with cell carriers
Solid tumors are complex organs that attract several types of non-neoplastic cells to their
stroma such as myeloid derived stem/progenitor cells and many types of immune cells
(Hanahan and Weinberg 2011). Several laboratories have taken advantage of these tumor
stromal cells as carriers for oncolytic viruses by utilizing their inherent tumor homing
capabilities and the protection which they provide against antiviral immunity in the
circulation. Recent examples of successful application of this approach to improve OV
include CIK loaded and CCL5 expressing vaccinia virus utilized in murine MC38 colorectal
cancer model (Sampath et al. 2013), neural stem cells loaded with adenovirus in the
treatment of murine glioma xenografts (Kim et al. 2013), VSV loaded MDSCs in a metastatic
MCA26 murine colon carcinoma model (Eisenstein et al. 2013) and hypoxia promoterdriven-adenovirus delivered by macrophages in a murine prostate cancer xenograft model
(Muthana et al. 2013).
Targeting by directed evolution
A novel and innovative oncolytic virus targeting and potentiating approach was
introduced by Hermiston laboratory, which used a “directed evolution” strategy. Pools of
different adenovirus serotypes were passaged in human tumor cell lines (colon, breast,
pancreatic, prostate) to enable viral recombination and then the most potent variants were
selected after subsequent rounds of limiting dilution and repassaging. The resulting virus
ColoAd1, a Ad3/Ad11p virus chimera was found to be 2–3 log units more potent and colon
carcinoma selective than ONYX-015, one of the clinically most advanced oncolytic
adenoviruses (Kuhn et al. 2008).
43
2.4.5 Naturally targeted oncolytic viruses
Cancers, per definition, consist of malignant cells that show an abnormally high rate of
proliferation and capability to resist cell death. In order to drive unregulated proliferation,
cancer cells induce overactive biosynthesis of different macromolecules such as nucleic
acids. Since viruses also need these building blocks for the synthesis of progeny particles it
is not surprising that metabolically active malignant cells in general support viral
replication better than quiescent healthy cells (Le Boeuf et al. 2013; Parato et al. 2012). While
most oncolytic viruses used to date have required manipulation in order to achieve a
sufficiently cancer-specific infection, a number of essentially unmodified viruses have also
been used to treat cancer. These viruses inherently replicate better in transformed cells and
thus can be defined as “naturally targeted oncolytic viruses”.
Receptor preference
Several tumor types have been reported to overexpress poliovirus receptor Necl-5, but due
to the well-known neurovirulence of poliovirus additional modifications are necessary and
thus poliovirus cannot be considered as a naturally targeted oncolytic virus (Goetz and
Gromeier 2010). Sindbis virus uses high-affinity laminin receptor (LAMR) for its entry into
mammalian cells, and the receptor has a known role in tumorigenesis. The unoccupied
form of the receptor has been reported to be over-expressed in numerous cancer types and
thus for long it was thought to account for tumor preference of Sindbis virus (Tseng et al.
2004). This view was, however, recently contradicted by Huang and colleagues, who
claimed that cancer defects in type I IFN response rather than the levels of viral receptor, is
the primary determinant of Sindbis tumor tropism (Huang, Guo, Hwang 2012). While the
relative importance of these two factors might be disputable, it seems conceivable that both
the number of Sindbis receptors on cancer cells and the intactness of cell autonomous
antiviral responses contribute to the tumor infectivity, albeit at different stages.
Defects in apoptosis signaling of cancer
The fact that most tumors have defects in their apoptosis mechanisms makes them also
sensitive to some viruses, as programmed cell death is a fundamental part of antiviral
defense. This seems to hold true at least for NDV. The replication of NDV was documented
to occur preferentially in malignant cells with defective apoptosis responses, and
interestingly, tumor targeting did not seem to depend on type I IFN defects (Mansour,
Palese, Zamarin 2011). However, some controversy regarding the tumor specificity
remains, as NDV is not insensitive to the effects of IFN (Buijs et al. 2014). Interestingly, in
this very recent study by Buijs et al., 11 different human pancreatic adenocarcinoma cell
lines were tested for their IFN signaling and while major differences were observed in IFN
induction upon NDV infection, IFN pretreatment inhibited replication of the virus in all
tested cell lines.
Defects in neoplastic type I IFN signaling
Since the IFN system and ISGs have an important role in growth restriction and apoptosis,
the responsible genes are occasionally deleted or epigenetically silenced in tumors. For
example, IRF1, a transcription factor important in immunity against several viruses,
including alphaviruses (Karki et al. 2012; Schoggins et al. 2011) is a well-known tumor
44
suppressor acting in concert with p53 (Tanaka et al. 1996). As a secondary effect, silencing
of one or more of these genes can compromise tumor antiviral defense partially or even
completely, rendering them as vulnerable targets for IFN sensitive oncolytic viruses such as
VSV (Naik et al. 2012), myxomavirus (Lun et al. 2010) and reovirus (Wilcox et al. 2001).
Interestingly, simple overexpression of one of the IFNA-receptor chains, IFNAR2c, can
reintroduce the antiproliferative effects of type I IFN to many cancer cell lines (Wagner et
al. 2004). According to one recent study, a large number of type I IFN receptors was needed
for the antiproliferative effects of type I IFN while clearly smaller number was adequate for
the antiviral response (Levin, Harari, Schreiber 2011). Different replicative alphaviruses
especially those based on Semliki Forest- and Sindbis viruses have been used as naturally
targeted oncolytic viruses. Sindbis replication preference in tumors has been reported to
result from defects some cancers have in their type I IFN production and responsiveness
(Huang, Guo, Hwang 2012). It is understandable that while defects in type I IFN
production of cancer cell lines might impair the antiviral defense in vitro, they do not
necessarily guarantee successful tumor targeting in vivo, as several tumor-resident nontransformed cell types such as macrophages are potent type I IFN sources (Liu et al. 2013).
Hence, tumors that have defects in their type I IFN response are better candidates for the
naturally targeted oncolytic viruses. Interestingly, U87 cells seem to have defects both in
the response and induction of type I IFN (Wollmann, Robek, van den Pol 2007).
There are multiple pathways which may lead to aberrant antiviral signaling in cancer.
Suppressors of cytokine signaling (SOCS) proteins are negative-feedback regulators of type
I IFN signaling. SOCS1 is a well-known tumor suppressor in the SOCS family of proteins,
which is silenced by hypermethylation in many cancers, but intriguingly, induced in
melanomas (Davey, Heath, Starr 2006). Melanoma is responsive and thus commonly
chosen target malignancy for oncolytic virotherapy in clinical trials (Mastrangelo et al. 1999;
Russell, Peng, Bell 2012). This may be at least partly due to the known immunogenicity of
melanomas, a fact supporting immuno-virotherapeutic approaches. In addition, tumors do
also inherently differ in their potency of launching cell autonomous antiviral attack. Indeed,
some melanomas escape interferon therapy by suppressing the expression of STAT1
(Pansky et al. 2000; Wong et al. 1997; Wong et al. 1998). While the role of reduced STAT1
expression in melanoma IFN resistance has been contradicted (Chawla-Sarkar et al. 2002) it
is nevertheless clear that melanomas do develop resistance to IFN therapy and, hence, the
escape variants could be attractive targets for type I IFN sensitive oncolytic viruses. In
addition to melanomas, also other tumors deregulate type I interferon response in order to
evade adaptive antitumor immunity, and this seems to be true at least for xenoantigenic
and thus for immunogenic human papillomavirus (HPV) induced tumors. HPV encodes
several proteins that inhibit the interferon response, apoptosis and MHC presentation, all of
which serve both the virus and tumor survival, and inadvertently, make the tumors good
targets for oncolytic virotherapy (Grabowska and Riemer 2012; Le Boeuf et al. 2012). Cells
infected with Epstein-Barr virus, the etiologic agent in Hodgkins’- and Burkitt's lymphomas
and in some nasopharyngeal carcinomas, have also been reported to be resistant to type I
IFN induced apoptosis (Ruf et al. 2005; Wu et al. 2009). Strikingly, serious IFN response
defects have also been found in a number of human lung adenocarcinomas and prostate
cancers (Dunn, Koebel, Schreiber 2006). Some human prostate tumors have been found to
45
be extremely sensitive to VA7-mediated oncolysis resulting in a total eradication of
implanted xenografts (Miika Martikainen, personal communication).
STAT1 induction has been recently implicated in the development of tumor resistance to
genotoxic stress (radiation and chemotherapy), and radioresistant tumors concomitantly
acquired resistance to the type I IFN sensitive HSV-1 mutant. Hence, the antiviral response
profile of heavily pretreated tumors, as is the case in most oncolytic virotherapy trials, may
be inherently unfavorable, proposing a paradigm shift in favor of selecting patients with
malignant but early stage tumors for entry into clinical trials (Khodarev et al. 2004;
Khodarev, Roizman, Weichselbaum 2012). Since overactive IFN/STAT pathway has also
been implicated in inflammation-mediated tumorigenesis of gastric cancer, hepatocellular
carcinoma and colorectal carcinoma, the patient selection for oncolytic virotherapy trials,
especially with type I IFN sensitive viruses, should be based on careful diagnostic screening
for tumor antiviral responses, and herein some tumor types might be inherently more
sensitive to virus replication. An ideal screening approach (but applicable only for small
scale trials) would be based on functional screenings, which always measure the objective
outcome (cancer cell death) regardless of the underlying genotype. These screens could
involve ex vivo infections of live tumor slice cultures or freshly prepared primary cells
pretreated with type I IFN. Nevertheless, for the purposes of large scale personalized
medicine, high throughput screens of applicable tumor resistance markers need to be
developed. As an example, constitutive high level expression of OAS and MxA ISGs has
been proposed as good candidate marker for tumor resistance against a type I IFN sensitive
oncolytic VSV (Moerdyk-Schauwecker et al. 2013).
Most viruses are sensitive to type I IFN and at least HSV, reovirus, VSV, poliovirus,
influenza- and adenoviruses replicate more efficiently in constitutively Ras-activated
tumors that often have defects both in the induction and response to type I IFNs (Christian
et al. 2012; Donnelly et al. 2013a; Park et al. 2010). Nevertheless, sometimes the defects are
virus-specific, and activation of the same CUG2-Ras pathway that mediated the
susceptibility to reovirus also conferred resistance to another type I IFN sensitive virus,
VSV through STAT1-OASL2 mediated signaling (Malilas et al. 2013). This apparent
discrepancy may stem from differences in the antiviral effects of OASL2 against reovirus
and VSV. At least the signaling pathways leading to the defects differ, and although the
constitutively activated Ras-MEK pathway is considered to be responsible for the IFN
defects facilitating VSV replication (Noser et al. 2007) Ras-p38 signaling mediates these
defects for reovirus (Park et al. 2010). Interestingly, the tumor specificity of parvovirus
(Paglino, Andres, van den Pol 2014), myxomavirus (Zemp et al. 2013) and possibly NDV
(Mansour, Palese, Zamarin 2011) seem to be type I IFN independent.
It has become evident that established cancer cell lines vary in their permissiveness to
oncolytic viruses. This permissiveness may depend particularly on cancer cell autonomous
IFN- antiviral defense capacity, especially when unmodified viruses such as VA7 are
used. However, in contrast to clonal cell lines, naturally occurring cancer displays
significant intratumoral and patient-to-patient heterogeneity. Thus, while it has been
occasionally possible to dissect the mechanisms responsible for the diminished antiviral
defenses in defined cell lines, reliable extrapolation of these findings to the clinical setting
46
has proved challenging. Furthermore, while it has been recently estimated in Bell’s
laboratory that 65-70% of tumors are unable to produce or respond to type I IFN (Forbes et
al. 2013), occasionally the pre-treatment experiments have been carried out using lower
affinity IFN- instead of the de facto early interferon IFN- (Stojdl et al. 2000) or
concentrations of type I IFN as low as 5 U/ml (Stojdl et al. 2003).
2.4.6 VA7 oncolytic virus
SFV is an Alphavirus with a high replication efficiency and broad host range. Fully
replication competent VA7 virus is based on the apathogenic and neurotropic SFV strain
A7(74) and the virus harbors a multiple cloning site for expression of ectopic therapeutic
and immunostimulative genes (Vaha-Koskela et al. 2003). Despite the high mutational rate
of RNA-viruses during replication, the multiple cloning site of VA7 was shown to be rather
stable, and 60% of plaques obtained from CNS homogenates after VA7-EGFP infection
were fluorescent (Vaha-Koskela et al. 2003). VA7 has proven to be safe in
immunocompetent hosts and powerful in several xenograft mouse models such as
orthotopic human A2058 melanoma (Vaha-Koskela et al. 2006), K7M3 osteosarcoma (Ketola
et al. 2008) and A549 lung cancer (Maatta et al. 2007; Maatta et al. 2008). VA7 has been
shown to be capable of homing to the tumors even in operatively difficult body locations
such as bone (Ketola et al. 2008). Since high viremia is an integral part of the lifecycle of
arthropod borne alphaviruses such as SFV, they are likely especially suitable for targeting
widely-spread metastatic disease. Furthermore, inherently neurotropic SFV is also well
suited for targeting primary and secondary brain tumors. SFV is not endemic in the
Western world, reducing the likelihood of pre-existing neutralizing immunity against the
vector (Mathiot et al. 1990).
Subcutaneous A2058 human melanoma xenografts were very sensitive to VA7 irrespective
of the route of administration (i.v., i.p., or i.t.), but the total eradication of tumors was not
achieved due to the presence of physical barriers caused by tumor compartmentalization
and the emergence of virus resistant cancer cell clones (Vaha-Koskela et al. 2006). The type I
IFN responsiveness of the A2058 tumors was not analyzed in the study. VA7 has
demonstrated in vivo efficacy also in certain prostate tumor xenografts, which according to
preliminary results also seem to have defects in their type I IFN response (Miika
Martikainen, personal communication). However, testing the VA7 safety and efficacy in
immunocompetent animal models is necessary as immunocompromised xenograft models
do not exhibit the tumor microenvironment highly populated with lymphocytes found in
many human tumors (Zemp et al. 2010) and also lack the components of antiviral and
antitumoral adaptive immunity.
SFV-based replicative and non-replicative (replicon) vectors have been utilized widely in
oncolytic immunovirotherapy (Quetglas et al. 2010). However, both local and systemic
administration of VA7 vector elicited MxA protein activation in subcutaneous A549
xenografts, indicative of IFN- induced responses against the virus. This likely
contributed to the poor result with the systemic VA7 delivery as only after intratumoral
47
injections large virus-positive areas were visible in immunohistochemistry 16 hours post
infection (Maatta et al. 2007). In line with these results, systemic application of VA7 did not
improve survival in the orthotopic A549 lung cancer model, whereas local VA7 injection
prolonged the survival of mice significantly (Maatta et al. 2008).
2.4.7 Armed oncolytic viruses
A variety of different transgenes have been incorporated into oncolytic viruses to boost the
efficacy of virotherapy, to allow noninvasive imaging and occasionally also to improve
safety of the treatment. As previously described, some of these transgenes are designed to
stimulate antitumor immunity (see chapter “2.4.2 Oncolytic virotherapy overlaps with gene
therapy and immunotherapy of cancer – emerging concept of oncolytic
immunovirotherapy”) (Naik et al. 2012). Other pre-clinical and clinical strategies include
incorporation of cytosine deaminase/thymidine kinase double suicide transgenes into
replicative adenovirus to treat prostate cancer (Freytag et al. 2002), adenoviruses releasing
anti-angiogenic (Thorne et al. 2006) or ECM degrading (Kim et al. 2006) factors for tumor
vasculature targeting and enhanced penetration respectively, sodium iodine symporter
(NIS) expressing adeno-, VSV- and measles viruses combined with radioiodide for
treatment and/or SPECT/CT imaging of several cancer types (Naik et al. 2012; Opyrchal et
al. 2012; Trujillo et al. 2013). Interestingly, both matrix metalloproteinase (MMP) (Schafer et
al. 2012) and MMP-inhibitor tissue inhibitor of metalloproteinases (TIMP) (Yang et al. 2011)
expressing viruses have exhibited enhanced oncolytic efficacy. The former was thought to
be a result of improved viral penetration thus resembling hyaluronidase co-treatment
(Ganesh et al. 2008) and the latter due to direct tumor suppression. It will be interesting to
determine in the future, whether TIMP expression from oncolytic viruses works partly
through immuno-stimulation, as MMP activity has been linked to immunosuppression via
stress ligand cleavage (Tang et al. 2013). A novel systems biology approach was recently
introduced in the Bell laboratory, which created oncolytic viruses expressing type I IFN
soluble receptor antagonists based on mathematical modeling (Le Boeuf et al. 2013). The
antagonists were produced late in the viral life cycle, so the initial tumor specificity relied
on the presence of a hospitable macromolecular environment in the metabolically hyperactive tumors. This approach was shown to work in vitro and a similar approach has been
demonstrated to enhance the in vivo efficacy of oncolytic HSV (Fu et al. 2012). Finally,
viruses engineered to express somatostatin (McCart et al. 2004) have been used to enable
positron emission tomography (PET) visualization, luciferase and enhanced green
fluorescence protein (EGFP) (Msaouel et al. 2009) to enable optical imaging, and the
transferrin receptor (Weissleder et al. 2000) and ferritin (Iordanova et al. 2013) to the
purpose of magnetic resonance imaging (MRI) of the infection.
48
2.4.8 Safety and ethics of oncolytic virotherapy
A wide variety of different oncolytic viruses have been tested in clinical trials and the safety
profile, even with the highest viral doses, has been excellent in comparison with classical
cancer therapeutics. Nevertheless, careful preclinical and phase I clinical safety assessment
are necessary for each new construct. Since viruses typically have restricted species tropism
and thus the infection can potentially cause dramatically differing effects in different hosts,
pre-clinical safety assessments should be carried out using more than one animal species.
Other safety factors which need to be considered include the potential toxicity of the virustransgene combination, viral evolution, risk of horizontal transmission and the prevalence
of antiviral medication, vaccinations and pre-existing immunity in the population (Miest
and Cattaneo 2014; Russell, Peng, Bell 2012).
The demand for careful and impartial biosafety assessments has been spurred especially by
two cases where gene therapy resulted in patient fatalities and serious harm. The first case
involved an 18 years old patient suffering from a rare genetic disease ornithine
transcarbamylase deficiency resulting in defective protein catabolism and accumulation of
toxic levels of ammonia. On September 17th 1999, he participated in a gene therapy trial,
where the defective enzyme was to be introduced into his liver by non-replicative
adenovirus 5-mediated gene transfer. The patient died from systemic inflammatory
response syndrome, “cytokine storm”, resulting in multiple organ failure 98h after hepatic
artery infusion of the vector (Raper et al. 2003). While all law suits were settled quietly, the
investigators were accused of failing to notify the patient about the deaths that had
occurred in experimental animals, the serious side effects that other patients receiving the
same treatment had experienced and the extent of financial and academic conflict of
interest of the principal investigators in the study. Furthermore, the study protocol was
violated and the patient had received a higher than approved dose of the vector, and was
not excluded from the trial even though his ammonia levels were above the accepted
threshold (Wilson 2010). Most of the viral vectors and oncolytic viruses induce potent
innate and adaptive immune responses, and these responses may be exacerbated if, as in
this case, the patient presents acquired immunity against the vector already prior to
treatment, or if the administration protocol is long enough for these memory responses to
develop. The trial in question was phase I dose escalation study, and the patient was
receiving the highest, ultimately fatal dose of the non-replicative vector. The dosing of
replicative oncolytic viruses is likely to increase as viral manufacturing technologies
develop thus raising new safety concerns, which will need to be addressed (Russell, Peng,
Bell 2012). Furthermore, careful regulation of the levels of replicative viruses is deemed to
be more challenging than dosing of classical pharmaceuticals or non-replicative vectors,
and it is conceivable that the original input dose may not always correlate linearly with the
level of the resulting viremia. Recently, high levels of mortality and morbidity of
experimental mice due to cytokine storm were reported when oncolytic VSV was combined
with the TLR agonist LPS (Rommelfanger et al. 2013).
A second well documented case of gene therapy-related major adverse events emerged
from two independent clinical trials aimed to treat X-linked severe-combined
immunodeficiency (SCID-X1) by gamma (onco)retroviral transfer of the missing IL-2
49
receptor -gene (Hacein-Bey-Abina et al. 2008; Howe et al. 2008). While the gene therapy
could cure most of the 20 infants suffering from the debilitating genetic disease, five of the
patients developed leukemia, of which four were chemotherapeutically treatable (Wang et
al. 2010a). The trigger for the multistep oncogenesis was proposed to be the activation of
LMO2 gene due to inherent enhancer activity originating from the LTR-regions of the
murine leukemia virus (MLV) gene transfer vector, which was stably integrated upstream
of the proto-oncogene. To reduce the likelihood of insertional oncogenesis, in the thirdgeneration of self-inactivating lentiviral vectors the viral LTR regions are designed to be
automatically inactivated; lentiviral vectors also have potentially safer insertional profile
than oncoretroviral vectors (Rothe, Modlich, Schambach 2013; Schambach et al. 2013). The
concerns of insertional oncogenesis may be especially prevalent when replicative oncolytic
viruses capable of genomic integration such as MLV are used, and the risks have to be
carefully balanced against the potential benefits achieved in the treatment of pre-existing
malignancies (Perez et al. 2012).
Some immunomodulatory cytokines such as interleukin-4 (IL-4) seem to be exceptionally
dangerous when expressed from oncolytic viruses, resulting in uncontrollable virus
replication (Chen et al. 2011; Jackson et al. 2001). These recombinant viruses have to be used
with an appropriate biosafety containment level, and the cross-species reactivity of the used
transgenes has to be carefully studied to limit the risk for human population. The ectopic
inclusion of one of the multiple viral genes targeted against IFN response (Devasthanam
2014) to create recombinant viruses with improved oncolytic efficacy poses obvious safety
concerns, but surprisingly so far the safety of the published constructs has been good (Fu et
al. 2012; Zamarin et al. 2009). Combining SFV with certain Th1 type immunosuppressants
such as linomide (Peltoniemi et al. 2002), CPA (Bradish, Allner, Fitzgeorge 1975),
aurothiolates (Scallan and Fazakerley 1999), cyclosporine (Fazakerley and Webb 1987) and
tunicamycin (Singh et al. 1987) has been reported to increase neurovirulence by presently
incompletely understood mechanisms. Low dose CPA is a commonly used additive in OV
preclinical and clinical regimes (see chapter “2.4.3 Combination therapy to improve OV
efficacy”). While several mechanisms such as depletion of antigen presenting cells
(Peltoniemi et al. 2002) and induction of neuronal membrane synthesis (Fazakerley 2002;
Scallan and Fazakerley 1999) have been proposed to lie behind the increased
neurovirulence, it will be interesting to determine whether depletion of T- and NK-cell
derived IFN- plays any role. IFN- is known to protect neurons against SFV infection, and
the latter cell class is present also in SCID mice (Keogh et al. 2003). Interestingly, the
colloidal gold compound, myocrisin, was reported to increase A7(74) neurovirulence to L10
levels by unknown mechanism, with highly increased viral titers found also in
macrophages (Oaten, Jagelman, Webb 1980).
SFV is not considered to be a significant human pathogen, and the occasionally observed
seropositivity among laboratory workers supports concept of low virulence of SFV in
humans (Willems et al. 1979). However, one case of fatal encephalitis after accidental
infection of a laboratory worker with SFV Osterrieth strain has been reported (Fazakerley
2002; Willems et al. 1979). The scientist had been suffering from purulent bronchitis for one
year before the incident, and thus was possibly immunosuppressed. Finally, SFV infection
in mice is known to cause transient autoimmune demyelinating disease (Fazakerley 2002;
50
Mokhtarian et al. 1999), and malformations or death in fetal mice (Mabruk et al. 1989) but
neither of these sequelae have been reported in SFV infected humans.
51
3 Aims of the study
The general aim of this thesis project was to explore the applicability of Semliki Forest virus
VA7 in the targeting and treatment of malignant gliomas and colorectal carcinoma in vitro
and in experimental mouse models. In addition, the original aims were to elucidate the
roles of host innate and adaptive immune responses in the therapeutic outcome. The
specific aims of the three parts (I-III) of the thesis were as follows:
(I)
(II)
(III)
To study whether VA7 after systemic administration would be able to home to
orthotopic U87-Fluc glioma xenografts, establish productive infection and
prolong survival of immunocompromised mice.
To investigate whether VA7 virotherapy would prolong survival of
immunocompetent or immunocompromised mice bearing orthotopic, syngeneic
GL261 and CT-2A mouse gliomas, and to explore the cell autonomous antiviral
signaling of virus-infected U87-Fluc, GL261 and CT-2A glioma cell lines.
To characterize two clones of CT26 mouse colon carcinoma differing in their
immunogenicity and type I interferon responsiveness upon infection and to
utilize the two-clone model in the analysis of the roles of innate and adaptive
immunity in VA7-mediated oncolysis.
52
53
4 Materials and methods
Several different techniques of biotechnology, molecular biology and virology were applied
in this thesis work in vitro, in vivo and ex vivo. The cell lines (Table 2) and mouse strains
(Table 3) used and their sources are listed, and some of the techniques are presented at a
general level where greater details can be found in the corresponding publications (I-III).
4.1. CELL LINES AND MOUSE STRAINS (I-III)
Table 2. Cell lines used in the study. The names of the cell lines are in alphabetical order, their
origins and host species and the type of work they were used in are described. The cells were
cultured in 5% CO2 in a humidified atmosphere using cell culture media described in the
respective publications (I-III).
Cell line
From
Description
A172
ATCC
human glioma
in vitro
I
BHK-21
ATCC
Thomas Seyfried,
Boston College, MA
baby hamster kidney
in vitro
I, II, III
C57BL/6 glioma
in vitro, in vivo
II
Balb/c colon carcinoma
in vitro, in vivo,
ex vivo
II
CT-2A
CT26WT
ATCC
CT26.CL25
ATCC
Balb/c colon carcinoma
passaged from previous by
Balb/c colon carcinoma
John Bell
Daniel Silbergeld,
University
Balb/c glioma
of Washington, WA
CT26LacZ
DBT
in vitro
II
II
I, II
in vitro
in vitro, in vivo,
ex vivo
I
human glioma
in vitro
I
African green monkey kidney
in vitro
II, III
Balb/c breast carcinoma
in vitro
Figure 6A
G26-24
Chicago, Glyn Dawson
MBA-13
Aimo Salmi, University of
Turku
U87
ATCC
C57BL/6 oligodendroglioma
SV40 transformed mouse brain
cells positive for CNPase
oligodendrocyte marker*
human glioma
U87-Fluc
established
human glioma
U251
Jarmo Wahlfors,
University of Kuopio
* (Vähä-Koskela 2007)
II
in vitro
C57BL/6 glioma
4T1
in vitro
in vitro, in vivo,
ex vivo
Figure 6A,B
Geza Safrany, NRIRR
ATCC
Pirkko Härkönen,
University of Turku
Ref.
in vitro, in vivo,
ex vivo
in vitro, in vivo
GL261
VERO(B)
Used
II
I, II
54
Table 3. Mouse strains used in the present study. The names of the mouse strains, their
breeders and phenotypes are described. All mice were adult (4w-10w) females.
Strain
Breeder
Phenotype
Balb/cOlaHsd Foxn1 -/-
Harlan
athymic (defective T-cell immunity)
I
athymic (defective T-cell immunity)
II
nu
B6.Cg/NTac-Foxn1
NE10 Taconic
Ref.
C57BL/6JOlaHsd
House bred in Kuopio immunocompetent C57BL/6 mice
II
JaxC57BL/6J
House bred in Kuopio immunocompetent C57BL/6 mice
Figure 6B
Balb/cAnNCrl
Charles River
III
Fox Chase SCID
Charles River
immunocompetent Balb/c mice
Prkdcscid/scid mutation causing
defective T- and B-cell immunity
III
4.2 VA7-EGFP PRODUCTION (I-III)
VA7-EGFP plasmid (Figure 4) was heat shock transformed into XL1-Blue chemicompetent
bacterial cells and cultured in LB-Ampicillin (50 μg/ml) solution for 16 hours at +37oC.
Plasmid DNA was extracted using High-Speed Plasmid Miniprep Kit (GeneAid) and eluted
in heated elution buffer according to the manufacturer’s instructions. To enable successful
in vitro transcription, the plasmid was linearised with SmaI-restriction enzyme. The linear
plasmid was purified using chloroform extraction (without phenol) to remove all protein
impurities, especially RNase. The volume of the DNA digest was topped to 100 μl with
m.q. H2O and then mixed with 10 μl of 3M NaAc (pH 5.2). A volume of 110 μl of
chloroform was then mixed into the solution and the mixture was shortly centrifuged (1
min, full speed). Only half of the aqueous phase was recovered to prevent chloroform
carryover. DNA was precipitated with absolute EtOH (2.5 x the sample volume) at -20oC
for 1 hour followed by 15 min, +4oC, 16 000 g centrifugation, washing the pellet with 70%
cold EtOH (5min, 16 000 g, +4oC), air drying and dissolving the DNA pellet into 15 μl of
m.q. H2O.
The linearised and purified VA7-EGFP DNA was then used for in vitro transcription by
mMESSAGE mMACHINE SP6-kit (Ambion) and the synthesized viral RNA was LiCl
precipitated according to the manufacturer’s instructions. A total of 500 000 BHK-21 cells
were plated onto 6-well plates in High Glucose (4500 mg/L) DMEM supplemented with 5%
FCS, 2mM L-glutamine, Penicillin-streptomycin and 25mM HEPES, and 2 μg of the viral
RNA was transfected per well using TransIT-mRNA kit (Mirus Bio) according to the
manufacturer’s instructions. The transfection success was verified by observing the
cytopathic effect and fluorescence by microscopy, and the primary virus supernatant was
collected from the cells 48 hours after transfection and stored at -70oC until used for
secondary infections.
55
BHK-21 cells were cultured in a large (175 cm2) tissue culture flask until approximately 90%
confluence and infected with the primary virus in 20 ml of fresh cell culture medium (the
same composition as in the primary virus productions). The secondary virus preparation
was collected at 24 h post infection, before the the appearance of massive cytopathic effect.
The cell debris was pelleted by centrifugation and removed, the supernatant was sterile
filtered using 0,2 μm filter and then the secondary virus was divided into 0,5-1 ml aliquots
and stored at -70oC.
Figure 4. Vector map of VA7-EGFP plasmid with XmaI/SmaI linearization site. NSP=non
structural protein, Opal=opal stop codon, 26S=sub-genomic promoter, C=capsid, E=envelope
genes, 6K=structural gene, MCS=multiple cloning site, EGFP=enhanced green fluorescent
protein AmpR= ampicillin resistance. XmaI (5’ overhang) and SmaI (blunt) linearization site. In
addition to the linearization site, the unique BamHI and AvrII enzyme sites for MCS cloning
along with some commonly used sites for restriction enzyme analysis are presented.
56
4.3 DETERMINATION OF SEMLIKI FOREST VIRUS INFECTIOUS TITERS
(I-III)
The ability of VA7 to replicate in different cells and tissues in vitro and in vivo was
measured by plaque assay. 200 000 VERO(B) cells were seeded in low glucose (1000 mg/L)
DMEM supplemented with 5% FCS, 2mM L-glutamine, penicillin-streptomycin and 25mM
HEPES to 12-well plates one day prior to virus infectivity titrations. Tenfold virus dilutions
were prepared in cold VERO(B) cell medium and 400 μl of dilutions from 1/104 to 1/109
were applied as duplicates to the VERO(B) monolayers without removing the old warm
medium. The plates were immediately moved back to the +37oC incubator. The plates were
kept in the incubator for longer than 30 minutes to allow all of the virus particles enter the
cells, but for less than 2 hours to ensure that budding of new virions had not begun. During
the incubation period 0,4% agarose was prepared in +42oC VERO(B) medium from preprepared and autoclaved 4% low melting temperature agarose stock melted in boiling
water. The 0,4% agarose was stored at +42oC water bath until the end of the incubation. The
plates were taken from the incubator and the medium replaced with the warm agarose that
was left to polymerase for 10 minutes in room temperature before returning the plates to
the incubator. The final virus incubation under the agarose layer was carried out for 48
hours. Agarose was then removed and the cell monolayers stained with Crystal Violet
solution for 5 minutes and additional color was washed away with water. The plaques
were then counted manually against light.
4.4 PREPARING MICE AND CELLS FOR TUMOR TRANSPLANTATIONS (IIII)
The cancer cells were cultured in large (175 cm2) tissue culture flasks at +37oC, 5% CO2 in a
humidified atmosphere until a sufficient amount of cells were acquired. The cells were
washed with warm DPBS and trypsinized. Trypsin was inactivated with 10 ml of +4oC
Opti-MEM (Life Technologies), and from this point on the cells were stored on ice all of the
time. The cells were pelleted in 14 ml Falcon tube by centrifugation (+4oC, 1000 RPM, 4 min)
and most of the supernatant was aspirated away. The cells were re-suspended in the
remaining volume of Opti-MEM and the total volume of cell/supernatant mixture was
measured with a pipette. The cells were first diluted 1/100 in DPBS and then 1/2 in Trypan
Blue to measure the viability. Cells in the resulting 1/200 dilution were counted using
Bürker chamber, and based on total cell count, the required amount of +4oC Opti-MEM was
added to reach the target cell concentration in the transplantation solution (10 000 cells/5μl,
50 000 cells/5μl or 100 000 cells/5μl for intracranial glioma inductions, and 300 000 cells/50μl
for intradermal CT26 implantations).
Mice were pharmacologically anesthetized for the intradermal tumor transplantations, and
the anesthesia was combined with isoflurane gas anesthesia to allow the more invasive
intracranial tumor inductions. 1.2% isoflurane concentration was used with air flow of 200 –
250 ml/min. Isoflurane dose was increased when remaining foot pad reflexes were
57
observed and decreased if the mouse had shortness of breath. The chemical anesthesia
cocktail diluted in sterile saline contained medetomidine (1mg/kg) and ketamine analgesics
(75 mg/kg), and the anesthetic was administered intraperitoneally. Analgesia was
supplemented with subcutaneous, saline diluted carprofen (5 mg/kg) administered
immediately after the transplantation together with intraperitoneal injection of the
antisedative drug atipamezole (1 mg/kg), also diluted in saline. The eyes of anesthetized
mice were kept moist with sterile saline, and the furless athymic “nude” mice were
protected against anesthesia related hypothermia by using tissue blankets and hot water
bottles in the cages. The weight of the mice was monitored daily and the mice were
euthanized if there was >20% weight loss or any signs of severe distress such as
neurological symptoms. Tumor development was measured using small animal MRI in the
case of intracranial gliomas, or with a caliper instrument in intradermal tumors. Pain
caused by high tumor burden was medicated with subcutaneous injections of carprofen (5
mg/kg) diluted in saline and buprenorphine (0.1 mg/kg) diluted in sterile water.
The intradermal tumor transplants were administered as a bolus using 1 ml syringes with
solid 30G needles to the shaved flanks of the mice. For the purpose of intracranial injections
the mouse scalp was disinfected with 70% EtOH followed by scalpel incision on the
immediate right side of the head midline, an area free of large blood vessels. A hole was
manually drilled through the skull using 21G needle 2 mm to the right of bregma (II). Later
the protocol was refined and the hole was moved 1 mm anterior, in order to avoid injection
to the ventricles (Figure 6B). A Hamilton syringe with a lock plunger was vertically
attached to a stereotactic apparatus or laboratory lift bench and lowered to a 3 mm depth
utilizing needle stopper made out of cut 10 μl pipette tip. The needle (opening towards the
right hemisphere) was lowered into the brain, kept still for 1 minute and the cells were
slowly injected in 5 μl volume by moving slowly the lock screw plunger. The needle was
kept still for an additional 1 minute after the injection to avoid reflux, the skull surface was
wiped with damp 70% EtOH swab and the skin wound sutured. Finally, the mice were
antisedated and 1 ml of sterile saline was injected intraperitoneally to prevent dehydration.
4.5 SDS-PAGE AND WESTERN BLOTTING (III)
The CT26 cells for SDS-PAGE and western blotting were cultured at least in duplicate in 10
cm plates. After the cells had received the virus and interferon treatments for the
predetermined time periods, they were washed twice with ice cold DPBS. Then as adapted
(with modifications) from the protocol by Simmons et al., (Simmons, Wollish, Heise 2010),
500 μl ice-cold radioimmunoprecipitation assay buffer (50 mM Tris-HCl [pH 8.0], 150 mM
NaCl, 1 mM EDTA, 1% Igepal, 0.5% deoxycholate, 0.1% SDS, Complete Mini protease
inhibitor cocktail tablets [ref. 11836170001, Roche], phosphatase inhibitor cocktail [ref.
04906845001, Roche]) was added to the plates, the cells were scraped into microcentrifuge
tubes and kept on ice for more than 5 minutes. The cell debris was centrifuged and the total
protein content was measured using commercially available assay (Biorad, cat. 500-006)
according to the manufacturer’s instructions. Cell lysates were stored in -70oC until needed
58
and then boiled in 1x loading buffer (10x stock: 0.45M Tris-HCl, pH 6.5, 0.5M DDT 10%
SDS, Bromophenol Blue, <50% glycerol) before running in 10% SDS-PAGE and wet
transferring to Hybond-ECL membranes (RPN2020D, GE Healthcare). The anti-OAS rabbit
polyclonal antibody was a generous gift from Ilkka Julkunen, (Helsinki, Finland) and used
at 1:5000 dilution with 1h incubation at room temperature. Rabbit polyclonal antibody
detecting total STAT1 (BD transduction laboratories, cat. 610119) was used in 1:4000
dilution (RT, 1h). Rabbit polyclonal anti STAT1-P (Tyr701) was purchased from Cell
Signaling Technology (cat. 91715) and used in 1:5000 dilution (O/N, +4oC). Mouse
monoclonal anti -actin (C4) used in 1:3000 dilution (O/N, +4oC) was purchased from Santa
Cruz Biotechnology (cat. sc-47778). Secondary detections were carried out with
simultaneous incubation of anti-mouse-cy3 (-Actin detection) and anti-rabbit-cy5
fluorophore-coupled antibodies using Amersham ECL Plex Western blotting system
combined with Typhoon scanner.
4.6 SFV DETECTION FROM TISSUES BY IMMUNOHISTOCHEMISTRY (IIII)
The VA7 infection pattern in tumors and different healthy tissues was examined post
mortem by immunohistochemistry. The mice were sacrificed with CO2 and perfused with
PBS (I, III) or 4% formalin (II). The tumors, tissues and organs of interest were immersed in
PBS diluted in 4% paraformaldehyde for 24 hours (+4oC), after which they were moved into
70% ethanol (+4oC) until embedded into paraffin and sliced into 7-μm sections with a
microtome. The sections were lifted on SuperFrost Plus glass slides (Menzel-Gläser) and
incubated in +37oC for overnight to ensure proper attachment of tissue sections. Then the
sections were deparaffinized and rehydrated by dipping into the following solution series:
2 x xylene (3 min each), 2 x 100% EtOH (3 min each), 2 x 96% EtOH (3 min each), 1 x 70%
EtOH (3 min) and PBS (1 min). Then the slices were subjected to antigen retrieval by boiling
in 10mM trisodium citrate buffer (pH 6) for 20 minutes. The slices were allowed to cool
down and endogenous peroxidase activity was quenched by keeping the slices in methanol
containing 0.3% H2O2. The slices were washed 3 x 5 minutes in PBS and blocked in normal
goat serum (from Vectastain ABC kit, Vector Laboratories, Burlingame, CA) diluted in PBS
for 30 minutes. The slices were incubated for 60 minutes at room temperature in primary
rabbit polyclonal anti-SFV antibody diluted 1/4000 in PBS. Then the slices were washed 3 x
5 minutes in PBS and incubated with biotin coupled Vectastain secondary antibody for 30
minutes at room temperature, followed by 3 x 5 minutes PBS washes and 30 minutes of
ABC reagent incubation at room temperature according to the manufacturer’s instructions.
The slices were washed again for 3 x 5 minutes in PBS and finally DAB (SigmaFast, ref.
D4293) stained for a few minutes until a brown color was visible. Staining was terminated
by washing the slices under tap water, and slices dehydrated by running the multistep
rehydration series at the beginning of protocol in the reverse order (70% EtOH Æ xylene).
The slices were left to dry for a few minutes and finally mounted using a few drops of
Permount (Fischer Scientific, ref. SP15-500) mounting medium and a glass cover slip. The
cover slip was left to attach overnight at room temperature before microscopy.
59
5 Results
5.1 PERIFERALLY ADMINISTERED VA7 HOMES INTO ORTHOTOPIC U87FLUC GLIOMA XENOGRAFTS AND DESTROYS THEM (I)
While VA7 Semliki Forest virus had previously demonstrated potency in several xenograft
cancer models (Ketola et al. 2008; Maatta et al. 2008; Vaha-Koskela et al. 2006), it was
deemed of interest to investigate whether as a neurotropic virus, it was able to target
orthotopic gliomas. Three different human glioma cell lines, namely U87, U251 and A172
were tested for their VA7-EGFP sensitivity by infecting the cells with various MOIs and
measuring the cell viability in the Calcein AM assay at 96 h p.i. While all of the cells were
infectable and died when the high MOI=10 was used, U87 cells were clearly most sensitive
towards low virus doses, i.e. even MOI=0.01 infection resulted in almost complete cell
death at the 96 h p.i. time point (I, Fig. 1a). With the MOI=1 infection, all the cell lines
expressed abundantly SFV structural proteins E1, E2 as well as the pE2-E3 precursor
protein p62 (I, Fig. 1c) (Uchime, Fields, Kielian 2013). Due to its potential for
virotherapeutic targeting in vitro, the U87 cell line was chosen for further in vivo studies.
Subsequently, the U87 cell line expressing firefly luciferase (Fluc) was created by lentiviral
(MOI=25) transduction (I, Fig. S1D). Importantly, the transgene insertion did not influence
U87-Fluc cell proliferation or virus sensitivity compared to the parental U87 line (I, Fig.
S1B, C).
U87-Fluc tumors were first implanted subcutaneously into Balb/c OlaHsd Foxn1 -/- nude
mice followed by intravenous VA7-EGFP injection. While the virus homed into the small
(average size 96.5 mm3) subcutaneous tumors, completely eliminating them, the large
tumors (average size 442 mm3) were more refractory (I, Fig. 3). Eradication of half of the
large subcutaneous U87-Fluc tumors required three intravenous virus injections at 7 day
intervals (I, Fig. 2). However, even the large tumors that did survive the therapy despite the
three virus injections appeared to be more necrotic than control tumors when examined
histologically with live Fluc positive cells being found mostly in the tumor rim areas (I, Fig.
3B, C). Furthermore, explant cell lines derived from the surviving tumors had remained
infectable (I, Fig. S3).
According to histological evaluation carried out for samples collected at d3 post
intravenous VA7-EGFP injection (I, Fig. 6) and biodistribution monitoring using ectopic
renilla luciferase expression from VA7 (VA7-Rluc) during the first survival experiment (I,
Fig. S4), intravenous VA7 injection led to the infection of intracranial U87-Fluc gliomas.
This infection resulted in tumor elimination and survival of 16/17 mice. The survival
experiment was carried out twice and mice were monitored on the first occasion for 120
days post implantation after receiving either VA7-EGFP (n=5) (I, Figs. 4, 5b) or VA7-Rluc
(n=6) injections (I, Fig. 5a, b) and on the second time for 90 days after VA7-Rluc injection
(n=6) (I, Fig. S6). In the first experiment, one VA7-Rluc treated mouse (mouse 5) showed
residual tumor-derived bioluminescence at the last time point imaged (I, Fig 5a), and a
60
cluster of tumor cells was detected in the brain post mortem by anti-Fluc IHC staining (I, Fig.
5c). In the first survival experiment 3/6 of the mice (including mouse 5) in the VA7-Rluc
treatment group were let to grow close the bioluminescence endpoint (approximately one
week before the appearance of neurological symptoms) before the intravenous virus
injection. In the second survival experiment, 1/6 (mouse 1) of the orthotopic tumors
escaped the virotherapy (I, Fig. S6). Infective VA7-Rluc could not be recovered from the
brains, heart, liver or the spleen after the 90 days monitoring period of the second survival
experiment (I, Table 1) and the organs revealed no pathological changes in hematoxylineosin staining apart from occasional wound healing or scarring seen at the presumable
tumor induction sites (I, Figs., S5, S7). Similarly, while all of the above organs of the
infected naïve mice contained low (max 4.7 x 104 PFU/g) viral titers at day 3 p.i., infectious
VA7-Rluc could be recovered only from the brains on day 6 (average virus titers of 2 x 104
PFU/g) and from none of the organs on day 9 (I, Table 1). No pathological changes were
seen in these organs of the infected naïve mice at day 3 p.i (I, Fig. S7) and infected mice did
not experience weight loss as compared to PBS controls during the 48 day follow up (I, Fig.
S2).
5.2 THE VA7 RESISTANCE OF GL261 MOUSE GLIOMAS REFLECTS THEIR
ROBUST TYPE I IFN SIGNALING (II)
Spurred by the encouraging results in the U87-Fluc glioma xenograft model, it was decided
to proceed to test the oncolytic potential of VA7 in the GL261 and CT-2A-Fluc syngeneic
mouse glioma models. Both of the cell lines were infected and lysed in vitro, but neither i.v.
nor i.c. VA7 injections resulted in tumor infection in vivo and no therapeutic benefit was
obtained (II, Fig. 1). Intracranial VA7 injection into GL261 tumors led to infection of the
corpus callosum (II, Fig. 2d). Immunosuppression with CPA or rapamycin did not improve
the GL261 tumor infectivity or mouse survival following VA7 i.v. injections (II, Fig. 2)
though both immunosuppressing agents did reduce the amount of virus neutralizing
antibodies (II, Table 1). Importantly, CPA markedly potentiated the infection of healthy
brain tissue (II, Fig. 2e). IHC data revealed that the GL261 tumors were actually more
resistant in vivo than the healthy brain parenchyma and the spread of the virus was often
halted at the border of tumors (II, Fig. 2d). This occurred even after CPA facilitated
pancerebral infection (II, Fig. 2e). Explant cells isolated from the GL261 tumors were also
infectable (II, Fig. 4c). To exclude the possibility that some remaining adaptive antiviral
immune responses had been responsible for the virus neutralization, intracranial GL261
tumors were next implanted and treated in athymic mice. The GL261 tumors were resistant
also in B6 nude mice and thus neither i.v. nor i.c. VA7 injections improved survival (II, Fig.
3c). Intracranial VA7 injection resulted in infection of the corpus callosum also in nude mice
(II, Fig. 3a, b).
After exclusion of a central role for adaptive immunity in the therapeutic outcome, the role
of cell autonomous type I IFN response was investigated. A dose of 100 U/ml IFN-
pretreatment inhibited VA7 replication efficiently both in GL261 and CT-2A cells, and IFN was shown to work equally well when added 30 minutes or 90 minutes after the virus
61
application to GL261 cells (II, Fig. 4a) but was ineffective 4h after virus administration (II,
Fig. 4b). In order to test the hypothesis that therapeutic susceptibility to VA7 infection
depended on tumor IFN- sensitivity it was decided to treat GL261 cells in parallel with
virotherapy amenable U87-Fluc cells (I) with human and mouse IFN- followed by VA7
infection. Mouse, but not human IFN- blocked infection in GL261 cells, even though a 5
fold higher concentration of human IFN- was applied (100U/ml vs 500 U/ml). IFN- from
neither species had any observable effect on U87-Fluc cells (II, Fig. 4d), whereas human
IFN- markedly restricted VA7 infection in a patient biopsy-derived primary glioma cell
line. Interestingly, despite the clear inhibition caused by human IFN- pretreatment of the
primary glioma cells, VA7 infection often escaped at later time points (Figure 5). GL261
cells seemed to be especially sensitive to the effects of mouse IFN- as the pretreatment did
not completely block VA7 infection in mouse G26-24 oligodendroglioma and mouse 4T1
mammary carcinoma cell lines (Figure 6A). Furthermore, a tendency towards improved
survival was observed after two intracranial VA7 injections in the orthotopic G26-24 glioma
model (Figure 6B).
Figure 5. Primary human glioma cells were susceptible to VA7-EGFP infection. Primary human
glioma cells were VA7-EGFP infected (MOI=0.01 or MOI=1) with or without 100 U/ml or 1000
U/ml hIFN- pretreatment. The cytopathic effect was monitored with phase contrast microscopy
(upper panels) and the infection with fluorescence microscopy (lower panels). All pretreatments
were carried out in duplicates with similar results. While both MOIs resulted in efficient infection
and there were only a few viable cells left at the 72 h p.i. time point, especially the higher dose
hIFN- treatment conferred marked protection of the cells. The lower dose of hIFN- could not
always prevent the infection, which spread to monolayer cells and spheres. While also the
spheres were readily infected, the cytopathic effect was not seen in them before the last 144h
time point.
62
GL261 cells were autonomous in their IFN- production (II, Fig. 5b) and response (II, Fig. 5
a) upon VA7 infection as shown by IFN- ELISA and OAS western blot analysis,
respectively. Interestingly, the D51 VSV mutant induced more IFN- than VA7 in both
GL261 and DBT mouse glioma cell lines; in the latter cell line VA7 caused no detectable
IFN- secretion. Confluent GL261 cultures were found to be more resistant to VA7 infection
than subconfluent cultures especially when a low MOI was used (II, Fig. 5c). In the mixed
confluence assay, the confluent centers of individual wells were resistant while the
subconfluent culture rims remained infectable (II, Fig. 5d). The confluent cultures were also
more efficient IFN- producers (II, Fig. 5e). CT-2A (II, Fig. S1) and U87-Fluc (II, Fig. 5f)
cultures remained susceptible to infection regardless of the seeded cell number. U87-Fluc
cultures contained spheres that could sustain the infection for a prolonged time before cell
lysis (II, Fig. 6). Similar results were obtained with VA7-EGFP infection of primary glioma
cells (Figure 5). Finally, an in vivo antiviral assay was performed where two separate
cultures of GL261 cells were infected with VA7-EGFP and one of them was used for
intracranial tumor implantation while the other was left in the incubator. While the in vitro
culture was efficiently infected, the growth of intracranial tumors was unperturbed as
compared to uninfected controls (II, Fig. 7).
63
Figure 6. mIFN- blocks VA7-EGFP infection efficiently in GL261 mouse astrocytoma cells but
only partially in 4T1 mouse breast carcinoma and G26-24 mouse oligodendroglioma cells. A.
4T1 and G26-24 cells were efficiently infected with both virus doses tested (MOI=1 and
MOI=0.01), but the infection of GL261 cells was incomplete when the lower MOI value was
used. The cytopathic effect was monitored by phase contrast microscopy (upper panels) and the
infection by fluorescence microscopy (lower panels). With respect to the untreated cells G26-24
were the most sensitive to infection, and a massive cytopathic effect was visible already at 24h
p.i. even with the lower virus dose. 100U/ml mIFN- pretreatment efficiently prevented the
infection of GL261 cells but only partially that of 4T1 and G26-24 cells. Confluent 4T1 cultures
were also efficiently infected. The treatments were carried out in duplicate with similar results.
B. C57BL/6 mice received intracranial G26-24 oligodendrogliomas and two i.c. VA7-EGFP
injections. The first injection was given 6 days after tumor transplantation and the second
injection one week after the first one. Treatment group displayed a tendency towards improved
survival (P=0.0753, Log-rank (Mantel-Cox) Test).
64
5.3 SUBCLONE OF RESISTANT CT26 COLON CARCINOMA IS HIGHLY
SENSITIVE TO VA7 DUE TO WEAK TYPE I IFN RESPONSE (III)
To test whether VA7 virotherapy could be functional in an immuno-competent host
presuming that the tumors per se would be susceptible, different syngeneic mouse cancer
cell lines were screened for their type I IFN sensitivity. It was soon discovered that two
clones of CT26 mouse colon carcinoma, namely CT26WT and CT26LacZ dramatically
differed in their VA7 sensitivities. Productive VA7-EGFP infection in vitro was faster in the
CT26LacZ cells and the supernatant titers peaked at 24h p.i., one day earlier than in
CT26WT cultures (III, Fig. 1a). VA7 infection also killed CT26LacZ cells significantly more
efficiently than CT26WT cells as revealed in alamar blue assay at 48 h p.i., especially when
low MOI, 0.05, 0.1 or 0.5, were used (III, Fig. 1b). This difference reflected also in the in vivo
susceptibility of intradermal CT26WT and CT26LacZ tumors, and while the CT26LacZ
tumors were efficiently eradicated, VA7 therapy had no detectable effect on CT26WT tumor
growth as compared to sham treatment (III, Fig. 1c). Moreover, VA7-EGFP infection at
MOI=0.01 resulted in a massive infection and cell death of CT26LacZ cultures despite 1000
U/ml IFN- pretreatment, whereas the same IFN- concentration prevented infection and
death of the CT26WT cells even at MOI=1 (III. Fig. 1d). Nevertheless, at 1000 U/ml IFN-
some CT26WT cells could be infected, whereas in the parallel GL261 culture the infection
was completely blocked (Figure 7).
To investigate whether IFN- depletion was sufficient to confer CT26WT cells susceptible
to VA7-EGFP infection, CT26LacZ and CT26WT cells were first plated under agarose layer
and then VA7-EGFP was added with or without anti-IFN- or anti-panIFN- antibodies
into a well stamped through the agarose. While CT26LacZ cells were successfully infected
even without antibody treatment, the CT26WT cells could be successfully infected only
when anti-IFN- antibody was used (III, Fig. 2a). The ELISA assay revealed that both cell
lines secreted IFN- in response to the infection (III, Fig. 2b). The secreted interferons were
also biologically active, and conditioned supernatants collected from both of the cell lines
protected CT26WT, but not CT26LacZ cells against VA7-EGFP infection (III, Fig. 2c). VA7EGFP infection was found to be self-limiting in CT26WT but not in CT26LacZ cultures
when a very low infectious dose of 10 PFU was used. Pretreatment at 100 U/ml IFN-
inhibited virus replication weakly also in the CT26LacZ cells in this setting (III, Fig. 2d).
Combining the Jak1/Jak2 inhibitor ruxolitinib with IFN- resulted in successful infection of
both cultures, and interestingly pre-treating the CT26WT cells with a combination of
ruxolitinib and IFN- (100 U/ml) before virus inoculation resulted in a more prominent
infection as compared to the pretreatment with ruxolitinib alone (III, Fig.S2). Intriguingly,
transfection of CT26LacZ cells with either the STAT1 or the LacZ overexpression plasmid
prevented the 10 PFU VA7-EGFP infections (III, Fig. S4A). In addition to the phenotypic
change in the infectivity, transfection of either of the plasmids modified also the gene
expression of CT26LacZ cells to resemble that of CT26WT as shown by STAT1 and, even
more so, by OAS1 expression in western analysis (III, Fig. S4B).
65
Figure 7. CT26LacZ mouse colon carcinoma cells have diminished mIFN- response compared to
CT26WT cells, but interferon pretreatment (1000 U/ml) inhibits VA7-EGFP infection most
efficiently in GL261 mouse glioma cells. The cytopathic effect was observed with phase contrast
microscopy (upper panels) and the infection with fluorescence microscopy (lower panels). All
samples were done in triplicate with similar results.
66
To shed more light on the mechanisms responsible for the dramatic difference in the
antiviral response between CT26WT and CT26LacZ, both the uninfected and VA7-EGFP
infected cell lines were subjected to genome wide RNA-Seq transcriptome analysis.
Surprisingly, CT26WT cells constitutively expressed several PRRs such as RIG-I and Zbp1
(DAI), type I IFN signaling mediators such as IRF7, IRF9 and STAT1, and ISGs such as
PKR, Viperin, Ifit1/3 and ISG15/20 (III, Fig. 3a). In line with these data, “Activation of IRF
by Cytosolic Pattern Recognition Receptors” and “Interferon Signaling” were the top 2
canonical pathways revealed in Ingenuity IPA analysis, when the transcriptomes of
uninfected CT26WT and CT26LacZ cells were compared (III, Table 1). Some of the
constitutively overexpressed interferon stimulated genes in CT26WT cells such as Mx1,
IRF7 and STAT1 were true ISGs which, according to current understanding, are induced
from the low basal level only by interferon signaling and not directly by PRR/IRF3.
However, the latter two genes have been evaluated in human cells and the possibility of
species-related differences remains (Andersen et al. 2008; Grandvaux et al. 2002; Muller et
al. 1994). Furthermore, some of the true-ISGs such as Mx1 observed to be constitutively
expressed in CT26WT cells are not normally expressed at all at basal levels without
interferon stimulation (Muller et al. 1994). While IFN- transcripts could not be detected at
basal level in CT26LacZ cells, very weak IFN- mRNA expression was found in CT26WT
cells. It is unclear whether any IFN- protein was produced from this very low basal
transcription, but if this were to be the case, the amounts were too small to be detected by
ELISA (unpublished data). As an interesting side note, ISG20 is an RNase and a known
alphavirus inhibitor (Ryman and Klimstra 2008) and type I IFN has been recently proposed
to inhibit SFV infection by destroying its positive sense RNAs (Nikonov et al. 2013).
Infection did not cause transcriptional changes in the CT26WT cells, but the levels of
inflammatory and antiviral genes Atf3, Cxcl1/11, Fos, IFN- and IL-6 were induced in the
CT26LacZ cells. The expression of STAT1, OAS1 and OAS3 was confirmed at the protein
level (III. Fig. 3b). Strikingly, instead of expected STAT1 and OAS induction in CT26WT
cells upon VA7 infection or after IFN- treatment, the levels of these ISGs were actually
transiently decreased together with STAT1 phosphorylation 3h after IFN- stimulation.
In order to test for tumor infectivity in vivo, Balb/c mice carrying intradermal CT26WT or
CT26LacZ tumors were sacrificed at 24h, 48h and 96h post intratumoral VA7-EGFP injection
and the tumors were subjected to IHC using anti-SFV antibody (III, Fig. 4a). SFV could not
be found in the CT26WT tumors at any time point, whereas a progressive infection pattern
was observed in the CT26LacZ tumors. The initial infection site was detected around the
needle cavity 24 h p.i., whereas perivascular virus localization coinciding with moderate
and heavy central necrosis was evident 48 h p.i. and 96 h p.i., respectively. In agreement
with the histology, live tumor slices obtained from the CT26LacZ tumors, but not from the
CT26WT tumors could be efficiently infected ex vivo (III, Fig. 4b).
Intradermally implanted CT26LacZ tumors grew better in immune-compromised SCID mice
as compared to their immunocompetent counterparts, but one i.t. VA7-EGFP injection
eradicated the tumors completely from both groups (III, Fig. 4c, Figure 8). Furthermore,
when CT26WT and CT26LacZ tumors were implanted into separate flanks of Balb/c mice,
either at the same time (not shown) or when the CT26WT cells were implanted 8 days later
than the CT26LacZ cells, the VA7-EGFP mediated destruction of CT26LacZ did not prevent
67
or even retard the CT26WT tumor growth (III, Fig. 5a, b). However, when CT26WT tumor
implantation was performed approximately two months after CT26LacZ clearance, 2/4 mice
rejected the challenge (III, Fig. 5c). Finally, in order to create tumors which would be
heterogenic in their type I IFN response, CT26WT and CT26LacZ cells were mixed in
varying proportions and then infected with VA7-EGFP in vitro and in the Balb/c model in
vivo. The survival time of the infected in vitro cultures correlated linearly with the
proportion of CT26WT cells (III, Fig. 6a), whereas as few as 10% of these cells was enough
to reduce the in vivo survival from 100% to approximately 16% (III, Fig. 6b).
Figure 8. The growth curves of individual intradermal CT26LacZ tumors (from III, Fig.4C). One
intratumoral VA7-EGFP injection destroyed the flank tumors in both Balb/c (B) and SCID mice
(D). While the sham treated tumors grew aggressively in SCID mice (C), the growth was erratic
in Balb/c mice (A).
68
69
6 Discussion
Before the present work was initiated (I-III), VA7 had already shown potential in different
subcutaneous and orthotopic cancer models (see Literature Review chapter “VA7 oncolytic
virus”). Human A2058 melanoma xenografts in SCID mice were especially amenable to the
treatment, and even peripheral VA7 administration had resulted in remarkable tumor
regressions (Vaha-Koskela et al. 2006). However, the remissions remained transient and
relapses were found to result from tumor compartmentalization and a few virus resistant
cancer cells. The emergence of resistant cancer cell clones was observed also in VA7-treated
human squamous cell lung cancer cell line SW-900 in vitro (Maatta et al. 2008). Furthermore,
transient, though statistically significant, tumor growth retardation was seen in orthotopic
rat BT4C gliomas (Maatta et al. 2007). Despite the transient remissions, survival of the rats
was not improved even when VA7 was combined with the immunosuppressive agent
dexamethasone, a result at least partly suggested to stem from the detected Nab responses
against the therapeutic virus. Interestingly, only intratumoral or local injection of the virus
resulted in any therapeutic benefits in subcutaneous and orthotopic A549 nude mouse
models, respectively (Maatta et al. 2007; Maatta et al. 2008). It was speculated that IgM type
Nabs, known to develop also in nude (but not in SCID) mice, could be responsible for the
discrepancy between the therapeutic efficacy of the local and systemic virus administration
routes, as A2058 xenografts in SCID mice were highly prone also to i.v. and even i.p. VA7
injections. However, it is likely that the kinetics of the type I IFN response also had a role,
as A549 cells are known both to produce and respond to type I IFN in response to virus
infection (Rautsi et al. 2007), and IFN- can rescue GL261 mouse glioma cells even when
applied 90 minutes post infection (II, Figure 4a). It is therefore plausible that in contrast to
intratumoral injection, peripheral infection route caused gradually developing VA7 viremia
resulting in a slow and focal delivery of the virus into the A549 tumors thus providing the
cancer cells with more time to react to the infection by type I IFN signaling. Furthermore,
while it is known that IFN- is not species cross-reactive between mouse and human, and
thus this interferon class produced by the healthy mouse cells such as pDCs in blood
presumably has little effect on human A549 tumors, some IFN- classes might exhibit more
cross-species reactivity (Alexenko, Ealy, Roberts 1999). VA7 administration i.v. delayed the
induction of MxA in the subcutaneous A549 tumors as compared to i.t. injection (Maatta et
al. 2007). MxA is a good biomarker of the type I IFN response since in contrast to many
other ISGs, it is not induced directly by viral infection but requires signaling through
IFNAR, i.e. it is a true ISG (Holzinger et al. 2007).
Systemically administered VA7 showed high safety and remarkable efficacy even against
large orthotopic U87-Fluc glioma xenografts (I). Even though VA7 destroyed 16/17 of i.c.
U87-Fluc tumors regardless of the original tumor size and small s.c. tumors were equally
amenable, only half of the large subcutaneous tumors could be destroyed. This required
increase of VA7 dosing up to three injections (I, Figure 2). The reason for the relative
resistance of these large tumors remained elusive as no virus resistant U87 populations
were observed which was in contrast to the earlier results with A2058 melanoma xenografts
70
(Vaha-Koskela et al. 2006). One possibility is that some tumor cells survived the limited
time span of infection in a semi-adherent state in the large necrotic areas (I, Fig. 3) out of the
reach of blood perfusion and, thus, the virus. Fully transformed malignant cells do not have
a contact inhibition and in addition can thrive in an anchorage independent manner for
example in pleural effusions without committing anoikis (Basak et al. 2009; Hanahan and
Weinberg 2011; Weinberg 2014). It is also possible that when the U87-Fluc tumors grew
over a certain threshold size, the number of CSC spheres became too high. These spheres
were readily infected with VA7 but they died rather slowly in comparison to monolayers
(II, Fig.6A). A similar result was obtained with primary human glioma samples (Figure 5).
Furthermore, as CSC formation is thought to be a dynamic and reversible process
(Hanahan and Weinberg 2011; Weinberg 2014), then the absence of virus-resistant
populations in the escaped tumors would have not been surprising.
Though speculative, clues from bioluminescence data obtained from two mice suggested
that innate immune responses may have played a role in the remarkable efficacy against the
orthotopic U87-Fluc tumors. For instance, on one occasion (I, Fig. S4, mouse 1) the
luminescence signal from the tumor disappeared 24h p.i., then returned to be briefly visible
between the 24 d and 42 d time points, to finally disappear again at the final imaging time
point of 101 days p.i. As day 6 p.i. was the last time point when infectious virus could be
recovered from the brain (I, Table 1), it is unlikely that tumor re-infection would have
accounted for this phenomenon. The tumor growth kinetics of mouse number 5 (I, Fig. S4)
was presumably similar, judging from the long term survival of the mouse (> 4 months)
despite the recurrence of tumor-derived bioluminescence at the final imaging point 64 days
p.i. These data highlight the power of preclinical bioluminescence imaging. Similar to
single-photon emission computed tomography (SPECT), bioluminescence imaging is very
sensitive and can readily detect even a few target cells. It is superior in this respect to the
traditional methods such as MRI or CT, which typically have a good resolution but only
mediocre sensitivity. However, optical imaging has the drawback that it requires xenogenic
reporter enzyme expression by the target cells, which are likely to be immunogenic in
immunocompetent hosts (Murakami et al. 2003). Interestingly, a similar pattern of initial
tumor regression and transient remission followed by short relapse before final elimination
was observed in one untreated Balb/c control mouse that received immunogenic intradermal
CT26LacZ tumor (Figure 8, mouse 7). The hypothesis of involvement of innate immunity in
the eradication of U87-Fluc tumors after VA7 infection will need to be confirmed in the
future by systematic antibody (or genetic) depletion studies of different innate immune cell
subclasses and their combinations, especially macrophages, neutrophils and NK-cells.
Another question which would be interesting, albeit difficult to answer, is whether U87
tumors with defects in their type I IFN response (II) would be equally sensitive to VA7 in a
fully immunocompetent host, or whether the efficacy would require pharmacological
inhibition of adaptive responses such as Nabs. Though perhaps counterintuitive, it is
conceivable that in vivo VA7 semi-permissive tumors such as DBT or G26-24 mouse gliomas
(Vaha-Koskela et al. 2013 and Figure 6, respectively) could become responsive to the
virotherapy when the neutralizing adaptive responses had been temporally attenuated. In
such a hypothetical model, the initial immunosuppression would be combined with virus
injections to achieve partial tumor infection, which would then be followed by immunemediated tumor destruction when immunosuppression was released. As a side note, G26-
71
24 is an oligodendroglioma (Lo Cicero et al. 2011) and tumors of this cellular origin could
be good targets for A7(74) SFV mediated virotherapy, since the virus naturally infects
perivascular brain oligodendrocytes and neurons, but not astrocytes (Fazakerley et al.
1993). Another important question although likely beyond any feasible means of
investigation, is how well the U87 cell line available from ATCC which is one of the most
extensively studied human glioblastomas, represents the original tumor isolated from a 44
old Caucasian male in the late 1960s (Ponten 2014). In the framework of this dissertation
and the field of oncolytic virotherapy in general, it would be especially interesting to know
if the documented defects in type I IFN induction and response in U87 cells (I, (Wollmann,
Robek, van den Pol 2007)) were an early aberration and thus a characteristic for the whole
tumor, or appeared later as a clonal trait as seems to be the case for CT26LacZ cells derived
from the VA7 resistant CT26WT line (III). It will be also crucial to explore in more detail the
nature of these defects in the future, as it seems that also U87 cells can respond to type I
IFN treatment under certain circumstances thus limiting viral infection (Alain et al. 2010
and Minna Niittykoski, personal communication). Similarly IFN- pretreatment can weakly
inhibit VA7-EGFP virus replication also in CT26LacZ cultures when minuscule virus
amounts were used (III, Fig.2D). Hence, the type I IFN signaling defects seen in U87 and
CT26LacZ cells are possibly more quantitative/epigenetic than qualitative/mutational in
their nature. In support of this hypothesis, type I IFN treatment was shown to induce ISG15
expression in U87 cells, but the cells were not successful in mounting antiviral state against
VSV infection (Wollmann, Robek, van den Pol 2007). This demonstrates that the signaling
circuits beginning from IFNAR do exist also in the U87 cells though rewired or suppressed.
Curiously, U87 cells seem to be handicapped both in type I IFN induction and response
(Wollmann, Robek, van den Pol 2007). Although only the latter of these defects is sufficient
for efficient tumor infection, defects in the type I IFN induction become relevant in
xenograft models. As IFN- shows only limited cross-species reactivity, the host derived
IFN- may not protect the xenograft tumors (II). For example, VERO green monkey kidney
cells used for VA7 titration (II, III) and baby hamster kidney cells (BHK) used for VA7
production (II, III), are both defective in type I IFN induction while the response to these
cytokines seem to be normal (Juang et al. 1998; MacDonald et al. 2007). Indeed, BHK cells
have been shown to make extremely sensitive xenograft targets for virotherapy (Tseng et al.
2004). Finally, VA7 proved to be safe even in athymic Balb/c mice. No viral induced
pathology could be detected in VA7-treated U87 tumor-bearing or in naïve mice (I, Fig. S7),
and the virus did not persist in the brains of infected control mice beyond day 6 p.i. (I,
Table 1). In line with these results, B6 nude mice did not develop any symptoms of virus
encephalitis, and no obvious virus induced pathologies could be seen in the brains (II, Fig.
3). Interestingly, a persisting subclinical A7(74) infection of nude mouse CNS has been
reported earlier (Amor et al. 1996) and the discrepancy might stem from the documented
attenuating nsp3 mutation that had occurred during construction of the plasmid prA7(74)
later used for construction of VA7 (Tuittila and Hinkkanen 2003). As an interesting side
note, U87 cultures typically retained their “net-like” morphology (II, Fig. 4d) even when at
confluence, while the freezing process in DMSO containing medium occasionally (but not
invariably) caused these cells to adapt to a monolayer growth pattern possibly reflecting
epigenetic gene regulatory switch in the transcriptional program (unpublished
72
observations). DMSO has long been known to be able to induce cell differentiation
(Tsiftsoglou and Sartorelli 1979).
VA7 failed to improve survival of C57BL/6J mice carrying orthotopic GL261 glioma tumors
even when immunosuppressive agents rapamycin or CPA was applied (II). Rapamycin is
thought to improve OV efficacy by reducing the amount of systemic type I IFN produced
by pDCs and hence to promote the infection of tumors, which are semi-competent in their
type I IFN response. However, if the tumors have managed to retain much of their IFN
signaling intact, this partial aid of rapamycin might be insufficient to confer therapeutic
benefit for the IFN sensitive viruses. Similar results were obtained when VA7 was
combined with another type I IFN signaling inhibitor valproic acid (VPA) (Zemp et al.
2010) in orthotopic GL261 tumor model (unpublished data). However, the experiment was
carried out twice and while VPA-VA7 combination resulted in a significant survival
advantage in the first pilot study, the encouraging result could not be repeated in a larger
scale experiment. Both CPA and rapamycin decreased the level of circulating Nabs (II,
Table 1) and unexpectedly CPA dramatically enhanced VA7 replication only in normal
brain tissue (II, Figure 2). The neurologic symptoms in the GL261 bearing mice treated with
the CPA/VA7 combination developed around d10 p.i., at the same time with the symptoms
of mice that received VA7 monotherapy resulting in no viral encephalitis. Hence, despite
enhanced viral CNS replication after CPA treatment, the high tumor burden was probably
the reason for the neurological symptoms requiring euthanasia of this cohort. The SFV-CPA
combination has been investigated once previously, and depending on the dosing schedule,
suppression of T- or B-cells was speculated to account for the increased virulence (Bradish,
Allner, Fitzgeorge 1975). A high dose CPA is widely immunosuppressive (Huyan et al.
2011; Lacki et al. 1997) and in the future it will be interesting to determine whether the
depletion of IFN- secreting T-cells and NK-cells has any role in the observed increased
neurovirulence, a mechanism which could potentially also explain age-dependent
neurovirulence (see Literature review, Chapter “2.3.1 Taxonomy, natural life cycle and
pathogenesis of Alphaviruses and Semliki Forest virus”).
Surprisingly, VA7 replication could not be detected in any of the GL261 tumors at any time
point, while the dissociated cell explants resumed their infectivity. GL261 cells both
produced and responded to IFN-, aborting VA7 infection even when 100 U/ml of IFN-
was applied as late as 90 minutes p.i., although, not beyond 4h p.i. Though speculation, it is
possible that the ISGs responsible for the termination of VA7 infection can target only the
steps occurring before this 4h p.i time point in the alphavirus lifecycle, for instance
coinciding with the switch from negative to positive strand RNA-synthesis taking place
between 4-6 h p.i. (Sawicki and Sawicki 1980; Sawicki et al. 1981a; Sawicki et al. 1981b).
However, as earlier discussed, many ISGs, including known anti-alphavirus effectors ZAP
and IFIT1, attack viral translation, making it difficult to understand why the block would
occur before bulk translation of viral proteins from the genomic and subgenomic RNAs had
been initiated. Indeed, an alternative plausible explanation for the temporal threshold of
IFN- effects could be the early effects of nsP2 induced host cell transcriptional shutdown,
a process known to be fully implemented around 6 h p.i. (Akhrymuk, Kulemzin, Frolova
2012). Whatever the underlying mechanism, GL261 cells seem to be especially resistant to
VA7 infection and they can abort the infection in response to IFN- more efficiently than
73
the 4T1 mouse breast carcinoma cells or G26-24 mouse oligodendroglioma cells (Figure 6A)
or CT26WT mouse colon carcinoma cells (Figure 7). GL261 and CT26WT tumors were
resistant to VA7 in vivo (II and III), as were also the 4T1 tumors (Markus Vähä-Koskela,
personal communication). However, two i.c. injections of VA7 displayed a tendency of
improved survival in G26-24 model (Figure 6B), which was not nevertheless statistically
significant. Considering the remarkable in vitro sensitivity of G26-24 cells, it could be
fruitful to test multiple i.c. virus injections in this model.
Intriguingly, i.c. injection of VA7 led to infection of the corpus callosum, both in the
immunocompetent and athymic mice (II, Figs. 2d, 3a, b). Infection with parental A7(74),
SFV replicon and even with a neuron-detargeted SFV4 resulted in the same infection
pattern of the white matter tract of the corpus callosum and these cells have been reported to
be oligodendrocytes (Fazakerley et al. 2006; Ylosmaki et al. 2013). If mice are administered
with poly I:C i.p. 3h before the i.c. infection, the corpus callosum remains protected. As
peripheral administration of both avirulent and virulent SVF strains reportedly results also
in infection of neurons, the observed dramatic oligodendrocytic infection upon
intracerebral administration warrants further studies. It will be interesting to examine
whether this phenomenon relates to a differential intracerebral fluid kinetics resulting from
virus administration induced inflammation, which could open access to oligodendrocyte
dense corpus callosum (Laule et al. 2004). Alternatively oligodendrocytes and possibly also
other cells of glial origin such as GL261 might be inherently more responsive to endocrine
type I IFN than neurons, and the latter cell types might respond better to type II IFN. If this
proves to be the case, it could also explain why several immunosuppressive Th1/IFN-
inhibitors increase alphavirus infection in neurons (see chapter “2.4.8 Safety and Ethics of
Oncolytic Virotherapy”). Whatever the underlying mechanism, SFV tropism for the
oligodendrocytes of corpus callosum might be exploitable for gene therapy purposes in this
brain region, especially as lentiviral vectors can be readily pseudotyped with SFV
glycoproteins (Froelich et al. 2011).
While defects in type I IFN induction may disable the antiviral defense of cancer cells in
vitro, this is not necessarily sufficient for in vivo tumor targeting because there are
additional type I IFN sources in the tumor microenvironment (Liu et al. 2013). These
findings are consistent with the present results, and while VA7 infection of DBT mouse
glioma did not lead to detectable IFN induction (II, Fig.5b) these tumors were refractory in
vivo (Vaha-Koskela et al. 2013). However, minute VA7 infection of DBT brain tumor tissue
has occasionally been seen in immunohistochemical staining (Vaha-Koskela et al. 2013) but
no signs of any infected cells were detected here in the GL261 brain tumors. On the
contrary, VA7 infection often terminated just at the border of parenchyma-to-malignant
tissue (II, Fig2d, e, Fig.3a, b). Further support for the importance of extratumoral type I IFN
sources in virus clearance was obtained when VA7 pre-infected GL261 cells were found to
form tumors with comparable kinetics to uninfected cells. Control cells left in the incubator
were efficiently infected (II, Fig. 7) demonstrating that once full and instantaneous infection
of tumors has been reached, the autonomous type I IFN production alone is not able to
protect the cells from cytolytic infection. However, it is important to realize that human
tumors which have lost autonomous type I IFN induction might falsely appear to be virus
susceptible when analyzed in nude/SCID mouse models. When these human tumors are
74
implanted into the mouse IFN milieu, they may de facto become deficient in their IFN
response as the key early player, IFN- is not species cross-reactive (II). If the tumors
cannot even partly compensate for the lack of exogenous IFN- by their own type I IFN
production, they may be artificially rendered vulnerable to viral infections. While both CPA
and rapamycin treatments reduced the quantity of circulating Nabs, GL261 tumors were
treated in athymic B6 mice to exclude the possibility of any remaining adaptive antiviral
immune responses. Tumor infectivity or therapeutic efficacy was not improved even by
direct intracranial virus injection (II, Fig. 3) further confirming that innate rather than
adaptive responses inhibited VA7 infection in the GL261 tumors.
Interestingly, confluent GL261 cultures were markedly more resistant than sub-confluent
cultures to VA7 infection at low MOI in vitro (II, Fig. 5c). This result could not be explained
only by the higher amounts of type I IFN in the supernatants of confluent cultures, as the
difference in infectivity remained in an assay where tissue culture plates had confluent
centers surrounded by sparse rim areas (Fig.5d). Interestingly, unlike confluent GL261
cultures (II, Fig.5C, D), confluent 4T1 cultures were not virus resistant (Figure 6A) thus
resembling CT-2A cells and VERO(B) cells (II, Fig.S1). Defective type I IFN induction
probably explains at least the sensitivity of the confluent VERO(B) cultures (Juang et al.
1998) and possibly also the CT-2A cells (II, data not shown). The exact mechanism for the
notable resistance of confluent GL261 cultures warrants further studies, and mechanisms
such as basolateral receptor location and antiviral signaling via gap-junctions may be
possible. It is also conceivable that cell-to-cell signaling of confluent in vitro GL261 cultures
activates the same virus hostile (possibly astrocytic) gene expression pattern present in the
respective in vivo tumors. Indeed, the conditions of both confluence and senescence have
been reported to induce ISGs, including STAT1 (Perou et al. 1999). Furthermore, type I IFN
gene program was recently shown to be activated when mixed cultures were created out of
glioblastoma cells and nervous tissue (Nayernia et al. 2013). Importantly, this effect was not
observed for glioblastoma or nervous monocultures, or for mixed cultures with MCF-7
breast carcinoma cells. Indeed, thes heterotypic interactions between gliomas and their
stromas might explain the notable discrepancy observed between the in vitro and in vivo
infectivity of GL261 gliomas (II). Supporting the spatio-temporal plasticity of GL261 cells,
the suppression of infectivity in vivo was reversible, and sparse explant cultures isolated
from the tumors were again infectable with VA7 (II, Fig.4c). The potential role of cell
culture medium also remains to be investigated in the future, as the RPMI medium used for
CT-2A cells has been reported to support baculovirus, adenovirus and lentivirus
transductions clearly better than DMEM used for GL261 cells (Mahonen et al. 2010). The
number of seeded cells did not markedly affect the infectivity of U87-Fluc cells, but these
cells do not typically form confluent monolayers (II, Fig. 5f).
Two clones derived from the same CT26 murine colon carcinoma cell line were here found
to differ both in terms of their autonomous antiviral defense and their immunogenicity,
which had important implications on VA7 mediated oncolytic virotherapy (III). It was
discovered that CT26LacZ cells differ dramatically in their IFN- responsiveness in
comparison to their parental cell line CT26WT, and this difference directly translated into a
difference in the therapeutic efficacy in vivo (III, Fig. 6B). Due to the presence of the
xenogenic LacZ antigen, CT26LacZ tumors were highly immunogenic in
75
immunocompetent Balb/c mice (Figure 8, III Fig. 4C). It was first hypothesized that
CT26LacZ tumor infection might stimulate robust anti-tumor adaptive immune responses,
which could be responsible to their eradication. However, VA7 proved equally powerful
against CT26LacZ tumors in immunosuppressed SCID mice demonstrating that the virusinduced tumor destruction occurs independently of adaptive immunity (III, Fig. 4C).
VA7 infection caused efficient production of IFN- both in CT26WT and CT26LacZ cell
clones, indicating that also CT26LacZ has intact PRR machinery (III, Fig. 2B). In contrast,
1000 U/ml IFN- pretreatment inhibited VA7 replication in CT26WT cells while having
only a minor (but not a nonexistent) effect on CT26LacZ cells, where VA7 continued to
replicate even after 10 000 U/ml IFN- pretreatment (III, Fig. 1D and S1). Interestingly,
some CT26WT cells were infected despite 1000 U/ml IFN- treatment, whereas the same
treatment completely blocked VA7 infection in GL261 cells (Figure 7). GL261 cells seem to
be exceptionally poised both in their type I IFN production and response and even one log
lower IFN- concentration applied 90 minutes after virus was shown here to be sufficient to
arrest the infection (II, Fig. 4). However, the in vivo outcome for CT26WT tumors was the
same as in the GL261 tumor model, and again no virus antigen could be detected by IHC in
CT26WT tumor tissue (III, Fig.4A). In line with these data, also ex vivo infection of live
CT26WT tumor slices was poor (III, Fig.4B).
SFV viremia is followed by virus penetrance into the tumors, and if the tumor is
susceptible, the initial focal plaques start to rapidly expand, resulting in a widely spread
intratumoral infection. Interestingly, the initial infection is often focal, regardless of the
route of administration ((Maatta et al. 2007; Vaha-Koskela et al. 2006), II, Fig. 6, III, Fig. 4A).
This suggests that though the i.t. VA7 administration route has been occasionally found to
be more efficient than systemic virus injection (Maatta et al. 2007), even after direct
intratumoral injection, much of the virus might come to the tumor from the blood. In an
attempt to dissect the mechanisms underlying the apparent incongruity between the in vitro
and in vivo infectivity of CT26WT tumors, the focal in vivo infection pattern was mimicked
in vitro by casting an agarose overlay on CT26WT and CT26LacZ cultures followed by VA7
injection into a stamped hole in the middle of the culture plate well. In this focal infection
model, VA7 replicated efficiently only in the CT26LacZ cells. However, the infection in the
CT26WT cultures could be rescued by adding anti-IFN- antibody (III, Fig. 2A). The
importance of type I IFN signaling in the termination of the infection was confirmed in an
experiment, where CT26LacZ and CT26WT cells were infected in parallel with very small
10 PFU amount of VA7-EGFP. The infection was productive in CT26LacZ cultures but self
limiting in CT26WT cells. The infection of CT26WT cells could be rescued using the
JAK/STAT inhibitor ruxolitinib, and interestingly, the infection was further enhanced when
combined with 100 U/ml IFN- treatment (III, Fig. 2D, Fig. S2).
Remarkably, when CT26LacZ cells were transfected with either the STAT1 or LacZ
expression plasmids, infection with 10 PFU of VA7-EGFP was inhibited (III, Fig. S4A).
Furthermore, the expression of OAS1 was observed to increase to the level seen in CT26WT
cells with both transfections (III, Fig.S4B). LacZ plasmid transfection induced also STAT1
expression to the levels seen in unstimulated CT26WT cells, whereas STAT1 expression was
higher after STAT1 expression plasmid transfection as expected. This experiment strongly
76
suggested that the diminished type I IFN responsiveness in the CT26LacZ cells had been
attributable to plastic epigenetic mechanisms rather than to irreversible mutations at the
genomic level. The exact PRRs responsible for responding to transfections remain to be
identified, and both cell lines express at least the DNA sensors, cGAS and IFI16
(unpublished observations), but CT26WT cells transcribe significantly more DAI (Zbp1)
than CT26LacZ cells do (III, Fig. 3A) (Lam, Stein, Falck-Pedersen 2014). Presently, it is
unclear whether transfection itself is sufficient to induce these changes, or whether
transfection of any RNA or any non-coding DNA would cause similar effects. However,
non-coding DNA has been shown to induce ISGs such as STAT1 in several cancer cell lines
in a type I IFN-dependent manner (Li et al. 2005). Importantly, the induction was reported
to occur independently of the DNA type (plasmid or PCR-amplified) or the method used
for transfection (Lipofectamine, electroporation or CaPO4), and furthermore the DNA or the
transfection reagent on their own did not evoke any induction. Interestingly, it is
documented that while OAS can be activated directly by IRF3, STAT1 does not respond to
IRF3 overexpression, suggesting that despite its un-stimulated basal expression, STAT1 is
induced as a true ISG (Grandvaux et al. 2002). The hypothesis that the increase in OAS1 and
STAT1 levels of CT26LacZ cells upon plasmid transfection is type I IFN-mediated thus
receives further support. It is, however, possible, that in addition to a type I IFN mediated
action, direct IRF3 activation may have played an additional role in ISG induction seen in
CT26LacZ cells after plasmid transfections, as the change in OAS1 levels was more
prominent than the change seen in STAT1 (III, Fig. S4B).
VA7 replicated efficiently in CT26LacZ tumors, and after initial infection evident around
the i.t. injection site, the infection was found to localize perivascularly (III, Fig. 4A). The
perivascular infection coincided with massive necrosis of the surrounding tumor
parenchyma. Thus, secondary tumor destruction resulting from compromised blood flow
probably contributed to the therapeutic efficacy as has been documented previously with
VSV in the CT26LacZ tumor model (Breitbach et al. 2007). VSV induced vascular shutdown
was found to be mediated by neutrophils attracted to the inflammation site and
interestingly, the VA7-infected CT26LacZ cells expressed CXCL1 and CXCL11 chemokines
(III, Fig. 3A). Importantly, rather than infection of the thin endothelial layer previously
suggested to underlie the efficacy of oncolytic vaccinia, VSV and reovirus after transient
VEGF/FGF “pulsing”(Breitbach et al. 2013; Kottke et al. 2010), profound VA7 infection of
perivascular CT26LacZ tumor parenchyma was seen beginning already at 48h p.i (III, Fig.
4). Though the target cells are different, the susceptibilities of tumor vasculature and
parenchyma to viral infection may be mechanistically similar, and indeed, RNA-Seq data
revealed statistically significant over-expression of both FGF- and PDGF-receptors in
CT26LacZ tumors (unpublished observations). In perfect agreement, both FGFR and
PDGFR are known to signal using Ras as a second messenger, and activated Ras/MEK
signaling has been demonstrated to suppress IFN-induced transcription (Brooks, Kilgour,
Smith 2012; Christian et al. 2012; Ostman and Heldin 2007).
There was a constitutively high transcription of numerous genes of inflammation-, PRRand IFN-signaling pathways observed in CT26WT- but not in CT26LacZ cells in the
genome wide analysis and some of these differences were confirmed at the protein level
(III, Fig.3). Comparative IPA pathway analysis of the transcriptomes of CT26WT and
77
CT26LacZ control cells further corroborated this observation, i.e. IRF activation by cytosolic
PRRs and interferon signaling were the top 2 canonical pathways differentially regulated
between the cell lines (III, Table 1). Several of the genes (or their human counterparts)
expressed at high levels in CT26WT cells have been reported to have alphavirus inhibitory
activity either alone, or in synergism with ZAP. These included many well-known
mediators and effectors of antiviral signaling such as ISG15, ISG20, Viperin, Ifi47, RIG-I,
PKR, IRF7, Ifitm3, MDA5, STAT1, IRF9, Ifi44l and OAS3 (Schoggins et al. 2011). In addition
to STAT1 phosphorylation, type I IFNs also induce STAT1 gene expression as a positivefeedback mechanism, and also the unphosphorylated STAT1 has been reported to
translocate into nucleus driving long-term expression of some ISGs (Cheon and Stark 2009).
Strikingly, the majority of those ISGs reported to be induced by unphosphorylated STAT1
were also constitutively expressed in CT26WT cells including, OAS1, OAS2, OAS3, STAT1,
IFI44, IFI44L, IFIH1 (MDA5), IFI35, IFIT3, Mx1, IRF7 and IFIT1 (p56) (III, Fig. 3A). This
suggests that the constitutive ISG expression in these cells may be driven by the high level
expression of unphosphorylated STAT1. IRF3 which is known to be expressed
constitutively and not induced by type I IFN, was expressed at equal levels in CT26WT and
CT26LacZ cells (Honda and Taniguchi 2006).
Constitutive expression of OAS and MxA in human tumor cells has been proposed as a
marker for tumor resistance against an IFN sensitive oncolytic virus VSV (MoerdykSchauwecker et al. 2013). Human MxA has been shown to have antiviral functions against
SFV in transgenic mice (Hefti et al. 1999) and it is induced in human xenograft tumors upon
VA7 therapy (Maatta et al. 2007). Indeed, the high expression of Mx1 and Mx2 genes as well
as of the different OAS genes was detected in the virus resistant CT26WT cells (III, Fig. 3A).
Mouse Mx1 and Mx2 genes are both antiviral ISGs, but Mx2 is non-functional in all
commonly used laboratory mouse strains due to the presence of an insertional mutation
resulting in a translational frameshift (Jin et al. 1999). The fact that in addition to actual
effectors, high basal levels of transcription of even Mx2 along with some IFN inducible
pseudogenes were found in CT26WT cells favors a ubiquitous, non-specific explanation for
the antiviral state in these cells. Furthermore, in addition to mouse Mx1 (as well as human
MxA), at least human IRF7 and ISG20 genes have been shown to be true ISGs, i.e. induced
only by interferon signaling and not directly via the PRR/IRF3 route (Grandvaux et al. 2002;
Holzinger et al. 2007; Muller et al. 1994). Thus it is possible that the constitutive expression
of many of the antiviral genes in the CT26WT cells was actually attributable to the very low
level IFN- transcription seen in these cells (unpublished observations), which may be
responsible for auto/paracrinic signaling and upregulation of multiple ISGs. A similar effect
has been reported to protect human hepatocytes from viral infection (Tsugawa et al. 2014)
and MEFs from malignant transformation (Abbas, Lichtman, Pallai 2007; Adekolu-John and
Fagbami 1983; Chen et al. 2009). Furthermore, low, but constitutive IFN- expression has
been reported to play a role in priming of bronchial epithelial cells to withstand influenza
virus attack (Hsu et al. 2012). The low constitutive type I IFN expression by tumors or
stromal cells has also been reported earlier, and very recently this mechanism was found to
account for the resistance of some tumors against oncolytic VSV (Liu et al. 2013; Tsai et al.
2011). However, it remains to be determined whether any IFN- protein was produced
from the very low level basal transcription in CT26WT cells, as it was not detectable by
ELISA (unpublished observations). Low levels of IFNs may be consumed by the cells and
78
thus IFN- production is often detected indirectly by its ability to induce true ISGs. A high
level of ISG15 expression has been found to protect tumors from IFN mediated apoptosis
(Tisserand et al. 2011), and elevated expression of ISG15 was seen in CT26WT cells,
potentially explaining the lack of IFN- treatment-induced cytotoxicity (III, Fig. 3A). It
would be worthwhile to examine whether ISG15 inhibition could re-introduce the natural
cytotoxicity of type I IFNs in CT26WT cells, a potentially powerful therapeutic effect
recently attributed to IAP antagonists (Beug et al. 2014). Finally, the basal type I IFN
expression and priming has been shown to enhance the effects of other antiviral cytokines
such as IFN- by increasing the expression of STAT1 (Gough et al. 2010). Indeed, it was
possible to detect higher constitutive expression of STAT1 in CT26WT cells compared to
CT26LacZ cells (III, Fig.3), and anti-IFN- antibody markedly enhanced VA7 replication in
the CT26WT cells in the focal infection assay under agarose overlay (III, Fig. 2A).
The cell autonomous virus resistance mechanism in CT26WT tumors may be similar to that
reported recently by Malilas et al. for VSV in NIH3T3 cells (Malilas et al. 2013), as a similar
high constitutive expression pattern of STAT1 and OASL2 genes was seen also in the
CT26WT cells (III, Fig. 3A). If high STAT1 mediated OASL2 expression is also revealed to
account for the CT26WT virus resistance, it will be interesting to determine whether it can
also be attributed to the high activity of the CUG2 gene, as CUG2 expression was
approximately twice higher in CT26WT control cells compared to CT26LacZ controls
(unpublished observations). Intriguingly, higher levels of both constitutive and IFN-
induced expression of several STAT1 dependent ISGs were recently reported by Cho et al.
to be responsible for the positive sense RNA-virus resistance in granular cell neurons as
compared to virus-permissive cortical neurons (Cho et al. 2013). In addition to many other
ISGs, exactly the same antiviral gene signature of IRGs reported by Cho and colleagues
(Irgm1, Irgm2, Igtp, Ifi47 and Tgtp1) was also expressed in CT26WT cells (III, Fig. 3A). Cho
and coworkers also presented preliminary data suggesting that epigenetic regulation might
be responsible for the differences in IFN/STAT1 signatures in different neuronal subtypes
resulting in differential susceptibility to RNA-virus infection. As an interesting side note,
Cho and colleagues also discovered that VEEV was resistant to most of the ISGs that had
potent antiviral effects against the other positive sense RNA-viruses tested, a result
consistent with other reports concerning the insensitivity of pathogenic VEEV to the
antiviral effects of type I IFN (Ryman and Klimstra 2008).
Amongst the several antiviral genes expressed at constitutively high levels in CT26WT
cells, there were two crucial downstream mediators of IFNAR signaling, IRF9 and STAT1,
which together with STAT2, form the ISGF3 complex. Reduced signaling by this complex
could explain the stoichiometric rather than absolute defects in the type I IFN response seen
in CT26LacZ cells, further proposing that a global difference in antiviral gene signatures
rather than single mutations underlie the virus resistance of CT26WT tumors. Indeed, an
increase in the expression of IRF9 and one of the type I IFN receptor chains, IFNAR2, have
been shown recently to correlate with anti-alphaviral state of mature neuronal cells (Farmer
et al. 2013). The known fact that IFN- binding to its receptor induces not only
phosphorylation of STAT1 protein but also an increase in its transcription level (Varble et
al. 2013) poses an obvious caveat to the data interpretation as the observed reduction in
STAT1 transcript and protein levels in CT26LacZ cells (III, Fig. 3) could be either the result
79
of or the cause of the type I IFN signaling defects in these cells. Reduced STAT1 expression,
as also observed in some melanomas (Wong et al. 1997) may benefit the CT26LacZ tumor
survival by reducing its immunogenicity. However, STAT1 induction has been implicated
in the development of tumor resistance to genotoxic stress (radiation and chemotherapy)
and radioresistant tumors concomitantly developed resistance to the type I IFN sensitive
HSV-1 mutant (Khodarev et al. 2004; Khodarev, Roizman, Weichselbaum 2012). Hence, the
antiviral response profile of heavily pretreated tumors, as in most oncolytic virotherapy
trials, may be inherently unfavorable. Furthermore, as the majority of preclinical rodent
cancer models in use are created by exposing the animals to high dosages of mutagens such
as 3-methylcholanthrene (MCA), there is a risk that some of these tumors inherently
express high levels STAT1 and are thus virus resistant. Thus, more natural genetic and
spontaneous cancer models should be favored, which would also better display the
heterogeneity seen in human tumors (Rodriguez-Madoz et al. 2014).
Data from IFN- pretreatment-, single-foci expansion assay-, RNA-Seq- and western blot
analysis and results from in vivo and ex vivo infection experiments together argue in favor of
a model, where CT26WT (but not CT26LacZ) cells constitutively reside in a semi-primed
but inactivated antiviral state. It is notable that the same antiviral state could be induced in
CT26LacZ cells by transfection of plasmid DNA (III, Fig. S4) but not with IFN- treatment
(III, Fig. 3B). In this model, focal infections in the CT26WT tumors in vivo lead to rapid IFN secretion and final paracrine priming (i.e. increase in the PRR and ISG levels) of the
neighboring uninfected tumor cells, which upon PRR engagement were then capable of
activating the ISGs by phosphorylation and aborting the infection. Interestingly, this
resembles the regulation of poliovirus tropism, in which tissues expressing constitutively
high levels of PRRs and ISGs, such as RIG-I, PKR and OAS (all upregulated also in
CT26WT cells) have been found to be non-permissive for poliovirus replication (IdaHosonuma et al. 2005). These events in CT26WT tumors are difficult to capture by standard
in vitro infection techniques in dissociated cell cultures without agarose cover as any
secreted IFN- is immediately diluted into the culture medium. The dilution possibly also
contributed to the absence of VA7 infection-induced transcriptional changes in CT26WT
cells in RNA-Seq analysis (III, Fig. 3A) and to the absence of STAT1 or OAS protein
induction or STAT1 phosphorylation in western analysis (III, Fig. 3B). In support of this
hypothesis, treatment with a high dose exogenous IFN- induced STAT1 phosphorylation
in CT26WT cells (III, Fig. 3B). Very small (10 PFU) amounts of VA7-EGFP in CT26WT
cultures could mimic the in vivo resistance of these tumors (III, Fig. 2D, Fig. S2, Fig. S4A),
but it remains to be determined if a small scale focal infection could also induce adequate
signaling changes, which could be then detected at mRNA or protein levels.
VA7 infection caused efficient eradication of CT26LacZ tumors also from SCID mice,
demonstrating that adaptive immune responses did not play any major role at least in the
acute therapeutic phase (III, Fig. 4C). Long term follow-up of the SCID mice was not
possible, as they gradually develop neurological symptoms upon VA7 infection (VahaKoskela et al. 2006). Indeed, the infected SCID mice had to be eventually sacrificed due to
ataxic symptoms appearing roughly at day 16 post infection, when the tumors had been
already completely destroyed by the infection. Notably, whereas the CT26LacZ tumors
grew aggressively in the SCID mice, tumor development in the immunocompetent Balb/c
80
mice was erratic (Figure 8). Some of the CT26LacZ tumors grew fairly quickly, whereas
others disappeared quite soon after i.t. saline injection. Tumor sizes in the third group were
increasing and decreasing spontaneously until finally reaching the criteria for sacrifice of
the animals. Strikingly, the transplanted CT26LacZ tumor of Balb/c mouse number 7
disappeared after sham treatment for 9 days until aggressive growth resumed again, tumor
reached a large size, and then spontaneously disappeared again without recurrence. A
similar growth pattern of the CT26LacZ tumors in immunocompetent Balb/c mice was seen
when tumors were transplanted for histological analyses (III, Fig. 4A), whereas all of the
CT26WT tumors grew aggressively (unpublished observations). This pronounced
immunogenicity of the CT26LacZ tumors was somewhat unexpected, as the growth pattern
has been previously reported by Dr Restifo, the creator of the two cell lines, to resemble
that of CT26WT tumors (Wang et al. 1995). The observed discrepancy possibly derives from
different routes of transplantation, i.e. subcutaneous used by Dr. Restifo and the
intradermal route used here; the latter has been reported to be more immunogenic due to
activation of resident skin dendritic cells (Bonnotte et al. 2003).
Double-flank and rechallenge experiments where the CT26LacZ and CT26WT cancer cells
were implanted into separate flanks of Balb/c mice either simultaneously, or the latter after
CT26LacZ infection, were carried out to determine if cross-immunity against CT26WT
tumors could be induced upon eradication of the virus-susceptible CT26LacZ tumors.
CT26LacZ tumor infections induced protective immunity against a delayed CT26WT
rechallenge in 50% of the treated mice, while the eradication of CT26LacZ tumors did not
influence the CT26WT outgrowth in the double-flank experiments upon simultaneous
tumor induction (III, Fig. 5). After infection, CT26LaZ cells are capable of secreting
efficiently at least IFN- (III, Fig. 2B), an early type I IFN that can be induced by IRF3
independently of positive feedback signaling via IFNAR (Honda and Taniguchi 2006). As
type I IFNs are generally considered as being necessary for evoking antigen-specific T-cell
responses in alphavirus based vaccination protocols (Leitner et al. 2006) and CT26WT and
CT26LacZ cells have been reported to express shared tumor antigens (Kershaw et al. 2001)
it was surprising to observe the lack of any cross-protective immunity against CT26WT
tumors inflicted by cure of the LacZ expressing tumors in double flank- (III, Fig. 5) and
clone-mixing (III, Fig. 6B) experiments. In the latter experiment as few as 10% of CT26WT
cells in the implantation mixture was sufficient to seriously compromise the efficacy of the
therapy. Hence, the dominant anti-tumor immune response generated during SFV
oncolysis was likely targeted at LacZ rather than against any of the previously identified
and shared tumor antigens (Castle et al. 2014; Kershaw et al. 2001). LacZ is a xenoantigen
and due to its alien origin and redundancy for tumor cell survival it might function in the
clonal-mix setting as an immune defense “decoy” diverting immune responses against the
dominant but ultimately ineffective target and lead to escape of CT26WT cells modeling
antigen loss variants. Immunoediting of cancer cell clones is known to occur in patients
undergoing antigen-specific immunotherapies, and malignant cell clones that have
undergone epigenetic/genetic suppression of the antigen or antigen-presenting machinery
are positively selected (DuPage et al. 2012). Tumors and their stroma can also be directly
immunosuppressive or evade immune attack by inflammation-induced transient cellular
dedifferentiation, resulting in an altered antigen landscape (Holzel, Bovier, Tuting 2013;
Mellman, Coukos, Dranoff 2011). Indeed, transcription of several immunosuppressive
81
genes was detected in the CT26 lines including PD-L1, TGF1, STAT3, RCAS1, IDO2, MCSF and strikingly high levels of Galectin-1 (unpublished observations) (Rabinovich,
Gabrilovich, Sotomayor 2007).
While oncolytic virotherapy may optimally be able to elicit broad-spectrum anti-tumor
immune responses through epitope spreading, these responses may occasionally be biased
against only a subset of immune-dominant epitopes. The number of shared epitopes
between the virus sensitive and resistant clones may also be an important factor. Notably,
in the present genome-wide transcriptome analysis between the CT26WT and CT26LacZ
cells it was found out that 10% of all the sequenced mRNAs were differentially transcribed
at the basal level, some of these potentially representing important epitope targets for
immunotherapy. Moreover, viral infection per se can also cause dramatic changes in gene
expression, potentially driving further apart the epitomes of the infected and uninfected
cancer cell clones. Tumors defective in the type I IFN response such as CT26LacZ whose
transcriptome reacts only mildly during the early phase of the infection (III, Fig. 3A) might
be optimal immuno-virotherapeutic targets also in this regard.
82
83
7 Summary and conclusions
In conclusion, VA7 Semliki Forest virus showed oncolytic potential in murine models of
glioma and colorectal carcinoma, but the therapeutic efficacy was strictly dependent on the
type I interferon responsiveness of the tumors (I-III). In the future it will be necessary to
develop fast and reliable methods for tumor prescreening in order to direct the therapy to
those patients who will most likely benefit from alphavirus-based oncolytic virotherapy.
Functional tests such as focal ex vivo infections of sliced live tumor tissue or dissociated
cultures with and without type I IFN pretreatments, are likely to yield the most reliable
information, especially if tumors (and possible metastases) are sampled in thorough
manner. In these tests, the type I IFN class and concentration should be chosen carefully,
with the high affinity early IFN- possibly being a better choice than any of the late IFN-
subclasses. Historically, a concentration as low as 5 U/ml of type I IFN has been used in
these screens to identify cancer cell lines with diminished responsiveness to type I IFN
(Stojdl et al. 2000). As a comparison, the minimum IFN concentration that efficiently
blocked VA7 replication in CT26WT cells was 1000U/ml of IFN-. Nevertheless, this level of
cell autonomous antiviral signaling was sufficient to seriously hamper the in vivo efficacy of
oncolytic virotherapy (III). Interestingly, in a recent analysis the majority of primary human
melanoma samples were shown to have defects in their type I IFN response as compared to
healthy melanocytes. While recombinant IFN- was used instead of IFN-, the tested
interferon concentration of 100 U/ml was rather high (Wollmann et al. 2013). Functional
tests are nevertheless too laborious and time consuming for routine large scale clinical
applications. Hence, high throughput applications based on cutting edge technologies such
as single cell genome and transcriptome sequencing (Navin et al. 2011) that are able to
account tumor heterogeneity could prove invaluable, presuming that appropriate
resistance and susceptibility markers are discovered. The detection of constitutive
transcription of interferon stimulative genes such as OAS and MxA may be a good
surrogate marker for the anticipated virus resistance (III). However, the lack of such a
signature should not be uncritically interpreted as a definitive signal for virus sensitivity, as
ISGs are typically induced by infection and interferons (II) (Maatta et al. 2007). Immunovirotherapeutic targeting of type I IFN responsive tumors can also be envisioned, as
alphavirus infection is naturally highly immunogenic. Paradoxically, much of this
immunogenicity stems from the robust type I IFN response evoked by the alphavirus
infection, which could, under optimal circumstances, be harnessed to achieve therapeutic
benefits.
84
85
References
Abbas A, Lichtman A, Pallai S. 2007. Cellular and molecular immunology. Sixth ed. Elsevier/ Saunders.
Adekolu-John EO and Fagbami AH. 1983. Arthropod-borne virus antibodies in sera of residents of kainji lake
basin, nigeria 1980. Trans R Soc Trop Med Hyg 77(2):149-51.
Ahmed A, Thompson J, Emiliusen L, Murphy S, Beauchamp RD, Suzuki K, Alemany R, Harrington K, Vile
RG. 2003. A conditionally replicating adenovirus targeted to tumor cells through activated RAS/P-MAPKselective mRNA stabilization. Nat Biotechnol 21(7):771-7.
Akhrymuk I, Kulemzin SV, Frolova EI. 2012. Evasion of the innate immune response: The old world
alphavirus nsP2 protein induces rapid degradation of Rpb1, a catalytic subunit of RNA polymerase II. J
Virol 86(13):7180-91.
Alain T, Lun X, Martineau Y, Sean P, Pulendran B, Petroulakis E, Zemp FJ, Lemay CG, Roy D, Bell JC, et al.
2010. Vesicular stomatitis virus oncolysis is potentiated by impairing mTORC1-dependent type I IFN
production. Proc Natl Acad Sci U S A 107(4):1576-81.
Alexenko AP, Ealy AD, Roberts RM. 1999. The cross-species antiviral activities of different IFN-tau subtypes
on bovine, murine, and human cells: Contradictory evidence for therapeutic potential. J Interferon
Cytokine Res 19(12):1335-41.
Alsharifi M, Lobigs M, Regner M, Lee E, Koskinen A, Mullbacher A. 2005. Type I interferons trigger systemic,
partial lymphocyte activation in response to viral infection. J Immunol 175(7):4635-40.
Alvarez-Breckenridge CA, Yu J, Price R, Wei M, Wang Y, Nowicki MO, Ha YP, Bergin S, Hwang C, Fernandez
SA, et al. 2012a. The histone deacetylase inhibitor valproic acid lessens NK cell action against oncolytic
virus-infected glioblastoma cells by inhibition of STAT5/T-BET signaling and generation of gamma
interferon. J Virol 86(8):4566-77.
Alvarez-Breckenridge CA, Yu J, Price R, Wojton J, Pradarelli J, Mao H, Wei M, Wang Y, He S, Hardcastle J, et
al. 2012b. NK cells impede glioblastoma virotherapy through NKp30 and NKp46 natural cytotoxicity
receptors. Nat Med 18(12):1827-34.
Amor S, Scallan MF, Morris MM, Dyson H, Fazakerley JK. 1996. Role of immune responses in protection and
pathogenesis during semliki forest virus encephalitis. J Gen Virol 77 ( Pt 2 )(Pt 2):281-91.
Andersen J, VanScoy S, Cheng TF, Gomez D, Reich NC. 2008. IRF-3-dependent and augmented target genes
during viral infection. Genes Immun 9(2):168-75.
Andersson M, Paabo S, Nilsson T, Peterson PA. 1985. Impaired intracellular transport of class I MHC antigens
as a possible means for adenoviruses to evade immune surveillance. Cell 43(1):215-22.
Andtbacka R, et al. 2013. Meeting abstract: OPTiM: A randomized phase III trial of talimogene laherparepvec
(T-VEC) versus subcutaneous (SC) granulocyte-macrophage colony-stimulating factor (GM-CSF) for the
treatment (tx) of unresected stage IIIB/C and IV melanoma. Journal of Clinical Oncology : Official Journal
of the American Society of Clinical Oncology 31.
Arudchandran R, Brown MJ, Song JS, Wank SA, Haleem-Smith H, Rivera J. 1999. Polyethylene glycolmediated infection of non-permissive mammalian cells with semliki forest virus: Application to signal
transduction studies. J Immunol Methods 222(1-2):197-208.
Ayala-Breton C, Barber GN, Russell SJ, Peng KW. 2012. Retargeting vesicular stomatitis virus using measles
virus envelope glycoproteins. Hum Gene Ther 23(5):484-91.
Bach P, Abel T, Hoffmann C, Gal Z, Braun G, Voelker I, Ball CR, Johnston IC, Lauer UM, Herold-Mende C, et
al. 2013. Specific elimination of CD133+ tumor cells with targeted oncolytic measles virus. Cancer Res
73(2):865-74.
Baer C, Claus R, Plass C. 2013. Genome-wide epigenetic regulation of miRNAs in cancer. Cancer Res
73(2):473-7.
Barbi J, Pardoll D, Pan F. 2013. Metabolic control of the treg/Th17 axis. Immunol Rev 252(1):52-77.
Barry G, Fragkoudis R, Ferguson MC, Lulla A, Merits A, Kohl A, Fazakerley JK. 2010. Semliki forest virusinduced endoplasmic reticulum stress accelerates apoptotic death of mammalian cells. J Virol 84(14):736977.
86
Barry G, Breakwell L, Fragkoudis R, Attarzadeh-Yazdi G, Rodriguez-Andres J, Kohl A, Fazakerley JK. 2009.
PKR acts early in infection to suppress semliki forest virus production and strongly enhances the type I
interferon response. J Gen Virol 90(Pt 6):1382-91.
Bartlett DL, Liu Z, Sathaiah M, Ravindranathan R, Guo Z, He Y, Guo ZS. 2013. Oncolytic viruses as
therapeutic cancer vaccines. Mol Cancer 12(1):103,4598-12-103.
Basak SK, Veena MS, Oh S, Huang G, Srivatsan E, Huang M, Sharma S, Batra RK. 2009. The malignant pleural
effusion as a model to investigate intratumoral heterogeneity in lung cancer. PLoS One 4(6):e5884.
Bayne LJ, Beatty GL, Jhala N, Clark CE, Rhim AD, Stanger BZ, Vonderheide RH. 2012. Tumor-derived
granulocyte-macrophage colony-stimulating factor regulates myeloid inflammation and T cell immunity
in pancreatic cancer. Cancer Cell 21(6):822-35.
Benedict CA and Ware CF. 2001. Virus targeting of the tumor necrosis factor superfamily. Virology 289(1):1-5.
Beug ST, Tang VA, LaCasse EC, Cheung HH, Beauregard CE, Brun J, Nuyens JP, Earl N, St-Jean M, Holbrook
J, et al. 2014. Smac mimetics and innate immune stimuli synergize to promote tumor death. Nat Biotechnol
32(2):182-90.
Bhat R and Rommelaere J. 2013. NK-cell-dependent killing of colon carcinoma cells is mediated by natural
cytotoxicity receptors (NCRs) and stimulated by parvovirus infection of target cells. BMC Cancer
13:367,2407-13-367.
Binder GK and Griffin DE. 2001. Interferon-gamma-mediated site-specific clearance of alphavirus from CNS
neurons. Science 293(5528):303-6.
Bonnotte B, Gough M, Phan V, Ahmed A, Chong H, Martin F, Vile RG. 2003. Intradermal injection, as
opposed to subcutaneous injection, enhances immunogenicity and suppresses tumorigenicity of tumor
cells. Cancer Res 63(9):2145-9.
Boozari B, Mundt B, Woller N, Struver N, Gurlevik E, Schache P, Kloos A, Knocke S, Manns MP, Wirth TC, et
al. 2010. Antitumoural immunity by virus-mediated immunogenic apoptosis inhibits metastatic growth of
hepatocellular carcinoma. Gut 59(10):1416-26.
Bradish CJ and Titmuss D. 1981. The effects of interferon and double-stranded RNA upon the virus-host
interaction: Studies with togavirus strains in mice. J Gen Virol 53(Pt 1):21-30.
Bradish CJ and Allner K. 1972. The early responses of mice to respiratory or intraperitoneal infection by
defined virulent and avirulent strains of semliki forest virus. J Gen Virol 15(3):205-18.
Bradish CJ, Allner K, Fitzgeorge R. 1975. Immunomodification and the expression of virulence in mice by
defined strains of semliki forest virus: The effects of cyclophosphamide. J Gen Virol 28(2):225-37.
Bradish CJ, Allner K, Maber HB. 1971. The virulence of original and derived strains of semliki forest virus for
mice, guinea-pigs and rabbits. J Gen Virol 12(2):141-60.
Breakwell L, Dosenovic P, Karlsson Hedestam GB, D'Amato M, Liljestrom P, Fazakerley J, McInerney GM.
2007. Semliki forest virus nonstructural protein 2 is involved in suppression of the type I interferon
response. J Virol 81(16):8677-84.
Breitbach CJ, Reid T, Burke J, Bell JC, Kirn DH. 2010. Navigating the clinical development landscape for
oncolytic viruses and other cancer therapeutics: No shortcuts on the road to approval. Cytokine Growth
Factor Rev 21(2-3):85-9.
Breitbach CJ, Paterson JM, Lemay CG, Falls TJ, McGuire A, Parato KA, Stojdl DF, Daneshmand M, Speth K,
Kirn D, et al. 2007. Targeted inflammation during oncolytic virus therapy severely compromises tumor
blood flow. Mol Ther 15(9):1686-93.
Breitbach CJ, Arulanandam R, De Silva N, Thorne SH, Patt R, Daneshmand M, Moon A, Ilkow C, Burke J,
Hwang TH, et al. 2013. Oncolytic vaccinia virus disrupts tumor-associated vasculature in humans. Cancer
Res 73(4):1265-75.
Brenner H, Kloor M, Pox CP. 2014. Colorectal cancer. Lancet 383(9927):1490-502.
Brenner MK, Gottschalk S, Leen AM, Vera JF. 2013. Is cancer gene therapy an empty suit? Lancet Oncol
14(11):e447-56.
Bridle BW, Boudreau JE, Lichty BD, Brunelliere J, Stephenson K, Koshy S, Bramson JL, Wan Y. 2009. Vesicular
stomatitis virus as a novel cancer vaccine vector to prime antitumor immunity amenable to rapid boosting
with adenovirus. Mol Ther 17(10):1814-21.
Bridle BW, Chen L, Lemay CG, Diallo JS, Pol J, Nguyen A, Capretta A, He R, Bramson JL, Bell JC, et al. 2013.
HDAC inhibition suppresses primary immune responses, enhances secondary immune responses, and
abrogates autoimmunity during tumor immunotherapy. Mol Ther 21(4):887-94.
87
Briolat V, Jouneau L, Carvalho R, Palha N, Langevin C, Herbomel P, Schwartz O, Spaink HP, Levraud JP,
Boudinot P. 2014. Contrasted innate responses to two viruses in zebrafish: Insights into the ancestral
repertoire of vertebrate IFN-stimulated genes. J Immunol 192(9):4328-41.
Brooks AN, Kilgour E, Smith PD. 2012. Molecular pathways: Fibroblast growth factor signaling: A new
therapeutic opportunity in cancer. Clin Cancer Res 18(7):1855-62.
Broz P and Monack DM. 2013. Newly described pattern recognition receptors team up against intracellular
pathogens. Nat Rev Immunol 13(8):551-65.
Buijs PR, van Eijck CH, Hofland LJ, Fouchier RA, van den Hoogen BG. 2014. Different responses of human
pancreatic adenocarcinoma cell lines to oncolytic newcastle disease virus infection. Cancer Gene Ther
21(1):24-30.
Burrell RA, McGranahan N, Bartek J, Swanton C. 2013. The causes and consequences of genetic heterogeneity
in cancer evolution. Nature 501(7467):338-45.
Byrnes AP, Durbin JE, Griffin DE. 2000. Control of sindbis virus infection by antibody in interferon-deficient
mice. J Virol 74(8):3905-8.
Cao W, Manicassamy S, Tang H, Kasturi SP, Pirani A, Murthy N, Pulendran B. 2008. Toll-like receptormediated induction of type I interferon in plasmacytoid dendritic cells requires the rapamycin-sensitive
PI(3)K-mTOR-p70S6K pathway. Nat Immunol 9(10):1157-64.
Castle JC, Loewer M, Boegel S, de Graaf J, Bender C, Tadmor AD, Boisguerin V, Bukur T, Sorn P, Paret C, et
al. 2014. Immunomic, genomic and transcriptomic characterization of CT26 colorectal carcinoma. BMC
Genomics 15(1):190,2164-15-190.
Cerullo V, Pesonen S, Diaconu I, Escutenaire S, Arstila PT, Ugolini M, Nokisalmi P, Raki M, Laasonen L,
Sarkioja M, et al. 2010. Oncolytic adenovirus coding for granulocyte macrophage colony-stimulating factor
induces antitumoral immunity in cancer patients. Cancer Res 70(11):4297-309.
Chattopadhyay A, Wang E, Seymour R, Weaver SC, Rose JK. 2013. A chimeric vesiculo/alphavirus is an
effective alphavirus vaccine. J Virol 87(1):395-402.
Chawla-Sarkar M, Leaman DW, Jacobs BS, Tuthill RJ, Chatterjee-Kishore M, Stark GR, Borden EC. 2002.
Resistance to interferons in melanoma cells does not correlate with the expression or activation of signal
transducer and activator of transcription 1 (Stat1). J Interferon Cytokine Res 22(5):603-13.
Cheema TA, Wakimoto H, Fecci PE, Ning J, Kuroda T, Jeyaretna DS, Martuza RL, Rabkin SD. 2013.
Multifaceted oncolytic virus therapy for glioblastoma in an immunocompetent cancer stem cell model.
Proc Natl Acad Sci U S A 110(29):12006-11.
Chen HM, Tanaka N, Mitani Y, Oda E, Nozawa H, Chen JZ, Yanai H, Negishi H, Choi MK, Iwasaki T, et al.
2009. Critical role for constitutive type I interferon signaling in the prevention of cellular transformation.
Cancer Sci 100(3):449-56.
Chen N, Bellone CJ, Schriewer J, Owens G, Fredrickson T, Parker S, Buller RM. 2011. Poxvirus interleukin-4
expression overcomes inherent resistance and vaccine-induced immunity: Pathogenesis, prophylaxis, and
antiviral therapy. Virology 409(2):328-37.
Cheon H and Stark GR. 2009. Unphosphorylated STAT1 prolongs the expression of interferon-induced
immune regulatory genes. Proc Natl Acad Sci U S A 106(23):9373-8.
Cho H, Proll SC, Szretter KJ, Katze MG, Gale M,Jr, Diamond MS. 2013. Differential innate immune response
programs in neuronal subtypes determine susceptibility to infection in the brain by positive-stranded
RNA viruses. Nat Med 19(4):458-64.
Choi SY, Collins CC, Gout PW, Wang Y. 2013. Cancer-generated lactic acid: A regulatory,
immunosuppressive metabolite? J Pathol 230(4):350-5.
Christian SL, Zu D, Licursi M, Komatsu Y, Pongnopparat T, Codner DA, Hirasawa K. 2012. Suppression of
IFN-induced transcription underlies IFN defects generated by activated ras/MEK in human cancer cells.
PLoS One 7(9):e44267.
Cloughesy TF, Yoshimoto K, Nghiemphu P, Brown K, Dang J, Zhu S, Hsueh T, Chen Y, Wang W, Youngkin
D, et al. 2008. Antitumor activity of rapamycin in a phase I trial for patients with recurrent PTEN-deficient
glioblastoma. PLoS Med 5(1):e8.
Conry RM, Khazaeli MB, Saleh MN, Allen KO, Barlow DL, Moore SE, Craig D, Arani RB, Schlom J, LoBuglio
AF. 1999. Phase I trial of a recombinant vaccinia virus encoding carcinoembryonic antigen in metastatic
adenocarcinoma: Comparison of intradermal versus subcutaneous administration. Clin Cancer Res
5(9):2330-7.
88
Csatary LK, Gosztonyi G, Szeberenyi J, Fabian Z, Liszka V, Bodey B, Csatary CM. 2004. MTH-68/H oncolytic
viral treatment in human high-grade gliomas. J Neurooncol 67(1-2):83-93.
Daemen T, Riezebos-Brilman A, Regts J, Dontje B, van der Zee A, Wilschut J. 2004. Superior therapeutic
efficacy of alphavirus-mediated immunization against human papilloma virus type 16 antigens in a
murine tumour model: Effects of the route of immunization. Antivir Ther 9(5):733-42.
Davey GM, Heath WR, Starr R. 2006. SOCS1: A potent and multifaceted regulator of cytokines and cellmediated inflammation. Tissue Antigens 67(1):1-9.
De Angelis R, Sant M, Coleman MP, Francisci S, Baili P, Pierannunzio D, Trama A, Visser O, Brenner H,
Ardanaz E, et al. 2014. Cancer survival in europe 1999-2007 by country and age: Results of EUROCARE--5a population-based study. Lancet Oncol 15(1):23-34.
Dempoya J, Matsumiya T, Imaizumi T, Hayakari R, Xing F, Yoshida H, Okumura K, Satoh K. 2012. Doublestranded RNA induces biphasic STAT1 phosphorylation by both type I interferon (IFN)-dependent and
type I IFN-independent pathways. J Virol 86(23):12760-9.
Der SD, Zhou A, Williams BR, Silverman RH. 1998. Identification of genes differentially regulated by
interferon alpha, beta, or gamma using oligonucleotide arrays. Proc Natl Acad Sci U S A 95(26):15623-8.
Deuber SA and Pavlovic J. 2007. Virulence of a mouse-adapted semliki forest virus strain is associated with
reduced susceptibility to interferon. J Gen Virol 88(Pt 7):1952-9.
Devasthanam AS. 2014. Mechanisms underlying the inhibition of interferon signaling by viruses. Virulence
5(2):270-7.
Diamond MS and Farzan M. 2013. The broad-spectrum antiviral functions of IFIT and IFITM proteins. Nat
Rev Immunol 13(1):46-57.
Dias JD, Hemminki O, Diaconu I, Hirvinen M, Bonetti A, Guse K, Escutenaire S, Kanerva A, Pesonen S,
Loskog A, et al. 2012. Targeted cancer immunotherapy with oncolytic adenovirus coding for a fully human
monoclonal antibody specific for CTLA-4. Gene Ther 19(10):988-98.
Diaz RM, Galivo F, Kottke T, Wongthida P, Qiao J, Thompson J, Valdes M, Barber G, Vile RG. 2007. Oncolytic
immunovirotherapy for melanoma using vesicular stomatitis virus. Cancer Res 67(6):2840-8.
Dienz O, Rud JG, Eaton SM, Lanthier PA, Burg E, Drew A, Bunn J, Suratt BT, Haynes L, Rincon M. 2012.
Essential role of IL-6 in protection against H1N1 influenza virus by promoting neutrophil survival in the
lung. Mucosal Immunol 5(3):258-66.
Dmitriev I, Krasnykh V, Miller CR, Wang M, Kashentseva E, Mikheeva G, Belousova N, Curiel DT. 1998. An
adenovirus vector with genetically modified fibers demonstrates expanded tropism via utilization of a
coxsackievirus and adenovirus receptor-independent cell entry mechanism. J Virol 72(12):9706-13.
Dommett RM, Klein N, Turner MW. 2006. Mannose-binding lectin in innate immunity: Past, present and
future. Tissue Antigens 68(3):193-209.
Donnelly O, Harrington K, Melcher A, Pandha H. 2013a. Live viruses to treat cancer. J R Soc Med 106(8):310-4.
Donnelly OG, Errington-Mais F, Steele L, Hadac E, Jennings V, Scott K, Peach H, Phillips RM, Bond J, Pandha
H, et al. 2013b. Measles virus causes immunogenic cell death in human melanoma. Gene Ther 20(1):7-15.
Dorer DE and Nettelbeck DM. 2009. Targeting cancer by transcriptional control in cancer gene therapy and
viral oncolysis. Adv Drug Deliv Rev 61(7-8):554-71.
Dranoff G, Jaffee E, Lazenby A, Golumbek P, Levitsky H, Brose K, Jackson V, Hamada H, Pardoll D, Mulligan
RC. 1993. Vaccination with irradiated tumor cells engineered to secrete murine granulocyte-macrophage
colony-stimulating factor stimulates potent, specific, and long-lasting anti-tumor immunity. Proc Natl
Acad Sci U S A 90(8):3539-43.
Duan X, Nauwynck HJ, Pensaert MB. 1997. Effects of origin and state of differentiation and activation of
monocytes/macrophages on their susceptibility to porcine reproductive and respiratory syndrome virus
(PRRSV). Arch Virol 142(12):2483-97.
Dunn GP, Koebel CM, Schreiber RD. 2006. Interferons, immunity and cancer immunoediting. Nat Rev
Immunol 6(11):836-48.
DuPage M, Mazumdar C, Schmidt LM, Cheung AF, Jacks T. 2012. Expression of tumour-specific antigens
underlies cancer immunoediting. Nature 482(7385):405-9.
Dupuy LC, Richards MJ, Ellefsen B, Chau L, Luxembourg A, Hannaman D, Livingston BD, Schmaljohn CS.
2011. A DNA vaccine for venezuelan equine encephalitis virus delivered by intramuscular electroporation
elicits high levels of neutralizing antibodies in multiple animal models and provides protective immunity
to mice and nonhuman primates. Clin Vaccine Immunol 18(5):707-16.
89
Eisenstein S, Coakley BA, Briley-Saebo K, Ma G, Chen HM, Meseck M, Ward S, Divino C, Woo S, Chen SH, et
al. 2013. Myeloid-derived suppressor cells as a vehicle for tumor-specific oncolytic viral therapy. Cancer
Res 73(16):5003-15.
El Maadidi S, Faletti L, Berg B, Wenzl C, Wieland K, Chen ZJ, Maurer U, Borner C. 2014. A novel
mitochondrial MAVS/caspase-8 platform links RNA virus-induced innate antiviral signaling to bax/bakindependent apoptosis. J Immunol 192(3):1171-83.
Eng KE, Panas MD, Murphy D, Karlsson Hedestam GB, McInerney GM. 2012. Accumulation of
autophagosomes in semliki forest virus-infected cells is dependent on expression of the viral
glycoproteins. J Virol 86(10):5674-85.
Eto Y, Yoshioka Y, Mukai Y, Okada N, Nakagawa S. 2008. Development of PEGylated adenovirus vector with
targeting ligand. Int J Pharm 354(1-2):3-8.
Farmer JR, Altschaefl KM, O'Shea KS, Miller DJ. 2013. Activation of the type I interferon pathway is enhanced
in response to human neuronal differentiation. PLoS One 8(3):e58813.
Fazakerley JK. 2002. Pathogenesis of semliki forest virus encephalitis. J Neurovirol 8 Suppl 2:66-74.
Fazakerley JK and Webb HE. 1987. Cyclosporine enhances virally induced T-cell-mediated demyelination. the
effect of cyclosporine on a demyelinating virus infection. J Neurol Sci 78(1):35-50.
Fazakerley JK, Cotterill CL, Lee G, Graham A. 2006. Virus tropism, distribution, persistence and pathology in
the corpus callosum of the semliki forest virus-infected mouse brain: A novel system to study virusoligodendrocyte interactions. Neuropathol Appl Neurobiol 32(4):397-409.
Fazakerley JK, Pathak S, Scallan M, Amor S, Dyson H. 1993. Replication of the A7(74) strain of semliki forest
virus is restricted in neurons. Virology 195(2):627-37.
Fink SL and Cookson BT. 2005. Apoptosis, pyroptosis, and necrosis: Mechanistic description of dead and
dying eukaryotic cells. Infect Immun 73(4):1907-16.
Fisher KD and Seymour LW. 2010. HPMA copolymers for masking and retargeting of therapeutic viruses.
Adv Drug Deliv Rev 62(2):240-5.
Fleming P. 1977. Age-dependent and strain-related differences of virulence of semliki forest virus in mice. J
Gen Virol 37(1):93-105.
Forbes NE, Abdelbary H, Lupien M, Bell JC, Diallo JS. 2013. Exploiting tumor epigenetics to improve
oncolytic virotherapy. Front Genet 4:184.
Fragkoudis R, Breakwell L, McKimmie C, Boyd A, Barry G, Kohl A, Merits A, Fazakerley JK. 2007. The type I
interferon system protects mice from semliki forest virus by preventing widespread virus dissemination in
extraneural tissues, but does not mediate the restricted replication of avirulent virus in central nervous
system neurons. J Gen Virol 88(Pt 12):3373-84.
Francisco LM, Sage PT, Sharpe AH. 2010. The PD-1 pathway in tolerance and autoimmunity. Immunol Rev
236:219-42.
Freytag SO, Khil M, Stricker H, Peabody J, Menon M, DePeralta-Venturina M, Nafziger D, Pegg J, Paielli D,
Brown S, et al. 2002. Phase I study of replication-competent adenovirus-mediated double suicide gene
therapy for the treatment of locally recurrent prostate cancer. Cancer Res 62(17):4968-76.
Frieman MB, Chen J, Morrison TE, Whitmore A, Funkhouser W, Ward JM, Lamirande EW, Roberts A, Heise
M, Subbarao K, et al. 2010. SARS-CoV pathogenesis is regulated by a STAT1 dependent but a type I, II and
III interferon receptor independent mechanism. PLoS Pathog 6(4):e1000849.
Froelich S, Tai A, Kennedy K, Zubair A, Wang P. 2011. Virus-receptor mediated transduction of dendritic cells
by lentiviruses enveloped with glycoproteins derived from semliki forest virus. PLoS One 6(6):e21491.
Fros JJ, van der Maten E, Vlak JM, Pijlman GP. 2013. The C-terminal domain of chikungunya virus nsP2
independently governs viral RNA replication, cytopathicity, and inhibition of interferon signaling. J Virol
87(18):10394-400.
Fros JJ, Liu WJ, Prow NA, Geertsema C, Ligtenberg M, Vanlandingham DL, Schnettler E, Vlak JM, Suhrbier A,
Khromykh AA, et al. 2010. Chikungunya virus nonstructural protein 2 inhibits type I/II interferonstimulated JAK-STAT signaling. J Virol 84(20):10877-87.
Fu X, Rivera A, Tao L, Zhang X. 2012. Incorporation of the B18R gene of vaccinia virus into an oncolytic
herpes simplex virus improves antitumor activity. Mol Ther 20(10):1871-81.
Fulci G, Breymann L, Gianni D, Kurozomi K, Rhee SS, Yu J, Kaur B, Louis DN, Weissleder R, Caligiuri MA, et
al. 2006. Cyclophosphamide enhances glioma virotherapy by inhibiting innate immune responses. Proc
Natl Acad Sci U S A 103(34):12873-8.
90
Fulci G, Dmitrieva N, Gianni D, Fontana EJ, Pan X, Lu Y, Kaufman CS, Kaur B, Lawler SE, Lee RJ, et al. 2007.
Depletion of peripheral macrophages and brain microglia increases brain tumor titers of oncolytic viruses.
Cancer Res 67(19):9398-406.
Fung KY, Mangan NE, Cumming H, Horvat JC, Mayall JR, Stifter SA, De Weerd N, Roisman LC, Rossjohn J,
Robertson SA, et al. 2013. Interferon-epsilon protects the female reproductive tract from viral and bacterial
infection. Science 339(6123):1088-92.
Galivo F, Diaz RM, Wongthida P, Thompson J, Kottke T, Barber G, Melcher A, Vile R. 2010. Single-cycle viral
gene expression, rather than progressive replication and oncolysis, is required for VSV therapy of B16
melanoma. Gene Ther 17(2):158-70.
Ganesh S, Gonzalez-Edick M, Gibbons D, Van Roey M, Jooss K. 2008. Intratumoral coadministration of
hyaluronidase enzyme and oncolytic adenoviruses enhances virus potency in metastatic tumor models.
Clin Cancer Res 14(12):3933-41.
Garcia-Castro J, Alemany R, Cascallo M, Martinez-Quintanilla J, Arriero Mdel M, Lassaletta A, Madero L,
Ramirez M. 2010. Treatment of metastatic neuroblastoma with systemic oncolytic virotherapy delivered by
autologous mesenchymal stem cells: An exploratory study. Cancer Gene Ther 17(7):476-83.
Garmashova N, Gorchakov R, Volkova E, Paessler S, Frolova E, Frolov I. 2007. The old world and new world
alphaviruses use different virus-specific proteins for induction of transcriptional shutoff. J Virol 81(5):247284.
Gil MP, Bohn E, O'Guin AK, Ramana CV, Levine B, Stark GR, Virgin HW, Schreiber RD. 2001. Biologic
consequences of Stat1-independent IFN signaling. Proc Natl Acad Sci U S A 98(12):6680-5.
Ginn SL, Alexander IE, Edelstein ML, Abedi MR, Wixon J. 2013. Gene therapy clinical trials worldwide to
2012 - an update. J Gene Med 15(2):65-77.
Glasgow GM, McGee MM, Sheahan BJ, Atkins GJ. 1997. Death mechanisms in cultured cells infected by
semliki forest virus. J Gen Virol 78 ( Pt 7)(Pt 7):1559-63.
Glasgow GM, McGee MM, Tarbatt CJ, Mooney DA, Sheahan BJ, Atkins GJ. 1998. The semliki forest virus
vector induces p53-independent apoptosis. J Gen Virol 79 ( Pt 10)(Pt 10):2405-10.
Goetz C and Gromeier M. 2010. Preparing an oncolytic poliovirus recombinant for clinical application against
glioblastoma multiforme. Cytokine Growth Factor Rev 21(2-3):197-203.
Safety and Immunogenicity of Venezuelan Equine Encephalomyelitis Vaccine (VEE C-84) as a Booster to VEE
TC-83 [Internet]; cMarch 6, 2014 [cited 2014 11th of June]. Available
from: http://www.clinicaltrials.gov/ct2/show/NCT00582088 .
Goldberg SM, Bartido SM, Gardner JP, Guevara-Patino JA, Montgomery SC, Perales MA, Maughan MF,
Dempsey J, Donovan GP, Olson WC, et al. 2005. Comparison of two cancer vaccines targeting tyrosinase:
Plasmid DNA and recombinant alphavirus replicon particles. Clin Cancer Res 11(22):8114-21.
Goubau D, Deddouche S, Reis e Sousa C. 2013. Cytosolic sensing of viruses. Immunity 38(5):855-69.
Gough DJ, Messina NL, Hii L, Gould JA, Sabapathy K, Robertson AP, Trapani JA, Levy DE, Hertzog PJ,
Clarke CJ, et al. 2010. Functional crosstalk between type I and II interferon through the regulated
expression of STAT1. PLoS Biol 8(4):e1000361.
Grabowska AK and Riemer AB. 2012. The invisible enemy - how human papillomaviruses avoid recognition
and clearance by the host immune system. Open Virol J 6:249-56.
Grandvaux N, Servant MJ, tenOever B, Sen GC, Balachandran S, Barber GN, Lin R, Hiscott J. 2002.
Transcriptional profiling of interferon regulatory factor 3 target genes: Direct involvement in the
regulation of interferon-stimulated genes. J Virol 76(11):5532-9.
Granot T, Yamanashi Y, Meruelo D. 2014. Sindbis viral vectors transiently deliver tumor-associated antigens
to lymph nodes and elicit diversified antitumor CD8+ T-cell immunity. Mol Ther 22(1):112-22.
Grieder FB and Vogel SN. 1999. Role of interferon and interferon regulatory factors in early protection against
venezuelan equine encephalitis virus infection. Virology 257(1):106-18.
Griffin D, Levine B, Tyor W, Ubol S, Despres P. 1997. The role of antibody in recovery from alphavirus
encephalitis. Immunol Rev 159:155-61.
Griffin DE. 2005. Neuronal cell death in alphavirus encephalomyelitis. Curr Top Microbiol Immunol 289:5777.
Griffith TS and Ferguson TA. 2011. Cell death in the maintenance and abrogation of tolerance: The five ws of
dying cells. Immunity 35(4):456-66.
91
Guan M, Rodriguez-Madoz JR, Alzuguren P, Gomar C, Kramer MG, Kochanek S, Prieto J, Smerdou C, Qian
C. 2006. Increased efficacy and safety in the treatment of experimental liver cancer with a novel
adenovirus-alphavirus hybrid vector. Cancer Res 66(3):1620-9.
Gurlevik E, Woller N, Struver N, Schache P, Kloos A, Manns MP, Zender L, Kuhnel F, Kubicka S. 2010.
Selectivity of oncolytic viral replication prevents antiviral immune response and toxicity, but does not
improve antitumoral immunity. Mol Ther 18(11):1972-82.
Hacein-Bey-Abina S, Garrigue A, Wang GP, Soulier J, Lim A, Morillon E, Clappier E, Caccavelli L, Delabesse
E, Beldjord K, et al. 2008. Insertional oncogenesis in 4 patients after retrovirus-mediated gene therapy of
SCID-X1. J Clin Invest 118(9):3132-42.
Haisma HJ and Bellu AR. 2011. Pharmacological interventions for improving adenovirus usage in gene
therapy. Mol Pharm 8(1):50-5.
Hanahan D and Weinberg RA. 2011. Hallmarks of cancer: The next generation. Cell 144(5):646-74.
Hefti HP, Frese M, Landis H, Di Paolo C, Aguzzi A, Haller O, Pavlovic J. 1999. Human MxA protein protects
mice lacking a functional alpha/beta interferon system against la crosse virus and other lethal viral
infections. J Virol 73(8):6984-91.
Helenius A, Kartenbeck J, Simons K, Fries E. 1980. On the entry of semliki forest virus into BHK-21 cells. J Cell
Biol 84(2):404-20.
Heo J, Reid T, Ruo L, Breitbach CJ, Rose S, Bloomston M, Cho M, Lim HY, Chung HC, Kim CW, et al. 2013.
Randomized dose-finding clinical trial of oncolytic immunotherapeutic vaccinia JX-594 in liver cancer. Nat
Med 19(3):329-36.
Hervas-Stubbs S, Perez-Gracia JL, Rouzaut A, Sanmamed MF, Le Bon A, Melero I. 2011. Direct effects of type I
interferons on cells of the immune system. Clin Cancer Res 17(9):2619-27.
Hochrein H and Wagner H. 2004. Of men, mice and pigs: Looking at their plasmacytoid dendritic cells
[corrected. Immunology 112(1):26-7.
Hodi Stephen. 2013. The 2013 european cancer congress (ECC 2013). ECC 2013 press release: Advanced
melanoma: Pooled analysis of long-term survival with ipilimumab; 28th September 2013; Amsterdam,
Netherlands. .
Holzel M, Bovier A, Tuting T. 2013. Plasticity of tumour and immune cells: A source of heterogeneity and a
cause for therapy resistance? Nat Rev Cancer 13(5):365-76.
Holzinger D, Jorns C, Stertz S, Boisson-Dupuis S, Thimme R, Weidmann M, Casanova JL, Haller O, Kochs G.
2007. Induction of MxA gene expression by influenza A virus requires type I or type III interferon
signaling. J Virol 81(14):7776-85.
Honda K and Taniguchi T. 2006. IRFs: Master regulators of signalling by toll-like receptors and cytosolic
pattern-recognition receptors. Nat Rev Immunol 6(9):644-58.
Hong DS, Angelo LS, Kurzrock R. 2007. Interleukin-6 and its receptor in cancer: Implications for translational
therapeutics. Cancer 110(9):1911-28.
Howe JA, Mymryk JS, Egan C, Branton PE, Bayley ST. 1990. Retinoblastoma growth suppressor and a 300kDa protein appear to regulate cellular DNA synthesis. Proc Natl Acad Sci U S A 87(15):5883-7.
Howe SJ, Mansour MR, Schwarzwaelder K, Bartholomae C, Hubank M, Kempski H, Brugman MH, PikeOverzet K, Chatters SJ, de Ridder D, et al. 2008. Insertional mutagenesis combined with acquired somatic
mutations causes leukemogenesis following gene therapy of SCID-X1 patients. J Clin Invest 118(9):3143-50.
Hsu AC, Parsons K, Barr I, Lowther S, Middleton D, Hansbro PM, Wark PA. 2012. Critical role of constitutive
type I interferon response in bronchial epithelial cell to influenza infection. PLoS One 7(3):e32947.
Huang J, Jochems C, Talaie T, Anderson A, Jales A, Tsang KY, Madan RA, Gulley JL, Schlom J. 2012. Elevated
serum soluble CD40 ligand in cancer patients may play an immunosuppressive role. Blood 120(15):3030-8.
Huang PY, Guo JH, Hwang LH. 2012. Oncolytic sindbis virus targets tumors defective in the interferon
response and induces significant bystander antitumor immunity in vivo. Mol Ther 20(2):298-305.
Hubackova S, Krejcikova K, Bartek J, Hodny Z. 2012. Interleukin 6 signaling regulates promyelocytic
leukemia protein gene expression in human normal and cancer cells. J Biol Chem 287(32):26702-14.
Huotari J and Helenius A. 2011. Endosome maturation. EMBO J 30(17):3481-500.
Huyan XH, Lin YP, Gao T, Chen RY, Fan YM. 2011. Immunosuppressive effect of cyclophosphamide on white
blood cells and lymphocyte subpopulations from peripheral blood of balb/c mice. Int Immunopharmacol
11(9):1293-7.
Hyde JL, Gardner CL, Kimura T, White JP, Liu G, Trobaugh DW, Huang C, Tonelli M, Paessler S, Takeda K, et
al. 2014. A viral RNA structural element alters host recognition of nonself RNA. Science 343(6172):783-7.
92
Ida-Hosonuma M, Iwasaki T, Yoshikawa T, Nagata N, Sato Y, Sata T, Yoneyama M, Fujita T, Taya C,
Yonekawa H, et al. 2005. The alpha/beta interferon response controls tissue tropism and pathogenicity of
poliovirus. J Virol 79(7):4460-9.
Iijima Y, Ohno K, Ikeda H, Sawai K, Levin B, Meruelo D. 1999. Cell-specific targeting of a thymidine
kinase/ganciclovir gene therapy system using a recombinant sindbis virus vector. Int J Cancer 80(1):110-8.
Ikeda K, Wakimoto H, Ichikawa T, Jhung S, Hochberg FH, Louis DN, Chiocca EA. 2000. Complement
depletion facilitates the infection of multiple brain tumors by an intravascular, replication-conditional
herpes simplex virus mutant. J Virol 74(10):4765-75.
Ikeda K, Ichikawa T, Wakimoto H, Silver JS, Deisboeck TS, Finkelstein D, Harsh GR,4th, Louis DN, Bartus RT,
Hochberg FH, et al. 1999. Oncolytic virus therapy of multiple tumors in the brain requires suppression of
innate and elicited antiviral responses. Nat Med 5(8):881-7.
Ilett EJ, Prestwich RJ, Kottke T, Errington F, Thompson JM, Harrington KJ, Pandha HS, Coffey M, Selby PJ,
Vile RG, et al. 2009. Dendritic cells and T cells deliver oncolytic reovirus for tumour killing despite preexisting anti-viral immunity. Gene Ther 16(5):689-99.
Iordanova B, Goins WF, Clawson DS, Hitchens TK, Ahrens ET. 2013. Quantification of HSV-1-mediated
expression of the ferritin MRI reporter in the mouse brain. Gene Ther 20(6):589-96.
Jackson RJ, Ramsay AJ, Christensen CD, Beaton S, Hall DF, Ramshaw IA. 2001. Expression of mouse
interleukin-4 by a recombinant ectromelia virus suppresses cytolytic lymphocyte responses and
overcomes genetic resistance to mousepox. J Virol 75(3):1205-10.
Jha BK, Dong B, Nguyen CT, Polyakova I, Silverman RH. 2013. Suppression of antiviral innate immunity by
sunitinib enhances oncolytic virotherapy. Mol Ther 21(9):1749-57.
Jiang Y, Oliver P, Davies KE, Platt N. 2006. Identification and characterization of murine SCARA5, a novel
class A scavenger receptor that is expressed by populations of epithelial cells. J Biol Chem 281(17):1183445.
Jin HK, Takada A, Kon Y, Haller O, Watanabe T. 1999. Identification of the murine Mx2 gene: Interferoninduced expression of the Mx2 protein from the feral mouse gene confers resistance to vesicular stomatitis
virus. J Virol 73(6):4925-30.
John LB, Howland LJ, Flynn JK, West AC, Devaud C, Duong CP, Stewart TJ, Westwood JA, Guo ZS, Bartlett
DL, et al. 2012. Oncolytic virus and anti-4-1BB combination therapy elicits strong antitumor immunity
against established cancer. Cancer Res 72(7):1651-60.
Jose J, Snyder JE, Kuhn RJ. 2009. A structural and functional perspective of alphavirus replication and
assembly. Future Microbiol 4(7):837-56.
Juang YT, Lowther W, Kellum M, Au WC, Lin R, Hiscott J, Pitha PM. 1998. Primary activation of interferon A
and interferon B gene transcription by interferon regulatory factor 3. Proc Natl Acad Sci U S A 95(17):983742.
Kamrud KI, Coffield VM, Owens G, Goodman C, Alterson K, Custer M, Murphy MA, Lewis W, Timberlake S,
Wansley EK, et al. 2010. In vitro and in vivo characterization of microRNA-targeted alphavirus replicon
and helper RNAs. J Virol 84(15):7713-25.
Karapanagiotou EM, Roulstone V, Twigger K, Ball M, Tanay M, Nutting C, Newbold K, Gore ME, Larkin J,
Syrigos KN, et al. 2012. Phase I/II trial of carboplatin and paclitaxel chemotherapy in combination with
intravenous oncolytic reovirus in patients with advanced malignancies. Clin Cancer Res 18(7):2080-9.
Karki S, Li MM, Schoggins JW, Tian S, Rice CM, MacDonald MR. 2012. Multiple interferon stimulated genes
synergize with the zinc finger antiviral protein to mediate anti-alphavirus activity. PLoS One 7(5):e37398.
Kaur S and Platanias LC. 2013. IFN-beta-specific signaling via a unique IFNAR1 interaction. Nat Immunol
14(9):884-5.
Keogh B, Atkins GJ, Mills KH, Sheahan BJ. 2003. Role of interferon-gamma and nitric oxide in the
neuropathogenesis of avirulent semliki forest virus infection. Neuropathol Appl Neurobiol 29(6):553-62.
Kershaw MH, Hsu C, Mondesire W, Parker LL, Wang G, Overwijk WW, Lapointe R, Yang JC, Wang RF,
Restifo NP, et al. 2001. Immunization against endogenous retroviral tumor-associated antigens. Cancer Res
61(21):7920-4.
Ketola A, Hinkkanen A, Yongabi F, Furu P, Maatta AM, Liimatainen T, Pirinen R, Bjorn M, Hakkarainen T,
Makinen K, et al. 2008. Oncolytic semliki forest virus vector as a novel candidate against unresectable
osteosarcoma. Cancer Res 68(20):8342-50.
Khodarev NN, Roizman B, Weichselbaum RR. 2012. Molecular pathways: Interferon/stat1 pathway: Role in
the tumor resistance to genotoxic stress and aggressive growth. Clin Cancer Res 18(11):3015-21.
93
Khodarev NN, Beckett M, Labay E, Darga T, Roizman B, Weichselbaum RR. 2004. STAT1 is overexpressed in
tumors selected for radioresistance and confers protection from radiation in transduced sensitive cells.
Proc Natl Acad Sci U S A 101(6):1714-9.
Kielian MC, Marsh M, Helenius A. 1986. Kinetics of endosome acidification detected by mutant and wild-type
semliki forest virus. EMBO J 5(12):3103-9.
Kim CK, Ahmed AU, Auffinger B, Ulasov IV, Tobias AL, Moon KS, Lesniak MS. 2013. N-acetylcysteine amide
augments the therapeutic effect of neural stem cell-based antiglioma oncolytic virotherapy. Mol Ther
21(11):2063-73.
Kim JH, Lee YS, Kim H, Huang JH, Yoon AR, Yun CO. 2006. Relaxin expression from tumor-targeting
adenoviruses and its intratumoral spread, apoptosis induction, and efficacy. J Natl Cancer Inst 98(20):148293.
Kim M, Sumerel LA, Belousova N, Lyons GR, Carey DE, Krasnykh V, Douglas JT. 2003. The coxsackievirus
and adenovirus receptor acts as a tumour suppressor in malignant glioma cells. Br J Cancer 88(9):1411-6.
Kim TW, Hung CF, Juang J, He L, Hardwick JM, Wu TC. 2004. Enhancement of suicidal DNA vaccine potency
by delaying suicidal DNA-induced cell death. Gene Ther 11(3):336-42.
Kirn D, Martuza RL, Zwiebel J. 2001. Replication-selective virotherapy for cancer: Biological principles, risk
management and future directions. Nat Med 7(7):781-7.
Klimstra WB, Nangle EM, Smith MS, Yurochko AD, Ryman KD. 2003. DC-SIGN and L-SIGN can act as
attachment receptors for alphaviruses and distinguish between mosquito cell- and mammalian cellderived viruses. J Virol 77(22):12022-32.
Knudsen ML, Johansson DX, Kostic L, Nordstrom EK, Tegerstedt K, Pasetto A, Applequist SE, Ljungberg K,
Sirard JC, Liljestrom P. 2013. The adjuvant activity of alphavirus replicons is enhanced by incorporating
the microbial molecule flagellin into the replicon. PLoS One 8(6):e65964.
Kohanbash G, McKaveney K, Sakaki M, Ueda R, Mintz AH, Amankulor N, Fujita M, Ohlfest JR, Okada H.
2013. GM-CSF promotes the immunosuppressive activity of glioma-infiltrating myeloid cells through
interleukin-4 receptor-alpha. Cancer Res 73(21):6413-23.
Kokernot RH, Casaca VM, Weinbren MP, McIntosh BM. 1965. Survey for antibodies against arthropod-borne
viruses in the sera of indigenous residents of angola. Trans R Soc Trop Med Hyg 59(5):563-70.
Koski A, Rajecki M, Guse K, Kanerva A, Ristimaki A, Pesonen S, Escutenaire S, Hemminki A. 2009. Systemic
adenoviral gene delivery to orthotopic murine breast tumors with ablation of coagulation factors,
thrombocytes and kupffer cells. J Gene Med 11(11):966-77.
Koski A, Kangasniemi L, Escutenaire S, Pesonen S, Cerullo V, Diaconu I, Nokisalmi P, Raki M, Rajecki M,
Guse K, et al. 2010. Treatment of cancer patients with a serotype 5/3 chimeric oncolytic adenovirus
expressing GMCSF. Mol Ther 18(10):1874-84.
Kottke T, Hall G, Pulido J, Diaz RM, Thompson J, Chong H, Selby P, Coffey M, Pandha H, Chester J, et al.
2010. Antiangiogenic cancer therapy combined with oncolytic virotherapy leads to regression of
established tumors in mice. J Clin Invest 120(5):1551-60.
Kottke T, Errington F, Pulido J, Galivo F, Thompson J, Wongthida P, Diaz RM, Chong H, Ilett E, Chester J, et
al. 2011. Broad antigenic coverage induced by vaccination with virus-based cDNA libraries cures
established tumors. Nat Med 17(7):854-9.
Kuhn I, Harden P, Bauzon M, Chartier C, Nye J, Thorne S, Reid T, Ni S, Lieber A, Fisher K, et al. 2008.
Directed evolution generates a novel oncolytic virus for the treatment of colon cancer. PLoS One
3(6):e2409.
Kundig TM, Hengartner H, Zinkernagel RM. 1993. T cell-dependent IFN-gamma exerts an antiviral effect in
the central nervous system but not in peripheral solid organs. J Immunol 150(6):2316-21.
Lacki JK, Mackiewicz SH, Leszczynski P, Muller W. 1997. The effect of intravenous cyclophosphamide pulse
on peripheral blood lymphocytes in lupus erythematosus patients. Rheumatol Int 17(2):55-60.
Lam E, Stein S, Falck-Pedersen E. 2014. Adenovirus detection by the cGAS/STING/TBK1 DNA sensing
cascade. J Virol 88(2):974-81.
Lam P, Khan G, Stripecke R, Hui KM, Kasahara N, Peng KW, Guinn BA. 2013. The innovative evolution of
cancer gene and cellular therapies. Cancer Gene Ther 20(3):141-9.
Laule C, Vavasour IM, Moore GR, Oger J, Li DK, Paty DW, MacKay AL. 2004. Water content and myelin
water fraction in multiple sclerosis. A T2 relaxation study. J Neurol 251(3):284-93.
94
Le Boeuf F, Niknejad N, Wang J, Auer R, Weberpals JI, Bell JC, Dimitroulakos J. 2012. Sensitivity of cervical
carcinoma cells to vesicular stomatitis virus-induced oncolysis: Potential role of human papilloma virus
infection. Int J Cancer 131(3):E204-15.
Le Boeuf F, Diallo JS, McCart JA, Thorne S, Falls T, Stanford M, Kanji F, Auer R, Brown CW, Lichty BD, et al.
2010. Synergistic interaction between oncolytic viruses augments tumor killing. Mol Ther 18(5):888-95.
Le Boeuf F, Batenchuk C, Vaha-Koskela M, Breton S, Roy D, Lemay C, Cox J, Abdelbary H, Falls T, Waghray
G, et al. 2013. Model-based rational design of an oncolytic virus with improved therapeutic potential. Nat
Commun 4:1974.
Le Mercier I, Chen W, Lines JL, Day M, Li J, Sergent P, Noelle RJ, Wang L. 2014. VISTA regulates the
development of protective antitumor immunity. Cancer Res 74(7):1933-44.
Lee EY, Schultz KL, Griffin DE. 2013. Mice deficient in interferon-gamma or interferon-gamma receptor 1
have distinct inflammatory responses to acute viral encephalomyelitis. PLoS One 8(10):e76412.
Lee JY, Myung SK, Song YS. 2013. Prognostic role of cyclooxygenase-2 in epithelial ovarian cancer: A metaanalysis of observational studies. Gynecol Oncol 129(3):613-9.
Leib DA, Harrison TE, Laslo KM, Machalek MA, Moorman NJ, Virgin HW. 1999. Interferons regulate the
phenotype of wild-type and mutant herpes simplex viruses in vivo. J Exp Med 189(4):663-72.
Leitner WW, Bergmann-Leitner ES, Hwang LN, Restifo NP. 2006. Type I interferons are essential for the
efficacy of replicase-based DNA vaccines. Vaccine 24(24):5110-8.
Leitner WW, Hwang LN, Bergmann-Leitner ES, Finkelstein SE, Frank S, Restifo NP. 2004. Apoptosis is
essential for the increased efficacy of alphaviral replicase-based DNA vaccines. Vaccine 22(11-12):1537-44.
Leitner WW, Hwang LN, deVeer MJ, Zhou A, Silverman RH, Williams BR, Dubensky TW, Ying H, Restifo
NP. 2003. Alphavirus-based DNA vaccine breaks immunological tolerance by activating innate antiviral
pathways. Nat Med 9(1):33-9.
Levin D, Harari D, Schreiber G. 2011. Stochastic receptor expression determines cell fate upon interferon
treatment. Mol Cell Biol 31(16):3252-66.
Li S, Wilkinson M, Xia X, David M, Xu L, Purkel-Sutton A, Bhardwaj A. 2005. Induction of IFN-regulated
factors and antitumoral surveillance by transfected placebo plasmid DNA. Mol Ther 11(1):112-9.
Lienenklaus S, Walisko R, te Boekhorst A, May T, Samuelsson C, Michiels T, Weiss S. 2008. PCR-based
simultaneous analysis of the interferon-alpha family reveals distinct kinetics for early interferons. J
Interferon Cytokine Res 28(11):653-60.
Lines JL, Pantazi E, Mak J, Sempere LF, Wang L, O'Connell S, Ceeraz S, Suriawinata AA, Yan S, Ernstoff MS,
et al. 2014. VISTA is an immune checkpoint molecule for human T cells. Cancer Res 74(7):1924-32.
Liu C, Russell SJ, Peng KW. 2010. Systemic therapy of disseminated myeloma in passively immunized mice
using measles virus-infected cell carriers. Mol Ther 18(6):1155-64.
Liu SY, Sanchez DJ, Aliyari R, Lu S, Cheng G. 2012. Systematic identification of type I and type II interferoninduced antiviral factors. Proc Natl Acad Sci U S A 109(11):4239-44.
Liu TC, Galanis E, Kirn D. 2007. Clinical trial results with oncolytic virotherapy: A century of promise, a
decade of progress. Nat Clin Pract Oncol 4(2):101-17.
Liu YP, Suksanpaisan L, Steele MB, Russell SJ, Peng KW. 2013. Induction of antiviral genes by the tumor
microenvironment confers resistance to virotherapy. Sci Rep 3:2375.
Ljungberg K, Whitmore AC, Fluet ME, Moran TP, Shabman RS, Collier ML, Kraus AA, Thompson JM,
Montefiori DC, Beard C, et al. 2007. Increased immunogenicity of a DNA-launched venezuelan equine
encephalitis virus-based replicon DNA vaccine. J Virol 81(24):13412-23.
Lo Cicero A, Schiera G, Proia P, Saladino P, Savettieri G, Di Liegro CM, Di Liegro I. 2011. Oligodendroglioma
cells shed microvesicles which contain TRAIL as well as molecular chaperones and induce cell death in
astrocytes. Int J Oncol 39(6):1353-7.
Lovett-Racke AE, Smith ME, Arredondo LR, Bittner PS, Ratts RB, Shive CL, Forsthuber TG, Racke MK. 2000.
Developmentally regulated gene expression of Th2 cytokines in the brain. Brain Res 870(1-2):27-35.
Lun X, Alain T, Zemp FJ, Zhou H, Rahman MM, Hamilton MG, McFadden G, Bell J, Senger DL, Forsyth PA.
2010. Myxoma virus virotherapy for glioma in immunocompetent animal models: Optimizing
administration routes and synergy with rapamycin. Cancer Res 70(2):598-608.
Lun X, Yang W, Alain T, Shi ZQ, Muzik H, Barrett JW, McFadden G, Bell J, Hamilton MG, Senger DL, et al.
2005. Myxoma virus is a novel oncolytic virus with significant antitumor activity against experimental
human gliomas. Cancer Res 65(21):9982-90.
95
Lun XQ, Jang JH, Tang N, Deng H, Head R, Bell JC, Stojdl DF, Nutt CL, Senger DL, Forsyth PA, et al. 2009.
Efficacy of systemically administered oncolytic vaccinia virotherapy for malignant gliomas is enhanced by
combination therapy with rapamycin or cyclophosphamide. Clin Cancer Res 15(8):2777-88.
Lun XQ, Zhou H, Alain T, Sun B, Wang L, Barrett JW, Stanford MM, McFadden G, Bell J, Senger DL, et al.
2007. Targeting human medulloblastoma: Oncolytic virotherapy with myxoma virus is enhanced by
rapamycin. Cancer Res 67(18):8818-27.
Maatta AM, Samaranayake H, Pikkarainen J, Wirth T, Yla-Herttuala S. 2009. Adenovirus mediated herpes
simplex virus-thymidine kinase/ganciclovir gene therapy for resectable malignant glioma. Curr Gene Ther
9(5):356-67.
Maatta AM, Makinen K, Ketola A, Liimatainen T, Yongabi FN, Vaha-Koskela M, Pirinen R, Rautsi O, Pellinen
R, Hinkkanen A, et al. 2008. Replication competent semliki forest virus prolongs survival in experimental
lung cancer. Int J Cancer 123(7):1704-11.
Maatta AM, Liimatainen T, Wahlfors T, Wirth T, Vaha-Koskela M, Jansson L, Valonen P, Hakkinen K, Rautsi
O, Pellinen R, et al. 2007. Evaluation of cancer virotherapy with attenuated replicative semliki forest virus
in different rodent tumor models. Int J Cancer 121(4):863-70.
Mabruk MJ, Glasglow GM, Flack AM, Folan JC, Bannigan JG, Smyth JM, O'Sullivan MA, Sheahan BJ, Atkins
GJ. 1989. Effect of infection with the ts22 mutant of semliki forest virus on development of the central
nervous system in the fetal mouse. J Virol 63(9):4027-33.
MacDonald GH and Johnston RE. 2000. Role of dendritic cell targeting in venezuelan equine encephalitis
virus pathogenesis. J Virol 74(2):914-22.
MacDonald MR, Machlin ES, Albin OR, Levy DE. 2007. The zinc finger antiviral protein acts synergistically
with an interferon-induced factor for maximal activity against alphaviruses. J Virol 81(24):13509-18.
MacMicking JD. 2012. Interferon-inducible effector mechanisms in cell-autonomous immunity. Nat Rev
Immunol 12(5):367-82.
Magge D, Guo ZS, O'Malley ME, Francis L, Ravindranathan R, Bartlett DL. 2013. Inhibitors of C5 complement
enhance vaccinia virus oncolysis. Cancer Gene Ther 20(6):342-50.
Mahonen AJ, Makkonen KE, Laakkonen JP, Ihalainen TO, Kukkonen SP, Kaikkonen MU, Vihinen-Ranta M,
Yla-Herttuala S, Airenne KJ. 2010. Culture medium induced vimentin reorganization associates with
enhanced baculovirus-mediated gene delivery. J Biotechnol 145(2):111-9.
Mak T and Saunders M. 2011. Primer to the immune response: Academic cell update edition (google eBook).
Academic Press.
Malilas W, Koh SS, Srisuttee R, Boonying W, Cho IR, Jeong CS, Johnston RN, Chung YH. 2013. Cancer
upregulated gene 2, a novel oncogene, confers resistance to oncolytic vesicular stomatitis virus through
STAT1-OASL2 signaling. Cancer Gene Ther 20(2):125-32.
Mansour M, Palese P, Zamarin D. 2011. Oncolytic specificity of newcastle disease virus is mediated by
selectivity for apoptosis-resistant cells. J Virol 85(12):6015-23.
Mantovani A, Sica A, Allavena P, Garlanda C, Locati M. 2009. Tumor-associated macrophages and the related
myeloid-derived suppressor cells as a paradigm of the diversity of macrophage activation. Hum Immunol
70(5):325-30.
Martin ME and Berk AJ. 1999. Corepressor required for adenovirus E1B 55,000-molecular-weight protein
repression of basal transcription. Mol Cell Biol 19(5):3403-14.
Maruggi G, Shaw CA, Otten GR, Mason PW, Beard CW. 2013. Engineered alphavirus replicon vaccines based
on known attenuated viral mutants show limited effects on immunogenicity. Virology 447(1-2):254-64.
Mastrangelo MJ, Maguire HC,Jr, Eisenlohr LC, Laughlin CE, Monken CE, McCue PA, Kovatich AJ, Lattime
EC. 1999. Intratumoral recombinant GM-CSF-encoding virus as gene therapy in patients with cutaneous
melanoma. Cancer Gene Ther 6(5):409-22.
Mathiot CC, Grimaud G, Garry P, Bouquety JC, Mada A, Daguisy AM, Georges AJ. 1990. An outbreak of
human semliki forest virus infections in central african republic. Am J Trop Med Hyg 42(4):386-93.
Matrai J, Chuah MK, VandenDriessche T. 2010. Recent advances in lentiviral vector development and
applications. Mol Ther 18(3):477-90.
Matsushita H, Vesely MD, Koboldt DC, Rickert CG, Uppaluri R, Magrini VJ, Arthur CD, White JM, Chen YS,
Shea LK, et al. 2012. Cancer exome analysis reveals a T-cell-dependent mechanism of cancer
immunoediting. Nature 482(7385):400-4.
96
McCart JA, Ward JM, Lee J, Hu Y, Alexander HR, Libutti SK, Moss B, Bartlett DL. 2001. Systemic cancer
therapy with a tumor-selective vaccinia virus mutant lacking thymidine kinase and vaccinia growth factor
genes. Cancer Res 61(24):8751-7.
McCart JA, Mehta N, Scollard D, Reilly RM, Carrasquillo JA, Tang N, Deng H, Miller M, Xu H, Libutti SK, et
al. 2004. Oncolytic vaccinia virus expressing the human somatostatin receptor SSTR2: Molecular imaging
after systemic delivery using 111In-pentetreotide. Mol Ther 10(3):553-61.
McFadden G, Mohamed MR, Rahman MM, Bartee E. 2009. Cytokine determinants of viral tropism. Nat Rev
Immunol 9(9):645-55.
Meerts P, Misinzo G, Nauwynck HJ. 2005. Enhancement of porcine circovirus 2 replication in porcine cell
lines by IFN-gamma before and after treatment and by IFN-alpha after treatment. J Interferon Cytokine
Res 25(11):684-93.
Mellman I, Coukos G, Dranoff G. 2011. Cancer immunotherapy comes of age. Nature 480(7378):480-9.
Metcalf TU and Griffin DE. 2011. Alphavirus-induced encephalomyelitis: Antibody-secreting cells and viral
clearance from the nervous system. J Virol 85(21):11490-501.
Metz SW, Feenstra F, Villoing S, van Hulten MC, van Lent JW, Koumans J, Vlak JM, Pijlman GP. 2011. Low
temperature-dependent salmonid alphavirus glycoprotein processing and recombinant virus-like particle
formation. PLoS One 6(10):e25816.
Miest TS and Cattaneo R. 2014. New viruses for cancer therapy: Meeting clinical needs. Nat Rev Microbiol
12(1):23-34.
Miyamoto S, Inoue H, Nakamura T, Yamada M, Sakamoto C, Urata Y, Okazaki T, Marumoto T, Takahashi A,
Takayama K, et al. 2012. Coxsackievirus B3 is an oncolytic virus with immunostimulatory properties that
is active against lung adenocarcinoma. Cancer Res 72(10):2609-21.
Moerdyk-Schauwecker M, Shah NR, Murphy AM, Hastie E, Mukherjee P, Grdzelishvili VZ. 2013. Resistance
of pancreatic cancer cells to oncolytic vesicular stomatitis virus: Role of type I interferon signaling.
Virology 436(1):221-34.
Mokhtarian F, Zhang Z, Shi Y, Gonzales E, Sobel RA. 1999. Molecular mimicry between a viral peptide and a
myelin oligodendrocyte glycoprotein peptide induces autoimmune demyelinating disease in mice. J
Neuroimmunol 95(1-2):43-54.
MOORE AE. 1954. Effects of viruses on tumors. Annu Rev Microbiol 8:393-410.
Moran TP, Burgents JE, Long B, Ferrer I, Jaffee EM, Tisch RM, Johnston RE, Serody JS. 2007. Alphaviral
vector-transduced dendritic cells are successful therapeutic vaccines against neu-overexpressing tumors in
wild-type mice. Vaccine 25(36):6604-12.
Msaouel P, Iankov ID, Allen C, Morris JC, von Messling V, Cattaneo R, Koutsilieris M, Russell SJ, Galanis E.
2009. Engineered measles virus as a novel oncolytic therapy against prostate cancer. Prostate 69(1):82-91.
Muller U, Steinhoff U, Reis LF, Hemmi S, Pavlovic J, Zinkernagel RM, Aguet M. 1994. Functional role of type I
and type II interferons in antiviral defense. Science 264(5167):1918-21.
Munch RC, Muhlebach MD, Schaser T, Kneissl S, Jost C, Pluckthun A, Cichutek K, Buchholz CJ. 2011.
DARPins: An efficient targeting domain for lentiviral vectors. Mol Ther 19(4):686-93.
Murakami T, Cardones AR, Finkelstein SE, Restifo NP, Klaunberg BA, Nestle FO, Castillo SS, Dennis PA,
Hwang ST. 2003. Immune evasion by murine melanoma mediated through CC chemokine receptor-10. J
Exp Med 198(9):1337-47.
Murchison EP, Tovar C, Hsu A, Bender HS, Kheradpour P, Rebbeck CA, Obendorf D, Conlan C, Bahlo M,
Blizzard CA, et al. 2010. The tasmanian devil transcriptome reveals schwann cell origins of a clonally
transmissible cancer. Science 327(5961):84-7.
Murphy AM, Sheahan BJ, Atkins GJ. 2001. Induction of apoptosis in BCL-2-expressing rat prostate cancer cells
using the semliki forest virus vector. Int J Cancer 94(4):572-8.
Murphy AM, Morris-Downes MM, Sheahan BJ, Atkins GJ. 2000. Inhibition of human lung carcinoma cell
growth by apoptosis induction using semliki forest virus recombinant particles. Gene Ther 7(17):1477-82.
Murray P, Rosenthal K, Pfaller M. 2013. Medical microbiology. 7th ed. Elsevier Saunders.
Muster T, Rajtarova J, Sachet M, Unger H, Fleischhacker R, Romirer I, Grassauer A, Url A, Garcia-Sastre A,
Wolff K, et al. 2004. Interferon resistance promotes oncolysis by influenza virus NS1-deletion mutants. Int J
Cancer 110(1):15-21.
Muthana M, Rodrigues S, Chen YY, Welford A, Hughes R, Tazzyman S, Essand M, Morrow F, Lewis CE.
2013. Macrophage delivery of an oncolytic virus abolishes tumor regrowth and metastasis after
chemotherapy or irradiation. Cancer Res 73(2):490-5.
97
Naik S, Nace R, Federspiel MJ, Barber GN, Peng KW, Russell SJ. 2012. Curative one-shot systemic virotherapy
in murine myeloma. Leukemia 26(8):1870-8.
Nandi S and Lesniak MS. 2009. Adenoviral virotherapy for malignant brain tumors. Expert Opin Biol Ther
9(6):737-47.
Nandi S, Ulasov IV, Tyler MA, Sugihara AQ, Molinero L, Han Y, Zhu ZB, Lesniak MS. 2008. Low-dose
radiation enhances survivin-mediated virotherapy against malignant glioma stem cells. Cancer Res
68(14):5778-84.
Nargi-Aizenman JL and Griffin DE. 2001. Sindbis virus-induced neuronal death is both necrotic and apoptotic
and is ameliorated by N-methyl-D-aspartate receptor antagonists. J Virol 75(15):7114-21.
Naslund TI, Uyttenhove C, Nordstrom EK, Colau D, Warnier G, Jondal M, Van den Eynde BJ, Liljestrom P.
2007. Comparative prime-boost vaccinations using semliki forest virus, adenovirus, and ALVAC vectors
demonstrate differences in the generation of a protective central memory CTL response against the P815
tumor. J Immunol 178(11):6761-9.
Navin N, Kendall J, Troge J, Andrews P, Rodgers L, McIndoo J, Cook K, Stepansky A, Levy D, Esposito D, et
al. 2011. Tumour evolution inferred by single-cell sequencing. Nature 472(7341):90-4.
Nayernia Z, Turchi L, Cosset E, Peterson H, Dutoit V, Dietrich PY, Tirefort D, Chneiweiss H, Lobrinus JA,
Krause KH, et al. 2013. The relationship between brain tumor cell invasion of engineered neural tissues
and in vivo features of glioblastoma. Biomaterials 34(33):8279-90.
Nguyen TL, Abdelbary H, Arguello M, Breitbach C, Leveille S, Diallo JS, Yasmeen A, Bismar TA, Kirn D, Falls
T, et al. 2008. Chemical targeting of the innate antiviral response by histone deacetylase inhibitors renders
refractory cancers sensitive to viral oncolysis. Proc Natl Acad Sci U S A 105(39):14981-6.
Nikonov A, Molder T, Sikut R, Kiiver K, Mannik A, Toots U, Lulla A, Lulla V, Utt A, Merits A, et al. 2013.
RIG-I and MDA-5 detection of viral RNA-dependent RNA polymerase activity restricts positive-strand
RNA virus replication. PLoS Pathog 9(9):e1003610.
Noser JA, Mael AA, Sakuma R, Ohmine S, Marcato P, Lee PW, Ikeda Y. 2007. The RAS/Raf1/MEK/ERK
signaling pathway facilitates VSV-mediated oncolysis: Implication for the defective interferon response in
cancer cells. Mol Ther 15(8):1531-6.
Oaten SW, Jagelman S, Webb HE. 1980. Further studies of macrophages in relationship to avirulent semliki
forest virus infections. Br J Exp Pathol 61(2):150-5.
O'Connor RJ and Hearing P. 2000. The E4-6/7 protein functionally compensates for the loss of E1A expression
in adenovirus infection. J Virol 74(13):5819-24.
Ogbomo H, Zemp FJ, Lun X, Zhang J, Stack D, Rahman MM, McFadden G, Mody CH, Forsyth PA. 2013.
Myxoma virus infection promotes NK lysis of malignant gliomas in vitro and in vivo. PLoS One
8(6):e66825.
Ohno K, Sawai K, Iijima Y, Levin B, Meruelo D. 1997. Cell-specific targeting of sindbis virus vectors
displaying IgG-binding domains of protein A. Nat Biotechnol 15(8):763-7.
Opyrchal M, Allen C, Iankov I, Aderca I, Schroeder M, Sarkaria J, Galanis E. 2012. Effective radiovirotherapy
for malignant gliomas by using oncolytic measles virus strains encoding the sodium iodide symporter
(MV-NIS). Hum Gene Ther 23(4):419-27.
O'Shea C, Klupsch K, Choi S, Bagus B, Soria C, Shen J, McCormick F, Stokoe D. 2005. Adenoviral proteins
mimic nutrient/growth signals to activate the mTOR pathway for viral replication. EMBO J 24(6):1211-21.
O'Shea CC. 2005. Viruses - seeking and destroying the tumor program. Oncogene 24(52):7640-55.
Ostman A and Heldin CH. 2007. PDGF receptors as targets in tumor treatment. Adv Cancer Res 97:247-74.
Ottolino-Perry K, Diallo JS, Lichty BD, Bell JC, McCart JA. 2010. Intelligent design: Combination therapy with
oncolytic viruses. Mol Ther 18(2):251-63.
Paessler S and Weaver SC. 2009. Vaccines for venezuelan equine encephalitis. Vaccine 27 Suppl 4:D80-5.
Paglino JC, Andres W, van den Pol AN. 2014. Autonomous parvoviruses neither stimulate nor are inhibited
by the type I interferon response in human normal or cancer cells. J Virol 88(9):4932-42.
Pan WY, Lo CH, Chen CC, Wu PY, Roffler SR, Shyue SK, Tao MH. 2012. Cancer immunotherapy using a
membrane-bound interleukin-12 with B7-1 transmembrane and cytoplasmic domains. Mol Ther 20(5):92737.
Pansky A, Hildebrand P, Fasler-Kan E, Baselgia L, Ketterer S, Beglinger C, Heim MH. 2000. Defective jakSTAT signal transduction pathway in melanoma cells resistant to growth inhibition by interferon-alpha.
Int J Cancer 85(5):720-5.
98
Parato KA, Breitbach CJ, Le Boeuf F, Wang J, Storbeck C, Ilkow C, Diallo JS, Falls T, Burns J, Garcia V, et al.
2012. The oncolytic poxvirus JX-594 selectively replicates in and destroys cancer cells driven by genetic
pathways commonly activated in cancers. Mol Ther 20(4):749-58.
Park EH, Park EH, Cho IR, Srisuttee R, Min HJ, Oh MJ, Jeong YJ, Jhun BH, Johnston RN, Lee S, et al. 2010.
CUG2, a novel oncogene confers reoviral replication through ras and p38 signaling pathway. Cancer Gene
Ther 17(5):307-14.
Peltoniemi J, Setala N, Broberg E, Roytta M, Hukkanen V, Salmi AA, Eralinna JP. 2002. Semliki forest virus
infection is enhanced in Th1-prone SJL mice but not in Th2-prone BALB/c mice during linomide-induced
immunomodulation. J Neuroimmunol 132(1-2):83-92.
Perez OD, Logg CR, Hiraoka K, Diago O, Burnett R, Inagaki A, Jolson D, Amundson K, Buckley T, Lohse D, et
al. 2012. Design and selection of toca 511 for clinical use: Modified retroviral replicating vector with
improved stability and gene expression. Mol Ther 20(9):1689-98.
Perou CM, Jeffrey SS, van de Rijn M, Rees CA, Eisen MB, Ross DT, Pergamenschikov A, Williams CF, Zhu SX,
Lee JC, et al. 1999. Distinctive gene expression patterns in human mammary epithelial cells and breast
cancers. Proc Natl Acad Sci U S A 96(16):9212-7.
Pesonen S, Diaconu I, Kangasniemi L, Ranki T, Kanerva A, Pesonen SK, Gerdemann U, Leen AM, Kairemo K,
Oksanen M, et al. 2012. Oncolytic immunotherapy of advanced solid tumors with a CD40L-expressing
replicating adenovirus: Assessment of safety and immunologic responses in patients. Cancer Res
72(7):1621-31.
Pickup M, Novitskiy S, Moses HL. 2013. The roles of TGFbeta in the tumour microenvironment. Nat Rev
Cancer 13(11):788-99.
Platanias LC. 2005. Mechanisms of type-I- and type-II-interferon-mediated signalling. Nat Rev Immunol
5(5):375-86.
Pol JG, Marguerie M, Arulanandam R, Bell JC, Lichty BD. 2013. Panorama from the oncolytic virotherapy
summit. Mol Ther 21(10):1814-8.
American Type Culture Collection (ATCC) [Internet]: ATCC; c2014 [cited 2014 . Available
from: http://www.lgcstandards-atcc.org/products/all/HTB-14.aspx?geo_country=fi .
Prestwich RJ, Ilett EJ, Errington F, Diaz RM, Steele LP, Kottke T, Thompson J, Galivo F, Harrington KJ,
Pandha HS, et al. 2009. Immune-mediated antitumor activity of reovirus is required for therapy and is
independent of direct viral oncolysis and replication. Clin Cancer Res 15(13):4374-81.
Protti MP and De Monte L. 2012. Cross-talk within the tumor microenvironment mediates Th2-type
inflammation in pancreatic cancer. Oncoimmunology 1(1):89-91.
Pukkala E, Sankila R, Rautalahti M. 2011. Syöpä suomessa 2011. 13th ed. Helsinki: Suomen Syöpäyhdistys.
Qiao J, Wang H, Kottke T, White C, Twigger K, Diaz RM, Thompson J, Selby P, de Bono J, Melcher A, et al.
2008. Cyclophosphamide facilitates antitumor efficacy against subcutaneous tumors following intravenous
delivery of reovirus. Clin Cancer Res 14(1):259-69.
Querido E, Blanchette P, Yan Q, Kamura T, Morrison M, Boivin D, Kaelin WG, Conaway RC, Conaway JW,
Branton PE. 2001. Degradation of p53 by adenovirus E4orf6 and E1B55K proteins occurs via a novel
mechanism involving a cullin-containing complex. Genes Dev 15(23):3104-17.
Quetglas JI, Ruiz-Guillen M, Aranda A, Casales E, Bezunartea J, Smerdou C. 2010. Alphavirus vectors for
cancer therapy. Virus Res 153(2):179-96.
Quetglas JI, Dubrot J, Bezunartea J, Sanmamed MF, Hervas-Stubbs S, Smerdou C, Melero I. 2012a.
Immunotherapeutic synergy between anti-CD137 mAb and intratumoral administration of a cytopathic
semliki forest virus encoding IL-12. Mol Ther 20(9):1664-75.
Quetglas JI, Fioravanti J, Ardaiz N, Medina-Echeverz J, Baraibar I, Prieto J, Smerdou C, Berraondo P. 2012b. A
semliki forest virus vector engineered to express IFNalpha induces efficient elimination of established
tumors. Gene Ther 19(3):271-8.
Rabinovich GA, Gabrilovich D, Sotomayor EM. 2007. Immunosuppressive strategies that are mediated by
tumor cells. Annu Rev Immunol 25:267-96.
Rani MR, Foster GR, Leung S, Leaman D, Stark GR, Ransohoff RM. 1996. Characterization of beta-R1, a gene
that is selectively induced by interferon beta (IFN-beta) compared with IFN-alpha. J Biol Chem
271(37):22878-84.
Raper SE, Chirmule N, Lee FS, Wivel NA, Bagg A, Gao GP, Wilson JM, Batshaw ML. 2003. Fatal systemic
inflammatory response syndrome in a ornithine transcarbamylase deficient patient following adenoviral
gene transfer. Mol Genet Metab 80(1-2):148-58.
99
Rautsi O, Lehmusvaara S, Salonen T, Hakkinen K, Sillanpaa M, Hakkarainen T, Heikkinen S, Vahakangas E,
Yla-Herttuala S, Hinkkanen A, et al. 2007. Type I interferon response against viral and non-viral gene
transfer in human tumor and primary cell lines. J Gene Med 9(2):122-35.
Amgen's Press Release [Internet]; c29th of July 2011 [cited 2014 . Available
from: http://www.amgen.com/media/media_pr_detail.jsp?releaseID=1590733 .
Ricklin D, Hajishengallis G, Yang K, Lambris JD. 2010. Complement: A key system for immune surveillance
and homeostasis. Nat Immunol 11(9):785-97.
Rodriguez-Madoz JR, Prieto J, Smerdou C. 2005. Semliki forest virus vectors engineered to express higher IL12 levels induce efficient elimination of murine colon adenocarcinomas. Mol Ther 12(1):153-63.
Rodriguez-Madoz JR, Zabala M, Alfaro M, Prieto J, Kramer MG, Smerdou C. 2014. Short-term intratumoral
interleukin-12 expressed from an alphaviral vector is sufficient to induce an efficient antitumoral response
against spontaneous hepatocellular carcinomas. Hum Gene Ther 25(2):132-43.
Rodriguez-Madoz JR, Liu KH, Quetglas JI, Ruiz-Guillen M, Otano I, Crettaz J, Butler SD, Bellezza CA, Dykes
NL, Tennant BC, et al. 2009. Semliki forest virus expressing interleukin-12 induces antiviral and
antitumoral responses in woodchucks with chronic viral hepatitis and hepatocellular carcinoma. J Virol
83(23):12266-78.
Rommelfanger DM, Wongthida P, Diaz RM, Kaluza KM, Thompson JM, Kottke TJ, Vile RG. 2012. Systemic
combination virotherapy for melanoma with tumor antigen-expressing vesicular stomatitis virus and
adoptive T-cell transfer. Cancer Res 72(18):4753-64.
Rommelfanger DM, Compte M, Diaz RM, Ilett E, Alvarez-Vallina L, Thompson JM, Kottke TJ, Melcher A, Vile
RG. 2013. The efficacy versus toxicity profile of combination virotherapy and TLR immunotherapy
highlights the danger of administering TLR agonists to oncolytic virus-treated mice. Mol Ther 21(2):348-57.
Rothe M, Modlich U, Schambach A. 2013. Biosafety challenges for use of lentiviral vectors in gene therapy.
Curr Gene Ther 13(6):453-68.
Ruf IK, Lackey KA, Warudkar S, Sample JT. 2005. Protection from interferon-induced apoptosis by epsteinbarr virus small RNAs is not mediated by inhibition of PKR. J Virol 79(23):14562-9.
Russell SJ, Peng KW, Bell JC. 2012. Oncolytic virotherapy. Nat Biotechnol 30(7):658-70.
Ryman KD and Klimstra WB. 2008. Host responses to alphavirus infection. Immunol Rev 225:27-45.
Ryman KD, Klimstra WB, Nguyen KB, Biron CA, Johnston RE. 2000. Alpha/beta interferon protects adult mice
from fatal sindbis virus infection and is an important determinant of cell and tissue tropism. J Virol
74(7):3366-78.
Sadler AJ and Williams BR. 2008. Interferon-inducible antiviral effectors. Nat Rev Immunol 8(7):559-68.
Salih HR, Holdenrieder S, Steinle A. 2008. Soluble NKG2D ligands: Prevalence, release, and functional impact.
Front Biosci 13:3448-56.
Sampath P, Li J, Hou W, Chen H, Bartlett DL, Thorne SH. 2013. Crosstalk between immune cell and oncolytic
vaccinia therapy enhances tumor trafficking and antitumor effects. Mol Ther 21(3):620-8.
Sane J, Guedes S, Kurkela S, Lyytikainen O, Vapalahti O. 2010. Epidemiological analysis of mosquito-borne
pogosta disease in finland, 2009. Euro Surveill 15(2):19462.
Sarkioja M, Pesonen S, Raki M, Hakkarainen T, Salo J, Ahonen MT, Kanerva A, Hemminki A. 2008. Changing
the adenovirus fiber for retaining gene delivery efficacy in the presence of neutralizing antibodies. Gene
Ther 15(12):921-9.
Sato Y, Goto Y, Narita N, Hoon DS. 2009. Cancer cells expressing toll-like receptors and the tumor
microenvironment. Cancer Microenviron 2 Suppl 1:205-14.
Sawai K and Meruelo D. 1998. Cell-specific transfection of choriocarcinoma cells by using sindbis virus hCG
expressing chimeric vector. Biochem Biophys Res Commun 248(2):315-23.
Sawicki DL and Sawicki SG. 1980. Short-lived minus-strand polymerase for semliki forest virus. J Virol
34(1):108-18.
Sawicki DL, Sawicki SG, Keranen S, Kaariainen L. 1981a. Specific sindbis virus-coded function for minusstrand RNA synthesis. J Virol 39(2):348-58.
Sawicki SG, Sawicki DL, Kaariainen L, Keranen S. 1981b. A sindbis virus mutant temperature-sensitive in the
regulation of minus-strand RNA synthesis. Virology 115(1):161-72.
Scallan MF and Fazakerley JK. 1999. Aurothiolates enhance the replication of semliki forest virus in the CNS
and the exocrine pancreas. J Neurovirol 5(4):392-400.
100
Schafer S, Weibel S, Donat U, Zhang Q, Aguilar RJ, Chen NG, Szalay AA. 2012. Vaccinia virus-mediated intratumoral expression of matrix metalloproteinase 9 enhances oncolysis of PC-3 xenograft tumors. BMC
Cancer 12:366,2407-12-366.
Schambach A, Zychlinski D, Ehrnstroem B, Baum C. 2013. Biosafety features of lentiviral vectors. Hum Gene
Ther 24(2):132-42.
Schiavoni G, Mattei F, Di Pucchio T, Santini SM, Bracci L, Belardelli F, Proietti E. 2000. Cyclophosphamide
induces type I interferon and augments the number of CD44(hi) T lymphocytes in mice: Implications for
strategies of chemoimmunotherapy of cancer. Blood 95(6):2024-30.
Schilte C, Couderc T, Chretien F, Sourisseau M, Gangneux N, Guivel-Benhassine F, Kraxner A, Tschopp J,
Higgs S, Michault A, et al. 2010. Type I IFN controls chikungunya virus via its action on nonhematopoietic
cells. J Exp Med 207(2):429-42.
Schlee M and Hartmann G. 2010. The chase for the RIG-I ligand--recent advances. Mol Ther 18(7):1254-62.
Schmeisser H, Bekisz J, Zoon KC. 2014. New function of type I IFN: Induction of autophagy. J Interferon
Cytokine Res 34(2):71-8.
Schoggins JW, Wilson SJ, Panis M, Murphy MY, Jones CT, Bieniasz P, Rice CM. 2011. A diverse range of gene
products are effectors of the type I interferon antiviral response. Nature 472(7344):481-5.
Schulz O, Diebold SS, Chen M, Naslund TI, Nolte MA, Alexopoulou L, Azuma YT, Flavell RA, Liljestrom P,
Reis e Sousa C. 2005. Toll-like receptor 3 promotes cross-priming to virus-infected cells. Nature
433(7028):887-92.
Schwartz O and Albert ML. 2010. Biology and pathogenesis of chikungunya virus. Nat Rev Microbiol
8(7):491-500.
Shankaran V, Ikeda H, Bruce AT, White JM, Swanson PE, Old LJ, Schreiber RD. 2001. IFNgamma and
lymphocytes prevent primary tumour development and shape tumour immunogenicity. Nature
410(6832):1107-11.
Shashkova EV, Doronin K, Senac JS, Barry MA. 2008. Macrophage depletion combined with anticoagulant
therapy increases therapeutic window of systemic treatment with oncolytic adenovirus. Cancer Res
68(14):5896-904.
BioWorld News Report [Internet]: BioWorld; c4th of September, 2013 [cited 2014 . Available
from: http://www.bioworld.com/content/transgene-falls-pexa-vec-phase-ii-miss-liver-cancer-0 .
Simmons JD, Wollish AC, Heise MT. 2010. A determinant of sindbis virus neurovirulence enables efficient
disruption of jak/STAT signaling. J Virol 84(21):11429-39.
Singh I and Helenius A. 1992. Role of ribosomes in semliki forest virus nucleocapsid uncoating. J Virol
66(12):7049-58.
Singh IR, Suomalainen M, Varadarajan S, Garoff H, Helenius A. 1997. Multiple mechanisms for the inhibition
of entry and uncoating of superinfecting semliki forest virus. Virology 231(1):59-71.
Singh VK, Damewood GP,4th, Friedman RM, Maheshwari RK. 1987. Tunicamycin enhances virus replication
and inhibits antiviral activity of interferon in mice: Correlation with natural killer cells. J Exp Pathol
3(1):19-33.
Siu RW, Fragkoudis R, Simmonds P, Donald CL, Chase-Topping ME, Barry G, Attarzadeh-Yazdi G,
Rodriguez-Andres J, Nash AA, Merits A, et al. 2011. Antiviral RNA interference responses induced by
semliki forest virus infection of mosquito cells: Characterization, origin, and frequency-dependent
functions of virus-derived small interfering RNAs. J Virol 85(6):2907-17.
Sivendran S, Pan M, Kaufman HL, Saenger Y. 2010. Herpes simplex virus oncolytic vaccine therapy in
melanoma. Expert Opin Biol Ther 10(7):1145-53.
Smit JM, Waarts BL, Kimata K, Klimstra WB, Bittman R, Wilschut J. 2002. Adaptation of alphaviruses to
heparan sulfate: Interaction of sindbis and semliki forest viruses with liposomes containing lipidconjugated heparin. J Virol 76(20):10128-37.
Smyth JW, Fleeton MN, Sheahan BJ, Atkins GJ. 2005. Treatment of rapidly growing K-BALB and CT26 mouse
tumours using semliki forest virus and its derived vector. Gene Ther 12(2):147-59.
Sobol PT, Boudreau JE, Stephenson K, Wan Y, Lichty BD, Mossman KL. 2011. Adaptive antiviral immunity is
a determinant of the therapeutic success of oncolytic virotherapy. Mol Ther 19(2):335-44.
Spurrell E, Gangeswaran R, Wang P, Cao F, Gao D, Feng B, Wold W, Tollefson A, Lemoine NR, Wang Y. 2014.
STAT1 interaction with E3-14.7K in monocytes affects the efficacy of oncolytic adenovirus. J Virol
88(4):2291-300.
101
Spuul P, Balistreri G, Hellstrom K, Golubtsov AV, Jokitalo E, Ahola T. 2011. Assembly of alphavirus
replication complexes from RNA and protein components in a novel trans-replication system in
mammalian cells. J Virol 85(10):4739-51.
Stojdl DF, Lichty B, Knowles S, Marius R, Atkins H, Sonenberg N, Bell JC. 2000. Exploiting tumor-specific
defects in the interferon pathway with a previously unknown oncolytic virus. Nat Med 6(7):821-5.
Stojdl DF, Lichty BD, tenOever BR, Paterson JM, Power AT, Knowles S, Marius R, Reynard J, Poliquin L,
Atkins H, et al. 2003. VSV strains with defects in their ability to shutdown innate immunity are potent
systemic anti-cancer agents. Cancer Cell 4(4):263-75.
Sugio K, Sakurai F, Katayama K, Tashiro K, Matsui H, Kawabata K, Kawase A, Iwaki M, Hayakawa T,
Fujiwara T, et al. 2011. Enhanced safety profiles of the telomerase-specific replication-competent
adenovirus by incorporation of normal cell-specific microRNA-targeted sequences. Clin Cancer Res
17(9):2807-18.
Tanaka N, Ishihara M, Lamphier MS, Nozawa H, Matsuyama T, Mak TW, Aizawa S, Tokino T, Oren M,
Taniguchi T. 1996. Cooperation of the tumour suppressors IRF-1 and p53 in response to DNA damage.
Nature 382(6594):816-8.
Tang H, Sampath P, Yan X, Thorne SH. 2013. Potential for enhanced therapeutic activity of biological cancer
therapies with doxycycline combination. Gene Ther 20(7):770-8.
Tesfay MZ, Ammayappan A, Federspiel MJ, Barber GN, Stojdl D, Peng KW, Russell SJ. 2014. Vesiculovirus
neutralization by natural IgM and complement. J Virol 88(11):6148-57.
Thomas C, Moraga I, Levin D, Krutzik PO, Podoplelova Y, Trejo A, Lee C, Yarden G, Vleck SE, Glenn JS, et al.
2011. Structural linkage between ligand discrimination and receptor activation by type I interferons. Cell
146(4):621-32.
Thompson MR, Kaminski JJ, Kurt-Jones EA, Fitzgerald KA. 2011. Pattern recognition receptors and the innate
immune response to viral infection. Viruses 3(6):920-40.
Thorne SH, Hermiston T, Kirn D. 2005. Oncolytic virotherapy: Approaches to tumor targeting and enhancing
antitumor effects. Semin Oncol 32(6):537-48.
Thorne SH, Tam BY, Kirn DH, Contag CH, Kuo CJ. 2006. Selective intratumoral amplification of an
antiangiogenic vector by an oncolytic virus produces enhanced antivascular and anti-tumor efficacy. Mol
Ther 13(5):938-46.
Tisserand J, Khetchoumian K, Thibault C, Dembele D, Chambon P, Losson R. 2011. Tripartite motif 24
(Trim24/Tif1alpha) tumor suppressor protein is a novel negative regulator of interferon (IFN)/signal
transducers and activators of transcription (STAT) signaling pathway acting through retinoic acid receptor
alpha (raralpha) inhibition. J Biol Chem 286(38):33369-79.
Trgovcich J, Aronson JF, Johnston RE. 1996. Fatal sindbis virus infection of neonatal mice in the absence of
encephalitis. Virology 224(1):73-83.
Trobaugh DW, Gardner CL, Sun C, Haddow AD, Wang E, Chapnik E, Mildner A, Weaver SC, Ryman KD,
Klimstra WB. 2014. RNA viruses can hijack vertebrate microRNAs to suppress innate immunity. Nature
506(7487):245-8.
Trujillo MA, Oneal MJ, McDonough SJ, Morris JC. 2013. Viral dose, radioiodide uptake, and delayed efflux in
adenovirus-mediated NIS radiovirotherapy correlates with treatment efficacy. Gene Ther 20(5):567-74.
Tsai YC, Pestka S, Wang LH, Runnels LW, Wan S, Lyu YL, Liu LF. 2011. Interferon-beta signaling contributes
to ras transformation. PLoS One 6(8):e24291.
Tseng D, Volkmer JP, Willingham SB, Contreras-Trujillo H, Fathman JW, Fernhoff NB, Seita J, Inlay MA,
Weiskopf K, Miyanishi M, et al. 2013. Anti-CD47 antibody-mediated phagocytosis of cancer by
macrophages primes an effective antitumor T-cell response. Proc Natl Acad Sci U S A 110(27):11103-8.
Tseng JC, Levin B, Hurtado A, Yee H, Perez de Castro I, Jimenez M, Shamamian P, Jin R, Novick RP, Pellicer
A, et al. 2004. Systemic tumor targeting and killing by sindbis viral vectors. Nat Biotechnol 22(1):70-7.
Tsiftsoglou AS and Sartorelli AC. 1979. Dimethyl sulfoxide-induced differentiation of friend erythroleukemia
cells in the absence of cytokinesis. Cancer Res 39(10):4058-63.
Tsugawa Y, Kato H, Fujita T, Shimotohno K, Hijikata M. 2014. Critical role of interferon-alpha constitutively
produced in human hepatocytes in response to RNA virus infection. PLoS One 9(2):e89869.
Tuittila M and Hinkkanen AE. 2003. Amino acid mutations in the replicase protein nsP3 of semliki forest virus
cumulatively affect neurovirulence. J Gen Virol 84(Pt 6):1525-33.
Tuittila MT, Santagati MG, Roytta M, Maatta JA, Hinkkanen AE. 2000. Replicase complex genes of semliki
forest virus confer lethal neurovirulence. J Virol 74(10):4579-89.
102
Tysome JR, Li X, Wang S, Wang P, Gao D, Du P, Chen D, Gangeswaran R, Chard LS, Yuan M, et al. 2012. A
novel therapeutic regimen to eradicate established solid tumors with an effective induction of tumorspecific immunity. Clin Cancer Res 18(24):6679-89.
Ubol S, Levine B, Lee SH, Greenspan NS, Griffin DE. 1995. Roles of immunoglobulin valency and the heavychain constant domain in antibody-mediated downregulation of sindbis virus replication in persistently
infected neurons. J Virol 69(3):1990-3.
Uchida H, Marzulli M, Nakano K, Goins WF, Chan J, Hong CS, Mazzacurati L, Yoo JY, Haseley A, Nakashima
H, et al. 2013. Effective treatment of an orthotopic xenograft model of human glioblastoma using an EGFRretargeted oncolytic herpes simplex virus. Mol Ther 21(3):561-9.
Uchime O, Fields W, Kielian M. 2013. The role of E3 in pH protection during alphavirus assembly and exit. J
Virol 87(18):10255-62.
Vähä-Koskela M. 2007. Construction and characterization of a replication-competent expression vector based
on avirulent semliki forest virus. Turku: Åbo Akademi.
Vaha-Koskela MJ, Heikkila JE, Hinkkanen AE. 2007. Oncolytic viruses in cancer therapy. Cancer Lett
254(2):178-216.
Vaha-Koskela MJ, Tuittila MT, Nygardas PT, Nyman JK, Ehrengruber MU, Renggli M, Hinkkanen AE. 2003.
A novel neurotropic expression vector based on the avirulent A7(74) strain of semliki forest virus. J
Neurovirol 9(1):1-15.
Vaha-Koskela MJ, Kallio JP, Jansson LC, Heikkila JE, Zakhartchenko VA, Kallajoki MA, Kahari VM,
Hinkkanen AE. 2006. Oncolytic capacity of attenuated replicative semliki forest virus in human melanoma
xenografts in severe combined immunodeficient mice. Cancer Res 66(14):7185-94.
Vaha-Koskela MJ, Le Boeuf F, Lemay C, De Silva N, Diallo JS, Cox J, Becker M, Choi Y, Ananth A, Sellers C, et
al. 2013. Resistance to two heterologous neurotropic oncolytic viruses, semliki forest virus and vaccinia
virus, in experimental glioma. J Virol 87(4):2363-6.
van Beusechem VW, Mastenbroek DC, van den Doel PB, Lamfers ML, Grill J, Wurdinger T, Haisma HJ,
Pinedo HM, Gerritsen WR. 2003. Conditionally replicative adenovirus expressing a targeting adapter
molecule exhibits enhanced oncolytic potency on CAR-deficient tumors. Gene Ther 10(23):1982-91.
van Boxel-Dezaire AH, Rani MR, Stark GR. 2006. Complex modulation of cell type-specific signaling in
response to type I interferons. Immunity 25(3):361-72.
van der Most RG, Currie AJ, Cleaver AL, Salmons J, Nowak AK, Mahendran S, Larma I, Prosser A, Robinson
BW, Smyth MJ, et al. 2009. Cyclophosphamide chemotherapy sensitizes tumor cells to TRAIL-dependent
CD8 T cell-mediated immune attack resulting in suppression of tumor growth. PLoS One 4(9):e6982.
van Pesch V, Lanaya H, Renauld JC, Michiels T. 2004. Characterization of the murine alpha interferon gene
family. J Virol 78(15):8219-28.
Varble A, Benitez AA, Schmid S, Sachs D, Shim JV, Rodriguez-Barrueco R, Panis M, Crumiller M, Silva JM,
Sachidanandam R, et al. 2013. An in vivo RNAi screening approach to identify host determinants of virus
replication. Cell Host Microbe 14(3):346-56.
Viaud S, Saccheri F, Mignot G, Yamazaki T, Daillere R, Hannani D, Enot DP, Pfirschke C, Engblom C, Pittet
MJ, et al. 2013. The intestinal microbiota modulates the anticancer immune effects of cyclophosphamide.
Science 342(6161):971-6.
Volpers C and Kochanek S. 2004. Adenoviral vectors for gene transfer and therapy. J Gene Med 6 Suppl
1:S164-71.
Vonderheide RH and Glennie MJ. 2013. Agonistic CD40 antibodies and cancer therapy. Clin Cancer Res
19(5):1035-43.
Wagner TC, Velichko S, Chesney SK, Biroc S, Harde D, Vogel D, Croze E. 2004. Interferon receptor expression
regulates the antiproliferative effects of interferons on cancer cells and solid tumors. Int J Cancer 111(1):3242.
Wahlberg JM and Garoff H. 1992. Membrane fusion process of semliki forest virus. I: Low pH-induced
rearrangement in spike protein quaternary structure precedes virus penetration into cells. J Cell Biol
116(2):339-48.
Wakimoto H, Fulci G, Tyminski E, Chiocca EA. 2004. Altered expression of antiviral cytokine mRNAs
associated with cyclophosphamide's enhancement of viral oncolysis. Gene Ther 11(2):214-23.
Wakimoto H, Ikeda K, Abe T, Ichikawa T, Hochberg FH, Ezekowitz RA, Pasternack MS, Chiocca EA. 2002.
The complement response against an oncolytic virus is species-specific in its activation pathways. Mol
Ther 5(3):275-82.
103
Wang GP, Berry CC, Malani N, Leboulch P, Fischer A, Hacein-Bey-Abina S, Cavazzana-Calvo M, Bushman
FD. 2010a. Dynamics of gene-modified progenitor cells analyzed by tracking retroviral integration sites in
a human SCID-X1 gene therapy trial. Blood 115(22):4356-66.
Wang M, Bronte V, Chen PW, Gritz L, Panicali D, Rosenberg SA, Restifo NP. 1995. Active immunotherapy of
cancer with a nonreplicating recombinant fowlpox virus encoding a model tumor-associated antigen. J
Immunol 154(9):4685-92.
Wang N, Dong Q, Li J, Jangra RK, Fan M, Brasier AR, Lemon SM, Pfeffer LM, Li K. 2010b. Viral induction of
the zinc finger antiviral protein is IRF3-dependent but NF-kappaB-independent. J Biol Chem 285(9):608090.
Wang W, Zhu NL, Chua J, Swenson S, Costa FK, Schmitmeier S, Sosnowski BA, Shichinohe T, Kasahara N,
Chen TC. 2005. Retargeting of adenoviral vector using basic fibroblast growth factor ligand for malignant
glioma gene therapy. J Neurosurg 103(6):1058-66.
Weber LW, Bowne WB, Wolchok JD, Srinivasan R, Qin J, Moroi Y, Clynes R, Song P, Lewis JJ, Houghton AN.
1998. Tumor immunity and autoimmunity induced by immunization with homologous DNA. J Clin Invest
102(6):1258-64.
Weinberg R. 2014. The biology of CANCER. 2nd ed. Taylor & Francis Group, LLC, an informa business, 711
Third Avenue, New Yourk, NY 10017, USA, and Park Square, Milton Park, Abingdon, OX14 4RN, UK:
Garland Science.
Weissleder R, Moore A, Mahmood U, Bhorade R, Benveniste H, Chiocca EA, Basilion JP. 2000. In vivo
magnetic resonance imaging of transgene expression. Nat Med 6(3):351-5.
Welsh RM, Bahl K, Marshall HD, Urban SL. 2012. Type 1 interferons and antiviral CD8 T-cell responses. PLoS
Pathog 8(1):e1002352.
Werneke SW, Schilte C, Rohatgi A, Monte KJ, Michault A, Arenzana-Seisdedos F, Vanlandingham DL, Higgs
S, Fontanet A, Albert ML, et al. 2011. ISG15 is critical in the control of chikungunya virus infection
independent of UbE1L mediated conjugation. PLoS Pathog 7(10):e1002322.
White LK, Sali T, Alvarado D, Gatti E, Pierre P, Streblow D, Defilippis VR. 2011. Chikungunya virus induces
IPS-1-dependent innate immune activation and protein kinase R-independent translational shutoff. J Virol
85(1):606-20.
Wilcox ME, Yang W, Senger D, Rewcastle NB, Morris DG, Brasher PM, Shi ZQ, Johnston RN, Nishikawa S,
Lee PW, et al. 2001. Reovirus as an oncolytic agent against experimental human malignant gliomas. J Natl
Cancer Inst 93(12):903-12.
Willems WR, Kaluza G, Boschek CB, Bauer H, Hager H, Schutz HJ, Feistner H. 1979. Semliki forest virus:
Cause of a fatal case of human encephalitis. Science 203(4385):1127-9.
Amgen Press Release [Internet]: Amgen; c4th of April 2014 [cited 2014 . Available
from: http://wwwext.amgen.com/media/media_pr_detail.jsp?year=2014&releaseID=1915897 .
Willmon C, Diaz RM, Wongthida P, Galivo F, Kottke T, Thompson J, Albelda S, Harrington K, Melcher A, Vile
R. 2011. Vesicular stomatitis virus-induced immune suppressor cells generate antagonism between
intratumoral oncolytic virus and cyclophosphamide. Mol Ther 19(1):140-9.
Wilson RF. 2010. The death of jesse gelsinger: New evidence of the influence of money and prestige in human
research. Am J Law Med 36(2-3):295-325.
Woller N, Knocke S, Mundt B, Gurlevik E, Struver N, Kloos A, Boozari B, Schache P, Manns MP, Malek NP, et
al. 2011. Virus-induced tumor inflammation facilitates effective DC cancer immunotherapy in a tregdependent manner in mice. J Clin Invest 121(7):2570-82.
Wollish AC, Ferris MT, Blevins LK, Loo YM, Gale M,Jr, Heise MT. 2013. An attenuating mutation in a
neurovirulent sindbis virus strain interacts with the IPS-1 signaling pathway in vivo. Virology 435(2):26980.
Wollmann G, Robek MD, van den Pol AN. 2007. Variable deficiencies in the interferon response enhance
susceptibility to vesicular stomatitis virus oncolytic actions in glioblastoma cells but not in normal human
glial cells. J Virol 81(3):1479-91.
Wollmann G, Davis JN, Bosenberg MW, van den Pol AN. 2013. Vesicular stomatitis virus variants selectively
infect and kill human melanomas but not normal melanocytes. J Virol 87(12):6644-59.
Wong LH, Hatzinisiriou I, Devenish RJ, Ralph SJ. 1998. IFN-gamma priming up-regulates IFN-stimulated
gene factor 3 (ISGF3) components, augmenting responsiveness of IFN-resistant melanoma cells to type I
IFNs. J Immunol 160(11):5475-84.
104
Wong LH, Krauer KG, Hatzinisiriou I, Estcourt MJ, Hersey P, Tam ND, Edmondson S, Devenish RJ, Ralph SJ.
1997. Interferon-resistant human melanoma cells are deficient in ISGF3 components, STAT1, STAT2, and
p48-ISGF3gamma. J Biol Chem 272(45):28779-85.
Wongthida P, Diaz RM, Galivo F, Kottke T, Thompson J, Melcher A, Vile R. 2011. VSV oncolytic virotherapy
in the B16 model depends upon intact MyD88 signaling. Mol Ther 19(1):150-8.
Wongthida P, Diaz RM, Galivo F, Kottke T, Thompson J, Pulido J, Pavelko K, Pease L, Melcher A, Vile R. 2010.
Type III IFN interleukin-28 mediates the antitumor efficacy of oncolytic virus VSV in immune-competent
mouse models of cancer. Cancer Res 70(11):4539-49.
Workenhe ST and Mossman KL. 2014. Oncolytic virotherapy and immunogenic cancer cell death: Sharpening
the sword for improved cancer treatment strategies. Mol Ther 22(2):251-6.
Workenhe ST, Pol JG, Lichty BD, Cummings DT, Mossman KL. 2013. Combining oncolytic HSV-1 with
immunogenic cell death-inducing drug mitoxantrone breaks cancer immune tolerance and improves
therapeutic efficacy. Cancer Immunol Res 1(5):309-19.
Wu L, Fossum E, Joo CH, Inn KS, Shin YC, Johannsen E, Hutt-Fletcher LM, Hass J, Jung JU. 2009. Epstein-barr
virus LF2: An antagonist to type I interferon. J Virol 83(2):1140-6.
Xia M, Gonzalez P, Li C, Meng G, Jiang A, Wang H, Gao Q, Debatin KM, Beltinger C, Wei J. 2014. Mitophagy
enhances oncolytic measles virus replication by mitigating DDX58/RIG-I-like receptor signaling. J Virol
88(9):5152-64.
Yang SW, Cody JJ, Rivera AA, Waehler R, Wang M, Kimball KJ, Alvarez RA, Siegal GP, Douglas JT,
Ponnazhagan S. 2011. Conditionally replicating adenovirus expressing TIMP2 for ovarian cancer therapy.
Clin Cancer Res 17(3):538-49.
Ye J, Livergood RS, Peng G. 2013. The role and regulation of human Th17 cells in tumor immunity. Am J
Pathol 182(1):10-20.
Yin J, Gardner CL, Burke CW, Ryman KD, Klimstra WB. 2009. Similarities and differences in antagonism of
neuron alpha/beta interferon responses by venezuelan equine encephalitis and sindbis alphaviruses. J
Virol 83(19):10036-47.
Yla-Herttuala S. 2012. Endgame: Glybera finally recommended for approval as the first gene therapy drug in
the european union. Mol Ther 20(10):1831-2.
Ylosmaki E, Martikainen M, Hinkkanen A, Saksela K. 2013. Attenuation of semliki forest virus neurovirulence
by microRNA-mediated detargeting. J Virol 87(1):335-44.
Ylosmaki E, Hakkarainen T, Hemminki A, Visakorpi T, Andino R, Saksela K. 2008. Generation of a
conditionally replicating adenovirus based on targeted destruction of E1A mRNA by a cell type-specific
MicroRNA. J Virol 82(22):11009-15.
Yu H, Kortylewski M, Pardoll D. 2007. Crosstalk between cancer and immune cells: Role of STAT3 in the
tumour microenvironment. Nat Rev Immunol 7(1):41-51.
Zamarin D, Holmgaard RB, Subudhi SK, Park JS, Mansour M, Palese P, Merghoub T, Wolchok JD, Allison JP.
2014. Localized oncolytic virotherapy overcomes systemic tumor resistance to immune checkpoint
blockade immunotherapy. Sci Transl Med 6(226):226ra32.
Zamarin D, Martinez-Sobrido L, Kelly K, Mansour M, Sheng G, Vigil A, Garcia-Sastre A, Palese P, Fong Y.
2009. Enhancement of oncolytic properties of recombinant newcastle disease virus through antagonism of
cellular innate immune responses. Mol Ther 17(4):697-706.
Zemp FJ, Corredor JC, Lun X, Muruve DA, Forsyth PA. 2010. Oncolytic viruses as experimental treatments for
malignant gliomas: Using a scourge to treat a devil. Cytokine Growth Factor Rev 21(2-3):103-17.
Zemp FJ, McKenzie BA, Lun X, Maxwell L, Reilly KM, McFadden G, Yong VW, Forsyth PA. 2013. Resistance
to oncolytic myxoma virus therapy in nf1(-/-)/trp53(-/-) syngeneic mouse glioma models is independent of
anti-viral type-I interferon. PLoS One 8(6):e65801.
Zhang J, Yang PL, Gray NS. 2009. Targeting cancer with small molecule kinase inhibitors. Nat Rev Cancer
9(1):28-39.
Zhang YQ, Tsai YC, Monie A, Wu TC, Hung CF. 2010. Enhancing the therapeutic effect against ovarian cancer
through a combination of viral oncolysis and antigen-specific immunotherapy. Mol Ther 18(4):692-9.
Zheng S, Ulasov IV, Han Y, Tyler MA, Zhu ZB, Lesniak MS. 2007. Fiber-knob modifications enhance
adenoviral tropism and gene transfer in malignant glioma. J Gene Med 9(3):151-60.
Ziegler RJ, Li C, Cherry M, Zhu Y, Hempel D, van Rooijen N, Ioannou YA, Desnick RJ, Goldberg MA, Yew
NS, et al. 2002. Correction of the nonlinear dose response improves the viability of adenoviral vectors for
gene therapy of fabry disease. Hum Gene Ther 13(8):935-45.
Janne Ruotsalainen
Interferon Response as a
Challenge and Possibility
for Developing Alphavirus
Based Oncolytic Virotherapy
In this study, the possibilities of
targeting malignant glioma with
oncolytic virotherapy were explored
in different syngeneic- and xenograft
mouse-models. The specific roles
of innate and adaptive immune
responses were examined using
two mouse colon carcinoma clones
with opposing patterns of cell
autonomous antiviral signaling and
immunogenicity.
Publications of the University of Eastern Finland
Dissertations in Health Sciences
isbn 978-952-61-1554-2