* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download HPV DNA PARTITIONING DURING MITOSIS AS FOLLOWED
Genomic library wikipedia , lookup
Gene therapy of the human retina wikipedia , lookup
Cancer epigenetics wikipedia , lookup
Primary transcript wikipedia , lookup
Cre-Lox recombination wikipedia , lookup
Point mutation wikipedia , lookup
Artificial gene synthesis wikipedia , lookup
Therapeutic gene modulation wikipedia , lookup
Extrachromosomal DNA wikipedia , lookup
Site-specific recombinase technology wikipedia , lookup
DNA vaccination wikipedia , lookup
Polycomb Group Proteins and Cancer wikipedia , lookup
Mir-92 microRNA precursor family wikipedia , lookup
History of genetic engineering wikipedia , lookup
Vectors in gene therapy wikipedia , lookup
No-SCAR (Scarless Cas9 Assisted Recombineering) Genome Editing wikipedia , lookup
HPV DNA PARTITIONING DURING MITOSIS AS FOLLOWED BY FLUORESCENCE MICROSCOPY by ROBERT J. CARTER LOUISE T. CHOW, COMMITTEE CHAIR WILLIAM J. BRITT THOMAS R. BROKER IGOR CHESNOKOV N. PATRICK HIGGINS KENT T. KEYSER TIM M. TOWNES A DISSERTATION Submitted to the graduate faculty of The University of Alabama at Birmingham, in partial fulfillment of the requirements for the degree of Doctor of Philosophy BIRMINGHAM, ALABAMA 2010 Copyright by Robert J. Carter 2010 HPV DNA PARTITIONING DURING MITOSIS AS FOLLOWED BY FLUORESCENCE MICROSCOPY Robert J. Carter Department of Biochemistry and Molecular Genetics ABSTRACT Human papillomaviruses (HPVs) are small, double-stranded deoxyribonucleic acid (DNA) tumor viruses capable of establishing persistent infections in the epithelia. After infecting actively-dividing basal cells, the papillomavirus (PV) genome is maintained as extrachromosomal nuclear plasmids. It is largely unknown how the viral genome is maintained in actively-dividing cells. Our lab demonstrated that several HPV genotypes appear to employ a distinct strategy to facilitate partitioning of HPV DNA into daughter cells during mitosis. Association of the HPV-11 origin of replication (ori)binding protein E2 with the mitotic apparatus via cellular adapter proteins is thought to mediate equal partitioning of HPV genomes into daughter cells. The focus of this dissertation project was to devise and test a technique allowing direct visualization of replicated HPV DNA during mitosis. First, I developed a tracking plasmid system for observing newly-replicated HPV ori-containing plasmids during mitosis. Certain DNA-binding proteins only bind to a unique DNA sequence. I incorporated DNA binding-site arrays into HPV ori-containing plasmids, expressed DNA-binding domain-GFP fusion proteins, and visualized HPV tracking plasmids using high-resolution, high-sensitivity microscopy. Usage of Gal4 DNA-binding site (Gal4-DBD)- and TetR GFP fusion proteins each failed. Alternatively, rtTA3-eGFP-nls facilitated efficient, doxycycline(dox)-inducible, direct, in situ fluorescent detection of replicated HPV ori-containing plasmid without side-effects. TetO ar- iii rays I successfully-utilized were smaller than those previously-reported by others. With my 48-copy, tandem TetO array, plasmid was only detectable after exposure of rtTA3eGFP-nls to dox. To our knowledge this is the first time that DNA has been induciblylabeled by fluorescence. Secondly, I report that we were able to follow the fate of HPV plasmids during mitosis using imagery. A short, pre-treamtent with dox prior to fixation allowed us to visualize replicated plasmids as small, discrete foci with relatively-uniform intensities. Control TetO array-only control plasmids mostly coalesced to form a small number of large, bright aggregates. We suggest that replicated HPV DNA can persist in dividing cells rather than being sequestered and degraded as was the case with control nonreplicated control plasmids. iv ACKNOWLEDGMENTS I thank my mentors, Drs. Louise T. Chow and Thomas R. Broker, for allowing me to conduct my research in their lab with their guidance and support. I believe firmly that my dissertation work could not have been completed in any other laboratory in the world. Their expertise, advice, patience, and criticism were instrumental in shaping who I have become as a researcher. I thank Tim Townes and Pat Higgins for their tremendous support and encouragement, without which I probably would have left science. Dr. Townes went in with our lab to acquire the DeltaVision Spectris on which I completed much of my dissertation work. My family and I appreciated Dr. Higgins’ scientific exactness, perfect integrity and constructive feedback given to me at pivotal moments during my dissertation work. He requested and shared contructs from David Sherratt at Oxford University that were vital my completing my thesis project. Peter Prevelige and Tim Townes provided substantial financial support that enabled our lab to remain entirely-intact, even during the leanest of economic times. I appreciate those in the Broker/Chow laboratory who grew to become close friends over the years. In particular, Wayne Wang and his wife have been close friends and helpful for many years. Also, Biing-Yuan Lin and Nicholas Genovese were there to encourage me to press on when I was ready to throw the towel in and walk away during my most difficult times here at UAB. I thank Brian Van Tine for mentoring me as his v rotation student and teaching me so much of what I know about microscopy. He was instrumental in my attending the advanced microscopy course at Cold Spring Harbor Laboratories in 2002. I also must mention Sanjib Banerjee, Lucas Yu, and Luan Dao and thank them for conversations and life lessons that have shaped me to become who I am today. The training I received from Jason Swedlow, Abby Dernburg, and John Murray and their assistants while at Cold Spring Harbor Laboratories has impacted every aspect of my research. Last but definitely not least, my family deserves special mention. They tolerated the chaotic, demanding lifestyle that comes with having a scientist as a husband and father. My wife Aimee and children Ethan and Elise have been sources of smiles each time I return home. vi TABLE OF CONTENTS Page ABSTRACT....................................................................................................................... iii ACKNOWLEDGMENTS ...................................................................................................v LIST OF FIGURES ........................................................................................................... ix LIST OF ABBREVIATIONS..............................................................................................x INTRODUCTION ...............................................................................................................1 HPV pathobiology and life cycle.............................................................................1 HPV genome structure and functions......................................................................4 Normal morphology of the epidermis and the impact of HPV infection ................9 Multiple roles for the upstream regulatory region (URR).....................................11 Control of gene expression by elements within Papillomavirus URR’s ........12 The URR contains the papillomavirus origin of replication (ori) ..................13 Viral genes and proteins ........................................................................................15 E1 ...................................................................................................................15 E2 ...................................................................................................................16 E1^E4 .............................................................................................................17 E5 ...................................................................................................................18 E6 ...................................................................................................................19 E7 ...................................................................................................................21 L1 and L2 .......................................................................................................22 The partitioning of papillomavirus DNA into daughter cells requires the URR and E2 proteins.............................................................................................................22 Cell division: the mitotic apparatus and chromosome segregation .......................24 HPV DNA replication and early region message splicing ....................................25 Experimental aims and questions addressed in this project ..................................27 EXPERIMENTAL WORK................................................................................................28 Tracking DNA inside mammalian cells using Gal4-DBD, TetR, and rtTA3 GFP fusion proteins................................................................................................28 Pre-requisites and experimental constraints to carefully consider during development and testing of HPV tracking plasmid systems.......................................32 Expression of the HPV E1, E2, and E4 proteins impacts cell cycle progression and can be cytotoxic........................................................................32 vii peBFP2-11-E1E2 plasmid supports efficient ori-specific plasmid replication ..........................................................................................................33 Tight association of a GFP fusion protein in a large array to an HPV ori plasmid can impede ori replication.........................................................37 Tracking plasmid detection sensitivity during mitosis erquires optimized expression of the eBFP2-11-E1E2 operon and rtTA3-eGFP-nls ...............40 Clear visualization of HPV DNA localization is obtained when rtTA3-eGFP-nls binds to arrays of tetO sites embedded in tracking plasmids ........................43 The cellular distribution pattern of the 48-copy TetO array-containing plasmid ..........................................................................................................43 Plasmid does not associate with transiently-expressed rtTA3-eGFP-nls to form foci until cells are treated with doxycycline ..................................47 The 11-URR/TetO array-containing tracking plasmid is efficiently-targeted to microtubules during mitosis ...............................................................51 DISCUSSION....................................................................................................................57 Possibilities for future research .............................................................................59 Additional HPV tracking experiments of interest .........................................59 An alternative to ChIP to pull down discrete chromatin sequence................61 Super-resolution fixed- and live-cell imaging on an OMX microscope .......66 MATERIALS AND METHODS.......................................................................................66 GENERAL LIST OF REFERENCES ...............................................................................79 viii LIST OF FIGURES Figure Page 1 Differentiation profile of normal epithelium ...................................................................3 2 HPV-11 genome organization and transcripts produced by the E1E2 operon ................7 3 HPV-11 URR features ...................................................................................................13 4 The replication efficiency of HPV-11 ori plasmids in the presence or absence of TetR and doxycycline ............................................................................................................35 5 The replication efficiency of HPV-11 URR-containing plasmids in the presence or absence of rtTA3-eGFP-nls and doxycycline ..................................................................39 6 The distribution pattern of the negative control, 48-copy TetO array plasmid in doxycycline (dox)-treated cells ............................................................................................45 7 The inability to detect replicated 11-URR/48-copy TetO array plasmid in 293 cells not treated with doxycycline (dox) .....................................................................................49 8 The distribution pattern of replicated 11-URR/48-copy TetO array tracking plasmid in doxycycline (dox)-treated 293 cells..............................................................................55 ix LIST OF ABBREVIATIONS ATP adenosine triphosphate ATPase adenosine triphosphatase BFP blue fluorescent protein Bp base pair BPV bovine papillomavirus BS binding site C carboxyl (terminus of a protein) Cdk cyclin-dependent kinase cDNA complementary deoxyribonucleic acid Cy-3 cyanine 3 DAPI 4’,6-diamidino-2-phenylindole DNA deoxyribonucleic acid E2BS E2-binding site EBNA1 Epstein Barr nucleic antigen-1 EBV Epstein Barr virus FK foreskin keratinocyte G1 gap 1 phase of the cell cycle G2 gap 2 phase of the cell cycle Gal4DBD Gal4 deoxyribonucleic acid binding domain GFP green fluorescent protein x LIST OF ABBREVIATIONS (Continued) H hinge domain of a protein such as HPV E2 HAT histone acetyltranferase HDAC histone deacetylase HPV human papillomavirus HR high-risk, referring to the neoplastic potential of a papillomavirus Hsp heat shock protein IRES internal ribosome entry site Kb kilo base (1000 nucleotides, a standard for representing the length of a segment of DNA) Kbp kilo base pair kDa kilo Dalton, a unit of mass of a macromolecule LR low-risk, a non-oncogenic papillomavirus M mitosis MAP microtubule-associated protein (not to be confused with mitogen-activated protein kinase, MAPK) MARS matrix attachment regions MCM minichromosome maintenance mRNA messenger ribonucleic acid MTOC microtubule-organizing center N amino (terminus of a protein) NCR non-coding region, also called the upstream regulatory region (URR) of a papillomavirus xi LIST OF ABBREVIATIONS (Continued) NLS nuclear localization sequence ORC origin recognition complex ORF open reading frame Ori origin of replication sequence PBS phosphate-buffered saline PCR polymerase chain reaction PFA paraformaldehyde PHK primary human keratinocyte PML promyelocytic leukemia pRb retinoblastoma protein, p105 Puro puromycin resistance gene PV papillomavirus PVTD papillomavirus transcription domain RNA ribonucleic acid RNase ribonuclease rt room-temperature rtTA3 reverse tetracycline-inducible transcriptional activator protein S synthesis phase of the cell cycle SSC sodium chloride, sodium citrate SV40 simian virus 40 TetR tetracycline-responsive repressor protein URR upstream regulatory region (also see NCR) xii 1 INTRODUCTION HPV pathobiology and life cycle Papillomaviruses are non-enveloped, icosahedral DNA tumor viruses with circular, double-stranded genomes of approximately 8 kb (Figure 2). Each PV is host speciesspecific with infectivity limited to either mucosal or cutaneous epithelium at specific body sites. As of 2010, more than 150 different HPV types have been recovered from infected human tissue and molecularly characterized (de Villiers et al. 2004). HPV infections are often sub-clinical. Active infections cause epithelial hyperproliferation, with viral lesions being referred to as warts, papillomas, condylomas, or low grade dysplasias, depending on the body site(s) being affected. Active infections usually regress but may reactivate at a later time. At a low frequency, infections by the “high-risk” oncogenic viruses such as HPV-16, HPV-18 and closely related types can progress to cause highgrade dysplastic lesions and carcinomas. Studies of men with penile cancer and women with cervical cancer frequently reveal that high-risk HPVs are a necessary underlying cause. Cervical cancer is the most prevalent cancer affecting women under the age of 35 (zur Hausen 2002). Globally, cervical cancer remains the second most common malignancy amongst women. It is projected that over 500,000 new cases of cervical cancer will occur each year, and worldwide, more than 275,000 women will likely die of cervical cancer annually (zur Hausen 2000; zur Hausen 2002; Ellenson and Wu 2004). Studies of individuals within this same risk group show that HIV infection and AIDS in- 2 creases the rate of HPV disease progression (Aoki and Tosato 2004). The immunosuppressed condition of individuals co-infected with HPV and HIV may wholly account for the elevated pathogenicity of HPV oncoproteins. Contrary to the disease caused by the so-called high-risk HPV strains, low-risk HPVs, such as HPV-6 and HPV-11, can cause benign condylomas of the anogenital tract as well as recurrent laryngeal papillomatosis. Although benign in nature, patients with laryngeal papillomas can develop lifethreatening airway blockages unless surgically removed. The papillomavirus life cycle can be sub-divided into two or possibly three phases. The first is the initial infection of basal cells and the establishment of the infection. The second phase is the maintenance of the viral genome in the long-living basal or parabasal cycling cells. The third phase involves high-level amplification of the viral genome; it occurs only in a subset of post-mitotic, differentiated cells comprising a squamous epithelium. The capsid proteins are expressed and the progeny virions assembled in the uppermost several layers of cells, while maturation of virus particles takes place in the squames or cornified envelopes of dessicated cells, from which the virions are released upon mechanical disruption (Figure 1). Monolayer cell cultures are insufficient for studying most aspects of HPV biology because of the absence of cellular differentiation. To overcome this obstacle, a method was devised to produce differentiated squamous epithelial tissue in vitro. By culturing primary human keratinocytes (PHKs) atop collagen beds at the air-media interface, the cells divide and differentiate as they are pushed upward and further away from the liquid medium interface. This system allows production of organotypic epithelial tissue that is almost indistinguishable from clinical specimens (Dollard et al. 1992). The differentiated strata are labeled alongside an image 3 of a cross-section from an organotypic raft culture (Figure 1, left-hand column). The viral gene expression profile and certain viral processes such as virion assembly and release are restricted to discrete layers of cells within the stratified squamous epithelium (Figure 1, right-hand column). With the organotypic raft culture system, the complete virus lifecycle from infection to the shedding of progeny virions can now be reproduced and studied in vitro (Chow and Broker 2006; Wang et al. 2009). Figure 1: Differentiation profile of normal epithelium. Viruses infect basal and parabasal cells where low-level maintenance of the viral genome proceeds. After expression of the E6 and E7 oncoproteins in the spinous layer has artificially induced and sustained reentry into S-phase, high-level amplification of the viral genome commences. Expression of capsid genes, virion assembly, and the sloughing off of infectious virus particles are restricted to the upper layers of the stratified squamous epithelium. 4 HPV genome structure and functions The HPV genome can be divided into three regions: the upstream regulatory region (URR), the early (E) genes, and the late (L) genes. All open reading frames (ORFs) are encoded by the same DNA strand of the viral genome. All mRNA transcription proceeds in the same direction. Utilization of several promoters, two polyA sites and alternative mRNA splicing together provide access to each of the various ORFs. The following is a brief summary of cis elements identified in the URR and proteins encoded and their known function(s) during the viral life cycle. Subsequent sections contain more detailed discussion of the URR, viral genes and proteins. URR: Comprising approximately 700 bp of the PV genome, it contains transcriptional regulatory elements, promoters and the viral origin of replication (ori). The ori is comprised of multiple tandem binding sites for the viral E1 and E2 proteins necessary for the initiation and elongation of viral DNA replication. E1: E1 is the replicative DNA helicase required for initiation and elongation during HPV genome replication. E1 is responsible for recruitment of the host DNA replication machinery to the origin, making the successful replication of the viral genome possible. The extremely-rare E1 transcript is derived from an unspliced transcript spanning the multi-cistronic E1E2 operon that comprises much of the E region. E2: A multi-function protein, E2 is the ori-recognition and binding protein required for recruiting E1 to the ori to nucleate the assembly of the pre-initiation replication complex. E2 is also a context-dependent positive as well as negative tran- scription factor and is now also thought to be responsible for segregation of ori- 5 containing plasmids. The E2 mRNA arises from alternative E1 intragenic splicing of transcripts spanning the E region. E2 message is more abundant than E1 message but is still present at very low levels. E4: The cytoplasmic E4 protein binds to the network of cytokeratin intermediate filaments. The spliced E1^E4 mRNA is the most abundant viral mRNA and originates from primary transcripts that span the much or the entire E region. Its role in viral DNA amplification and in other stages of the viral life cycle is controversial and remains to be resolved. E5: E5 is a small trans-membrane protein primarily found in the endoplasmic reticulum and Golgi membranes that enhances growth factor receptor signal transduction, but its role in the viral life cycle is not particularly understood. E6: The high-risk HPV E6 protein causes the degradation of p53 protein and a number of other host proteins. The HPV E6 proteins from both the high- and low-risk genotypes can inactivate the p53 transcription activities. As well, there are many other host protein binding partners. The exact roles of E6 in the viral life cycle are yet to be elucidated. E7: The high-risk HPV E7 protein causes the degradation of the pRB family of proteins. In contrast, the low-risk HPV E7 only destabilizes p130, a pRB-related protein which prevents the differentiated cells from reentering the cell cycle, but it does not cause turnover of pRB. Thus, this E7 activity promotes S-phase re-entry by the differentiated spinous cells, creating a milieu conducive to viral DNA am- 6 plification. This restricts high-copy viral genome amplification to the upper strata of the squamous epithelium. In addition, E5, E6, and E7 proteins all contribute to the down-regulation of the host immune surveillance and defensive responses. L1 and L2: Major and minor capsid proteins, respectively. 7 Figure 2: HPV-11 genome organization and transcripts produced by the E1E2 operon. The HPV genome is roughly 8 Kb in length and possesses 8 open reading frames. The major promoters are found within the URR or within the coding region. The main transcripts produced by the E1E2 operon portion of the genome include E1, E1^E2, and E1^E4, and these encode the E1, E2, and E1^E4 proteins, respectively. The eBFP2-11E1E2 operon construct I generated results in blue-tagged E1 and E1^E4 proteins but native E2 protein. The PV URR (approximately 500-700 bp in length) contains promoters as well as numerous elements that bind host transcription factors (Bernard 2002). It also contains the viral origin of replication (ori), which consists of several binding sites for viral E1 and E2 proteins (Chow and Broker 2006). Early region genes include: E1, E2, E4, E5, E6, and E7. The encoded proteins are thought to regulate viral transcription and DNA replication, but the actual roles of E4, E5 and E6 protein are not completely understood. The late region encodes the L1 and L2 capsid proteins. The L1 or L1 plus L2 proteins can self-assemble into empty particles (virus–like particles), but DNA packaging into virions 8 requires both L1 and L2. In benign lesions, HPVs typically persist in host cells as extrachromosomal nuclear plasmids. In cancers, the viral DNA is often integrated into host chromosomes in the E1 or E2 ORF, disrupting the expression of all downstream genes. Transcriptional initiation from alternate promoters coupled with alternative RNA splicing and utilization of one of two polyadenylation sites result in more peptides than predicted from genes. While most major PV messages and peptides have been at least partially characterized, little is known about the significance of minor transcripts and the peptides that they potentially encode (reviewed by (Chow and Broker 2006)). HPVs have evolved to take advantage of biological processes unique to differentiated squamous epithelial tissues. Submerged cultures of undifferentiated, immortalized or transformed epithelial cell lines do not support viral reproduction. Investigators have learned to isolate neonatal foreskin primary human keratinocytes (PHKs) and grow these cells into differentiated organotypic raft cultures that are almost completely indistinguishable from native skin (Chow and Broker 1997). Prior to inducing PHK differentiation into raft cultures, one or more HPV genes can be introduced via retrovirus-mediated gene transfer to investigate the viral protein functions. Alternatively, full-length HPV genomic DNA can be excised from recombinant plasmids and transfected into PHKs. Raft cultures of cells immortalized by transfected high risk HPV DNA support the production of a small number of virus particles. More recently, a new strategy that enables high-titer production of HPV-18 virus in PHKs was devised. PHKs are transfected with HPV-18 whole-genomic plasmids that also contain an antibiotic selection marker flanked by loxP sites. Following selection of the transfected monolayer PHKs for drug-resistance, Cre recombinase is expressed. The selection marker excises and results in a plasmid contain- 9 ing the intact HPV-18 genome and one loxP site. These cells, when grown as raft cultures, recapitulate the life cycle of HPV-18 and generate high-titer, infectious virus particles (Chow et al. 2009; Wang et al. 2009). This efficient, reproducible system makes it possible to perform mutagenic studies of individual viral genes in the context of the whole HPV in the squamous epithelium, independent of the immortalization functions. The results show for the first time that E6 protein is essential for viral DNA amplification (Chow et al. 2009; Wang et al. 2009). Using this system, the role E5 plays during the virus life cycle is now being elucidated in our lab. Understanding the intricacies of the HPV infection program will require this new method for culturing, generating and passaging HPV virus particles, with and without modifications to the viral genome. This will be crucial to attaining a better understanding of virus-host interactions. The regulation of virus gene expression and viral genome amplification by the URR is important during the HPV infectious cycle and is crucial for appropriate differentiation-dependent expression of the different HPV proteins throughout the stratified epithelium. Normal morphology of the epidermis and the impact of HPV infection In uninfected skin (Figure 1), the basal cells are usually relatively quiescent whereas the parabasal cells divide daily. The upwardly-displaced daughters of the parabasal cells exit the cell cycle and begin to differentiate, generating several layers of spinous cells. In the cutaneous skin, a layer or two of granular cells develop above the spinous cells and are further differentiated into layers of cornified envelopes. This differentiation program takes place over a period of 2-3 weeks in vivo but is achievable in 9 days to 2 weeks in raft cultures after lifting the assembly to the media-air interface. In 10 the absence of HPV, the spinous and granular cells are post-mitotic, having exited the cell cycle never to re-enter it. In HPV-infected tissues, when basal cells enter S-phase, low-level replication of viral DNA proceeds to keep pace with replication of the host genome. Thus, low level production of the E1 and E2 messages and protein products is necessary during this phase of the virus lifecycle. Mutational analyses suggest that the E6 and E7 proteins, but not the E4 and E5 proteins, are necessary for stable maintenance of the viral genome in transfected keratinocytes. Upon differentiation, HPV E7 begins to be expressed in the lowerto mid-spinous layers. Following retrovirus-mediated gene transfer and ectopic expression, the E7 oncoprotein overrides the cellular homeostasis conferred by p130 and promotes unscheduled S-phase re-entry in post-mitotic, differentiated primary keratinocytes in organotypic cultures. Thus, the E7 protein conditions the differentiated cells so that the HPV genome can amplify to a high copy-number (Cheng et al. 1995; Banerjee et al. 2006; Genovese et al. 2008). The function of E6 is not known, as E6 is neither necessary nor sufficient to promote S phase reentry by the differentiated keratinocytes in raft cultures. However, constitutive expression of E6 and E7 in the PHK raft cultures result in a tissue morphology resembling high-grade dysplasia (Halbert et al. 1992). Thus, E6 and E7 are likely to contribute to the increased thickness of the parabasal strata in productively-infected lesions. An E6-mutant genome cannot immortalize PHKs and thus could not be studied in organotypic raft cultures until recently. Our lab has now established the ability to recapitulate the HPV life cycle in a manner that can ultimately yield high-titer production of infectious virions in organotypic cultures of primary human keratinocytes (Chow et al. 11 2009; Wang et al. 2009). This capability has bolstered the toolbox for studying HPV immensely. Our research group has begun to examine the roles of E6 and E5 in the context of an intact HPV genome in the organotypic cultures. Multiple roles for the papillomavirus upstream regulatory region (URR) Within each known type of papillomavirus (PV), approximately 500-700 bp of contiguous genomic DNA is commonly referred to as the upstream regulatory region (URR), long control region (LCR) or non-coding region (NCR). The URR immediately precedes the E6 gene and has no open reading frame. The E6-proximal portion of the URR contains one of the major transcription promoters overlapping the PV origin of replication (ori). The transcription profiles in PV-infected squamous epithelial cells change dramatically as these cells undergo differentiation. High-level viral gene expression and viral genome amplification occur mainly in the middle to upper strata of the squamous epithelium, whereas virion packaging occurs only in the superficial cells. These observations are significant because almost every aspect of the viral life cycle is partially or totally shielded from immune surveillance. Recognition sites for many transcription and replication factors have been identified in the URR. Understanding how individual elements within the URR contribute to the PV life cycle remains a formidable task to this day (Chow and Broker 1997; Parker et al. 1997; Ozbun and Meyers 1998; Zhao et al. 1999a; Zhao et al. 1999b; Sen et al. 2002; Bromberg-White and Meyers 2003; McLaughlin-Drubin et al. 2004; McLaughlin-Drubin and Meyers 2004; Sen et al. 2004). I will describe briefly (1) the elements that control 12 viral gene expression and (2) the elements that constitute the papillomavirus origin of replication (ori), which also functions to ensure that viral genome copies are partitioned equally between daughter cells in the dividing basal and parabasal cells. Control of gene expression by elements within papillomavirus URR’s Numerous regulatory elements in the URR are vital to the differentiationdependent life cycle of PVs. Just upstream of the E6 gene is the E6 promoter, generally referred to as P1 or more specifically as P94, P95, or P97 in HPV types -11, -18, or -16, respectively, according to the map location of the first transcribed nucleotide in the RNA transcript. The enhancer region includes binding sites for host transcription factors such as YY-1, NF-1, AP1, Oct-1, KRF, AP2, GRE, and Sp1 (Figure 3) (Bernard 2002). Varied functionality of these protein binding sites has been observed and is attributable to the sequence context and the HPV type. Placement of the reporter gene lacZ under the differentiation-dependent control of the E6 promoter limits β-galactosidase expression mainly to the spinous cells of the differentiated strata. Mutation, deletion, or replacement of known URR promoter or enhancer elements often has a negative impact on promoter activity (Parker et al. 1997; Zhao et al. 1997; Zhao et al. 1999a; Zhao et al. 1999b; Sen et al. 2002; Sen et al. 2004). The URR sequence immediately upstream of the E6 promoter includes several E2 binding sites (E2BSs) (Figure 3). The E2 protein was originally thought to modulate the expression of the E6 and E7 oncoproteins, as E2 binding to the E6 promoter-proximal binding site can prevent assembly of the pre-initiation complex onto the E6 promoter which is situated a few bp downstream (Hou et al. 2000) and can prevent the binding of 13 SP1, which is also critical for HPV E6 promoter activity (Dong et al. 1994). However, promoter repression by E2 was only observed in cell lines harboring integrated but not extrachromosomal HPV plasmid DNA (Bechtold et al. 2003). In vivo, there is some doubt regarding the ability of E2 to regulate viral promoters in cells containing extrachromosomal plasmid DNA (Bechtold et al. 2003). Rather, the primary function of E2 is to initiate viral DNA replication. HPV-11 URR Features 93 Figure 3: HPV-11 URR features. List of host protein binding sites (top list) and viral protein binding sites (bottom list). (adapted from (Zhao et al. 1999a), (Hou et al. 2000). The URR Contains the Papillomavirus Origin of DNA Replication (Ori) Several E2BSs flanking a series of overlapping E1 binding sites (E1BSs) comprise the minimal PV origin of replication (ori). Mutations in the E2BSs are deleterious to viral DNA replication (Chiang et al. 1992c; Kuo et al. 1994; Stubenrauch et al. 1998a; Stubenrauch et al. 1998b). The HPV E1 DNA helicase and E2 proteins interact directly, and the role of the E2 protein is to recruit the E1 protein to the ori. Both will be discussed in detail in later sections. The host cell supplies all additional replication en- 14 zymes, protein factors, and deoxynucleoside triphosphate substrates. Thus, PV DNA replication proceeds only when the differentiated cells re-enter the cell cycle. Recent work in our lab has demonstrated that viral DNA amplification occurs in G2 phase following the completion of the host DNA replication in S phase (Wang et al. 2009). Our lab believes that E2-E2BS interactions are also instrumental for faithful partitioning of the viral genome during mitosis in the basal/parabasal cell layers, thereby facilitating persistent HPV infections. 15 Viral Genes and Proteins E1: Replicative DNA helicase The PV E1 phosphoprotein (68kDa MW) is an ATPase with DNA helicase activity. It binds to the ori, albeit with low specificity and low affinity. It is recruited to the ori via an interaction with the E2 protein which binds to E2BSs with high affinity and specificity. Indeed, transient replication of PV ori-containing plasmids in transfected cells depends upon the presence of E2 (Ustav and Stenlund 1991; Chiang et al. 1992a). The E1 protein is necessary for the initiation of viral genome replication (Sun et al. 1990; Blitz and Laimins 1991; Ustav and Stenlund 1991; Yang et al. 1991; Chiang et al. 1992a; Seo et al. 1993; Yang et al. 1993; Kuo et al. 1994; MacPherson et al. 1994; Liu et al. 1995; Raj and Stanley 1995; Sanders and Stenlund 1998). E1 oligomerizes into hexameric or dihexameric rings, and intermediates of assembly are thought to be dimers or trimers (Liu et al. 1998; Sedman and Stenlund 1998; Conger et al. 1999; Fouts et al. 1999; Titolo et al. 2000). The ori-binding activity of E1 is enhanced through interactions with chaperone proteins such as Hsp40 and Hsp70 and, in the presence of Hsp40, ori-bound E1 protein is a dihexamer (Liu et al. 1998; Fouts et al. 1999; Titolo et al. 2000) which functions as a potent bidirectional helicase in the presence of RPA (replication protein A) and topoisomerase I (Lin et al. 2002). The ori-bound E1 protein recruits DNA polymerase-α/primase and RPA to the viral ori to initiate replication (Park et al. 1994; Masterson et al. 1998; Conger et al. 1999; Han et al. 1999; Amin et al. 2000). Although assembly of the preinitiation complex onto the viral ori is inherently linked to the presence of E2 protein, the 16 formation of the E1 hexamer and PV DNA replication cannot actually commence until E2 becomes dissociated from the PV ori sequence (Lusky et al. 1993; Yang et al. 1993; MacPherson et al. 1994; Lin et al. 2002; Abbate et al. 2004). E1 is transcribed from a differentiation-dependent promoter (E1 or P3 promoter) located immediately upstream of the E1 ORF overlapping the E7 gene. The great majority of this transcript is alternatively-spliced to encode E2 or E4 protein. Only rare unspliced mRNAs encode the E1 protein. In the high-risk HPV, E1 and E2 proteins can also be translated from transcripts initiated from the E6 (P1) promoter. E2: transcription factor; viral DNA replication factor; viral persistence factor The 43kDa E2 and amino-terminus-truncated E2-related proteins were first identified as factors that bound with specificity to conserved 12-nt elements within the URR to regulate expression of the E6 promoter (Spalholz et al. 1985; Androphy et al. 1987; Hirochika et al. 1987; Chin et al. 1988; Giri and Yaniv 1988; Gius et al. 1988; Hirochika et al. 1988; Chin et al. 1989; Ham et al. 1991; Chiang et al. 1992a; Dong et al. 1994). It was subsequently shown that E2 is essential for initiation of viral DNA replication (Ustav and Stenlund 1991; Chiang et al. 1992a; Chiang et al. 1992c; Liu et al. 1995). Recently, work from our lab showed that E2 tethers the HPV ori-DNA to the mitotic apparatus to facilitate partitioning of the viral genome into daughter cells and thereby to ensure viral persistence, and it has now been shown that E2 interacts with several proteins of the mitotic apparatus(Van Tine et al. 2004; Parish et al. 2006). E2 can be functionally sub-divided into three distinct regions: the amino-terminus (N), the hinge (H), and the carboxyl-terminus (C). HPV E2N is a transcriptional trans- 17 acting domain and associates with mitotic spindles. E2H associates with nuclear matrix proteins. E2C is responsible for DNA sequence-specific binding and protein dimerization and interacts very strongly with the mitotic apparatus; specifically, the E2C spindlelocalization domain has been mapped to a short stretch of peptides near the H domain (Dao et al. 2006). Alternate promoter utilization and mRNA splicing result in multiple E2-related proteins, most prominently the E8^E2C protein (also simply called E2C) expressed from a promoter within the E1 ORF and containing a different N terminal domain of about 12 amino acids encoded by the E8 ORF. Spliced transcripts that originate from the E1 promoter and remove much of the E1 ORF can encode the full-length E2 protein. Using alternative mRNA splicing or an alternative promoter, other minor E2-related proteins such as E1M^E2C of unknown function are expressed. The utilization of P1 promoter coupled with RNA splicing can also produce a transcript with potential to encode E2, as reported for some of the HR HPV types. E1^E4: Cytoskeletal binding protein. The highly-abundant E1^E4 peptide of about 85 residues is encoded by a spliced mRNA derived from a primary transcript which serves as the mRNA for the E1 protein. Splicing links the initiation codon and approximately four following triplets to a site near the beginning of the E4 ORF (Nasseri et al. 1987). Various properties have been attributed to the E1^E4 peptide, although the role of this small peptide during the PV life cycle is not well defined. The E4 protein binds to the network of cytokeratin intermediate filaments and causes the collapse of this cytoskeletal framework in transfected cell lines, 18 a phenomenon not seen in patient specimens or in productive raft cultures (Doorbar et al. 1991; Pray and Laimins 1995; Wang et al. 2004; Doorbar 2005). Many PV E4 proteins contain a domain capable of blocking cell cycle progression during G2-phase when overexpressed alone in cycling cell lines by sequestration of cyclin B/cdk1 (Davy et al. 2002). Whether this property of E4 is relevant to viral DNA amplification in differentiated epithelium is still a matter of debate. Other reports describe E1^E4 interactions with the mitochondria (Raj et al. 2004). The E1^E4 protein undergoes proteolytic cleavage and release of the amino terminal domain, followed by a conformational change that enables some of the interactions with host proteins. The unspliced E1 mRNA and the alternatively-spliced E2 mRNA are each of lowabundance relative to the E1^E4 mRNA. Thus, alternate splicing to create the E1^E4 message can regulate the expression of the HPV E1 and E2 replication proteins and replication efficiency (Deng et al. 2003), and such a role in balancing production of the replication proteins may indeed be achieved by the E1^E4 mRNA splicing. E5: Enhancement of growth factor receptor signal transduction (Crusius et al. 1999; Leykauf et al. 2004; Longworth and Laimins 2004) HPV E5 is a small peptide which primarily localizes to the endoplasmic reticulum (ER) membranes (Conrad et al. 1993; Sparkowski et al. 1995; Disbrow et al. 2003; Krawczyk et al. 2008). HPV E5 does not interact with or activate EGFR, and its activation of receptor tyrosine kinases (identity unknown) is ligand-dependent. The MAP kinase signaling cascade can lead to the transformation of NIH 3T3 cells. E5 also enhances the abilities of the E6 and E7 proteins to immortalize cells (Crusius et al. 2000; Suprynowicz et al. 2005). E5 is believed to play a role during the establishment of infec- 19 tions by inducing basal and suprabasal cells to cycle more readily, thus expanding the infected cell population. Its role in viral DNA amplification in differentiated cells is being investigated in our lab using the newly-established PHK raft culture system. . E6: Inactivation or degradation of host proteins, including p53, many PDZ containing proteins, RasGap protein, and TIP60, and up-regulation of hTERT expression. E6 is one of two human papillomavirus oncoproteins. The 151-residue protein folds to form two zinc finger domains. This oncoprotein is sufficient to immortalize mammary epithelial cells (Liu et al. 1995) but not dermal or cervical keratinocytes (Münger and Howley 2002). E6 protein can inactivate the p53 tumor suppressor protein, which normally functions to halt cell-cycle progression when DNA damage is present. In cells where p53 detects DNA damage but the cell does not repair it, p53 triggers caspasemediated apoptosis (Kastan et al. 1995). The high-risk HPV E6 protein can inactivate p53 in different ways. Firstly, E6 can target p53 for degradation via a complex between E6 and E6-AP which functions as a p53-specific ubiquitin ligase (Scheffner et al. 1993). Secondly, E6 efficiently inhibits the binding of p53 to its target sequences (Lechner and Laimins 1994; Thomas et al. 1995; Thomas et al. 1999). Thirdly, E6 binds to chromatinbound complexes of p53-p300 to prevent acetylation of p53 target genes, and this further silences transcription of genes that would otherwise be activated during a p53-mediated DNA damage response (Thomas and Chiang 2005). To the contrary, E6 proteins from low-risk PV types do not destabilize p53, but it retains the ability to inactivate p53 transcription activation function (Thomas and Chiang 2005). Fourth, the C-terminus in highrisk E6 oncoproteins contains a PDZ ligand that interacts with PDZ-domain containing 20 proteins and causes them to be polyubiquitinated (Nakagawa and Huibregtse 2000; Pim et al. 2000; Grm and Banks 2004; Hengstermann et al. 2005; Matsumoto et al. 2006; Handa et al. 2007; Shai et al. 2007). Moreover, the N-terminus of E6 destabilizes Tatinteracting protein 60 kDa (TIP60) (Jha et al. 2010). A catalytic sub-unit of the nucleosomal acetyltransferase of H4 (NuA4) complex, TIP60 associates with the HPV major early promoter, acetylates histone H4, and results in the recruitment of Brd4 and repression of E6 expression. By destabilizing TIP60, the HPV E6 promoter and other cellular promoters are de-repressed. During a DNA damage response, TIP60-induced histone modifications can promote p53-mediated apoptosis. By destabilizing TIP60 both low- and high-risk E6 proteins enhance cellular proliferation and survival (Jha et al. 2010). E6 proteins of the high-risk HPV types can up-regulate the transcription of hTERT, a component of the telomerase complex, thus providing a critical contribution to immortalization of cells (Kiyono et al. 1998; Oh et al. 2001; Lee et al. 2002; Gewin et al. 2004; Horner et al. 2004; Jeong Seo et al. 2004; Plug-DeMaggio et al. 2004). The E6 coding region in low-risk HPV types is unspliced, but all high-risk types contain alternative intragenic and intergenic splices, one of the key distinguishing hallmarks of these HPVs. The splice leads to a peptide that contains the amino terminus of E6, but terminates shortly after the splice acceptor to create one of several possible forms of E6* peptide, depending on the HR HPV genotype. The E6* peptide can counteract some of the functions of the full length E6 protein. The splice itself balances the production of E6 vs. E6*, and it also affects the level of translation from the downstream E7 ORF by creating an untranslated spacer between the end of E6* and E7 (Schneider- 21 Gadicke and Schwarz 1986; Bedell et al. 1989; Hummel et al. 1992; Woodworth et al. 1992). E7: Inactivation of pRB family of proteins (facilitates unscheduled S-phase re-entry) E7 is a major viral oncoprotein. E7 proteins expressed by high-risk PVs bind to and inactivate the retinoblastoma protein (pRB) which controls the G1-S transition in submerged, proliferating cells. Thus, over-expression of the HR HPV E6 and E7 genes can immortalize PHKs in vitro (Woodworth et al. 1992; Steenbergen et al. 1996) and initiate viral carcinogenesis in vitro. Cervical carcinoma cell lines such as HeLa, Caski, and SiHa rely upon the continuous expression of this viral oncoprotein as well as E6 for proliferation (Schneider-Gadicke and Schwarz 1986; Tsunokawa et al. 1986) as ectopic expression of E2 to suppress the E6 promoter in these cell lines represses E6/E7 transcription and cause the cells to senesce (Dowhanick et al. 1995; Francis et al. 2000; Nishimura et al. 2000; Wells et al. 2000; DeFilippis et al. 2003; Wells et al. 2003). In contrast, E7 proteins from low-risk PVs have a much lower affinity for pRB. However, E7 of both HR and LR HPVs can induce unscheduled S-phase reentry in terminally-differentiated spinous cells (Gage et al. 1990; Cheng et al. 1995; Banerjee et al. 2006) because the target of E7 in the differentiated keratinocytes is the pRB-related p130 protein which normally maintains the homeostasis of the differentiated cells (Sdek et al. 2006; Genovese et al. 2008). 22 L1 and L2: Major and minor capsid proteins, respectively L1 and L2 are the major and minor papillomavirus capsid proteins, respectively. Expression of these proteins can be observed in cells of the upper spinous layers where encapsidation of the viral genome takes place. The appropriate differentiation-dependent expression of L1 and L2 in the context of the full HPV genome can result in extremelystable, highly-infectious virions. Papillomavirus capsids are non-enveloped, icosahedral, and seen as 43-56 nm in diameter, depending on the staining method used for electron microscopy. Mature virus particles are released as corneocytes are sloughed off. During the next round of infection, the low-pH, reducing atmosphere in the endosome causes the virus coat to break down, thereby permitting release of the viral genome into the invaded cell. When expressed alone in eukaryotic cells, L1 self-assembles into virus-like particles (VLPs) suitable for use in vaccines. Immunization with non-infectious VLPs stimulates a high-titer antibody response that is protective against immunological challenge by native, infectious virions (Kirnbauer et al. 1992; Howley 1996; Schiller and Davies 2004; Chow and Broker 2006). The Partitioning of papillomavirus DNA into daughter cells requires the URR and E2 protein HPV DNA normally replicates and persists extrachromosomally as plasmid DNA without ever becoming integrated into host chromosomes during the virus life cycle. On very rare occasions, HPV DNA will erroneously integrate into a host cell chromosome, 23 primarily found in HR-HPV caused cancers. It has been proposed that integration of viral DNA disrupts the autoregulation of the E6 promoter and leads to upregulated expression of E6 and E7 oncoproteins, as a result of capturing a stable host polyA sequence located downstream of the integration site. Many aspects of the mechanisms of replication of the PV genome are able to be studied accurately in monolayer cells or in cell-free systems because PV DNA replication can occur when E1 and E2 are supplied in trans from expression vectors (Chiang et al. 1991; Ustav and Stenlund 1991; Yang et al. 1991; Chiang et al. 1992b; Kuo et al. 1994). Monolayer cells are also suitable while studying segregation of HPV DNA into daughter cells during mitosis. Recent findings from our lab identified novel interactions between the HPV-11 E2 protein and the mitotic spindles, and these interactions have been proposed to ensure faithful partitioning of HPV ori DNA into daughter cells during cell division (Van Tine et al. 2004; Dao et al. 2006). The ability of HPV E2 proteins to tether viral ori-containing DNA to the mitotic spindle apparatus could have serious biological ramifications in a cell containing integrated HPV DNA. The HPV ori could function as an additional centromere. The ability of E2 to bind to an integrated viral ori while simultaneously interacting with the mitotic spindles could induce a DNA breakage-fusion-bridge cycle causing genomic instability, and this genomic instability would not end until E2 expression is disrupted (Dao et al. 2006). This could explain why E2 is no longer expressed in all transformed cells lines that harbor integrated HPV DNA. The E2 proteins from multiple PV types have been found to interact with specific host proteins, and these interactions are thought to tether PV DNA to mitotic chromosomes to enable viral DNA partitioning (You et al. 2004; Baxter et al. 2005; 24 Brannon et al. 2005; McPhillips et al. 2005; Zheng et al. 2005; McPhillips et al. 2006; Oliveira et al. 2006; Parish et al. 2006; Yu et al. 2007; Cardenas-Mora et al. 2008). Cell division: the mitotic apparatus and chromosome segregation Mitotic cell division in mammalian cells requires a vast array of intricate, interdependent events and checkpoints to ensure that chromosomes segregate equally to yield sister cells with identical diploid genomes (Decordier et al. 2008). Chromosomal migrations are carefully controlled by bipolar spindles of microtubules nucleated by a pair of centrosomes. Daughter strands of a replicated chromosome known as sister chromatids pair up and become physically joined together early during mitosis at a chromosomal locus known as the centromere. During mitosis, a complex scaffold of proteins assembles onto each centromere to form a structure known as a kinetochore. Kinetochores mark the major contact site between mitotic spindle microtubules and chromosomes (Burke and Stukenberg 2008). Regulation of the microtubule dynamics by centrosomes at the cell poles and kinetochores at microtubule-kinetochore interfaces ensures that chromosomes will traffic and segregate appropriately during cell division (Weaver and Cleveland 2006; Morrison 2007; Musacchio and Salmon 2007; Burke and Stukenberg 2008; Decordier et al. 2008). Prior to sister chromatid separation and chromosome segregation, paired sister chromatids move toward a metaphase plate and remain there until anaphase promoting complex (APC) inhibition is relieved (Burke and Stukenberg 2008). A cohesin ring keeps one sister kinetochore attached to the other until entry into anaphase (Decordier et 25 al. 2008). For the mitotic checkpoint signal at pro-metaphase to be relieved, each kinetochore must exist in a spindle-attached state, and sister kinetochores must be bi-oriented with one kinetochore being drawn by the spindle toward one centrosome while the sister is drawn to the opposite pole. Bi-orientation of sister chromatids causes the mitotic checkpoint apparatus to “sense” tension between kinetochore pairs. A single pair of sister kinetochores that are improperly-attached or wrongly-oriented is sufficient to prevent entry into anaphase. Proteins involved in maintenance and/or relief of the mitotic checkpoint prior to anaphase include Mad1, Mad2, CenpE, CenpF, BubR1, AuroraB, INCENP and Survivin, just to name a few (Brown et al. 1996; Yao et al. 2000; Mao et al. 2003; Mao et al. 2005). Once all sister chromatids are bi-oriented, APC inhibition is relieved, the cohesin ring complex is cleaved, sister chromatids begin segregating toward opposite spindle poles, cytokinesis occurs, and daughter cells almost always have inherited identical cellular DNA content (Burke and Stukenberg 2008; Decordier et al. 2008). HPV DNA replication and early region message splicing Replication of the HPV genome is one portion of the virus life cycle that can be re-capitulated fairly well in monolayer cells. For HPV genome replication to initiate successfully and then be completed, the E1 and E2 proteins must associate with each other and also their respective recognition sites in the papillomavirus (PV) URR. The normal relative abundance of E1 and E2 is dictated by alternative splicing of their common primary transcript as the E1E2 operon is transcribed. The E1 and E2 messages are minor RNA species though. The relative amount of a third product of the same primary tran- 26 script that is translated as the E1^E4 protein is much more abundant than the E1 and E1^E2 messages (Figure 2). Expression of E1 and E2 from separate plasmids prevents E1^E4 production and maximizes ori plasmid replication efficiency. Although transfection of an 11-E1E2 operon plasmid cuts expression of E1 and E2 drastically, the relative ratios of these replication proteins remain wild-type, and ori plasmid replication is still very strong. Use of an 11-E1E2 plasmid for replication studies supports robust replication of ori plasmids at levels similar to when separate 11-E1 and 11-E2 plasmids are used. E1, E2, and E1^E4 each cause cytotoxicity in transfected cells. Attempts by others in our lab to make inducible E1-expressing cells have failed. Even the slightest induction of E1 expression caused all cells to die. Inducible E2 cell lines have been generated but also exhibit substantial signs of cytotoxicity. The only way to be certain that the relative abundance of E1 and E2 is held constant in replication studies is to utilize the E1E2 operon. Utilization of the E1E2 operon leads to the production of E1^E4 which is undesirable and also cytotoxic, but the ability to normalize the ratio E1 expression ratio to that of E2 is very helpful. 27 Experimental aims and questions addressed in this project The goals of the present work are: 1) Develop plasmids that can be tracked by direct, in situ immunofluorescence imaging. 2) Visualize the partitioning of HPV ori-containing plasmids in dividing cells. 3) Determine whether the E2BSs which facilitate efficient PV DNA replication must also interact with E2 during mitosis to ensure that HPV genomes segregate into daughter cells evenly. 28 EXPERIMENTAL WORK Tracking DNA inside mammalian cells using Gal4-DBD, TetR, and rtTA3 GFP fusion proteins Hybridization conditions for visualizing discrete nucleic acid targets by fluorescence in situ hybridization (FISH) often disturb or destroy other cellular features of interest. As the Kd of the lac repressor-lac operator complex is 10-13M, integrated tandem arrays of at least 256 lac operator sites have permitted non-destructive, direct, fluorescent in situ visualization of target DNA in both fixed and living cells using chimeric fusions between the repressor and a tracer such as green fluorescent protein (GFP) (Miller and Reznikoff 1980; Robinett et al. 1996; Tsukamoto et al. 2000). These techniques can also yield better resolution than can FISH. More recently, it was reported that single membrane-bound protein molecules could be visualized in bacteria by fluorescence (Yu et al. 2006). However, the signal-to-noise ratio for most experimental purposes requires binding site arrays that are quite large. For in situ DNA localization experiments in bacteria involving detection of one or more discrete genetic loci, it has been the convention to use ~240 tandem copies of the operator (Lau et al. 2003; Wang et al. 2008a; Wang et al. 2008b). In yeast, in situ localization of DNA has relied upon as few as 112 copies of the lac operator (Michaelis et al. 1997). In mammalian eukaryotic cells, where cell size and volume are greater, refractive scattering of fluorescent signal has challenged investigators. In some instances, researchers relied upon clonal cell lines where multiples of the 256-copy lac operator array are stably integrated at a single genetic locus. With a plas- 29 mid containing 256 lac operators, cell lines have been isolated and tested with singlelocus insertions of either 10, 50, or 1000 copies of a lac operator-containing plasmid (Tsukamoto et al. 2000; Janicki et al. 2004). The net effects were cell lines containing a chromosomal insertion at a single locus of either 2.56*103, 1.28*104, or 2.56*105 lac operator sites. This is not feasible for studies where extrachromosomal, replicating HPV ori-containing plasmids are to be followed. Since limiting any HPV tracking plasmid to ~8kb in size is preferable, ~40-50 copies of an appropriate operator site would be ideal. Researchers have used either the Lac Repressor (LacI) or the TetR proteins fused to either CFP, GFP, or YFP fluorescent proteins for visualizing DNA (Robinett et al. 1996; Michaelis et al. 1997; Tsukamoto et al. 2000; Ono et al. 2004; Ebersbach et al. 2005; Higuchi and Uhlmann 2005; Fiebig et al. 2006; Yu et al. 2006; Wang et al. 2008a; Wang et al. 2008b). Our lab has previously used the yeast Gal4 DNA-binding domain (Gal4-DBD)-GFP fusion protein to observe HPV ori-containing plasmid which also contains 40 copies of a Gal4 binding site (Gal4BS) array (Van Tine et al. 2004). Using the high-sensitivity, high-resolution DeltaVision Spectris microscope system, I found that the Gal4-DBD/Gal4BS system has a serious drawback and can produce artifacts. In the absence of a Gal4BS array, Gal4-DBD-GFP was able to form very fine foci around the kinetochores and associate with cellular chromatin during mitosis. In light of this observation, I abandoned the Gal4-DBD/Gal4BS system for my further studies of HPV DNA replication and segregation. The tracking protein needs to bind reversibly to a very specific DNA sequence to allow ori-specific replication and only be induced to bind its recognition sequence until just prior to fixing cells for use in microscopy. These types of in situ DNA detection experiments remain technically and experimentally challenging. The 30 lac repressor has constitutively-high affinities for its operator nucleotide sequence, and this is inappropriate for our experimental purposes. Researchers have turned to the family of tetracycline(tet)/doxycycline(dox) inducible proteins. Among those proteins that can inducibly bind to or release from the tet operator are TetR, tTA, and rtTA. The usefulness of the Tet repressor-GFP fusion protein (TetR-GFP) and the tetracycline operator (tetO) array have been described and discussed in depth. Under opposite conditions, VP16 fusion proteins known as tTA and rtTA proteins are each capable of binding to promoter-proximal tetO sites to activate transcription. TetR and tTA exhibit high affinity for tetO sites only in the absence of doxycyclin (dox) while rtTA only binds tetO sites efficiently after exposure to dox (Gossen and Bujard 1992; Michaelis et al. 1997). Compared to rtTA in mammalian cells, a variant known as rtTA3 with the S12G, F86Y and A209T substitutions exhibited 5-fold better activity and 25-fold greater sensitivity, in response to the addition of dox (Das et al. 2004). For any tagging protein to be useful for visualizing HPV DNA replication and/or segregation, rtTA3 and TetR must not interfere with the replication competency of the URR-TetO plasmid when the wild-type relative amounts of E1 and E2 are expressed via co-transfected HPV-11-E1E2 operon-containing plasmid. Expression of neither TetR nor rtTA appeared toxic to the cells. Neither dox-induced nor the uninduced TetR and rtTA proteins interfered with the replication efficiency of the control plasmid containing only the HPV 11-URR, as assessed by Southern blot hybridization. In contrast, when either TetR or rtTA was allowed to bind to the tetO sites in the URR-containing plasmid for the entirety of the experiment, replication was severely depressed. 31 To study the segregation of replicated HPV plasmid DNA, I tested TetR and rtTA. Doxycycline was used as the inducer drug in lieu of tetracycline because it is water-soluble and more stable. In their high-affinity conformation, both TetR and rtTA3 are capable of binding to the same TetO DNA recognition sequence. The tetO DNA sequence recognized by TetR and rtTA3 is as follows: +9..+8.......+6......+4........+2........0........-2........-4.........-6.........-8...-9 5’- T C C C T A T C A G T G A T A G A 3’- A G G G A T A G T C A C (Gossen and Bujard 1992) T A T C T G A -3’ C T -5’ 32 Pre-requisites and experimental constraints to carefully consider during development and testing of HPV tracking plasmid systems Expression of the HPV E1, E2, and E4 proteins impacts cell cycle progression and can be cytotoxic The abundance of E1 protein relative to that of E2 is normally dictated by alternative intragenic and intergenic RNA splicing from primary transcripts spanning multiple cistrons. When ectopically-expressed, the use of the HPV-11 E1E2 operon allows both HPV replication proteins to be expressed from a single plasmid. In addition, the E1^E4 mRNA and protein are the predominant products. Using the more-stable blue fluorescent protein variant referred to as eBFP2, I generated the eBFP2-11-E1E2 expression plasmid which encodes predomantly the eBFP2-E1^E4 protein along with eBFP2-E1 and the native E2 protein (Ai et al. 2007). I was able to identify the cells successfully transfected with the plasmid expressing the abundant cytoplasmic eBFP2-E1^E4 protein. To assess whether the eBFP2-tagged 11-E1E2 operon can support the replication of an HPV ori-containing plasmid, I conducted transient replication assays in 293 cells. If positive, the eBFP2-11-E1E2 operon would be useful for visually verifying that E1 and E2 are both expressed in cells being imaged. If E1 and E2 expression from the same plasmid is feasible, the total number of plasmids being transfected during experiments can also be reduced. For all HPV ori plasmid replication assays being presented, low molecular weight plasmid was isolated 60 hours post-electroporation. Since each transfected plasmid originated and was purified from DNA methylase-positive bacterial cell strains, all plasmids were methylated at the time of transfection and therefore susceptible to di- 33 gestion by DpnI, whereas ori DNA newly-replicated in 293 cells will not be methylated and will be resistant to Dpn1 digestion. To assess ori replication in 293 cells, half the plasmid purified from each transfection was digested by a single restriction enzyme (oddnumbered lanes) while the remaining half was simultaneously-treated with DpnI endonuclease (even-numbered lanes). All non-replicated plasmid will be susceptible to degradation by DpnI. By running non-DpnI-treated (odd-numbered lanes) and DpnI-treated samplesfrom the same experiment in adjacent lanes, bands representing non-replicated and replicated plasmid DNA will be readily-visible upon Southern blot hybridization. Whenever high-level replication of the plasmid is observed, the sum total intensity of ori band(s) should be high and almost identical in both the DpnI and non-DpnI-treated lanes. This assay is highly-sensitive and can show whether expressed E1 and E2 proteins are functional and at adequate levels. This assay can also show whether ori plasmid replication is affected by the expression of fluorescent fusion proteins like TetR-eGFPnls or rtTA3-eGFP-nls or the treatment with a drug such as doxycyclin (dox). With the TetO array-containing HPV ori plasmid, we can also determine whether ori plasmids can replicate while TetR-eGFP-NLS is bound to the plasmid via a tetO array. peBFP2-11-E1E2 plasmid supports efficient ori-specific plasmid replication Fig. 4a shows that peBFP2-11 E1E2 supported replication of a minimal ori plasmid (spanning HPV-11 nts 7730-99) as well as when the HPV-11 eGFP-E1 and native E2 proteins were expressed from separate vectors. Expression of the low levels eBFP1-E1 and the native E2 replication proteins at their natural relative ratio (lanes 3,4) was almost 34 as good as when the proteins were individually over-expressed to high levels at nonnative ratios (lanes 1,2). To assess whether the peBFP2-11-E1E2 would be suitable for experimental studies of HPV DNA replication and segregation, it was co-transfected with a plasmid which contained the HPV-11 URR (nts 7070-99) spanning the full ori. Fig. 4b showed peBFP211-E1E2 was able to support the replication of plasmid containing only the HPV URR(lanes 1-6). 35 Figure 4: The replication efficiency of HPV-11 ori plasmids in the presence or absence of TetR and doxycycline. a) 293 cells were transiently-transfected with a minimal HPV11 ori plasmid. Additionally, plasmids encoding HPV-11 GFP-E1 or native E2 proteins (lanes 1, 2) or, alternatively, the eBFP2-11-E1E2 plasmid were co-transfected (lanes 3, 4). Robust ori replication was observed under both conditions b) 293 cells were transiently transfected with HPV-11 ori containing plasmid with (lanes 7-12) or without (lanes 1-6) an array of 48 copies of TetO sequence. Additional cells were transientlytransfected with a non-ori-containing plasmid with an array of 48 copies of TetO sequence (lanes 13-18). In each experiment, the eBFP2-11-E1E2 expression plasmid was also co-transfected. As indicated (+ or -), in some experiments, the expression vector of TetR-eGFP-nls was co-transfected. Cells were cultured in DMEM either in the presence or absence of 0.5 μg/ml doxycycline, also indicated (+ or -). Low-molecular weight 36 plasmid DNA was purified 60 hrs post-transfection, digested with ApaLI alone (oddnumbered lanes) or along with DpnI (even-numbered lanes), and analyzed by Southern blot hybridization. HindIII-cut, gel-purified 11-URR/eBFP2-11-E1E2/48-copy tetO array plasmid served as the template to generate the radio-probe. A single asterisk points to the linearized DNA band which corresponds to the ori plasmid (lanes 1-6), the 48-copy TetO array plasmid (lanes 13-18) or the plasmid with both the ori and the TetO array (lanes 712). A double-asterisk marks the open-circle DNA band of these same plasmids due to incomplete ApaLI digestion. 37 Tight association of a GFP fusion protein in a large array to an HPV ori plasmid can impede ori replication. Whether TetR-eGFP-nls can alter the ability of HPV-11 URR-only plasmids to replicate efficiently was next assessed by Southern (Figure 4b) using low molecular weight plasmid purified from 293 cells 60 hours post-electroporation. Compared with baseline replication levels (Figure 4b, lanes 1,2), URR-containing plasmid replication was unaffected by TetR-eGFP-nls expression either alone (Figure 4b: lanes 3,4) or together with dox treatment (Figure 4b, lanes 5,6). Replication of the plasmid containing both the HPV URR and the TetO array sequence (Figure 4b, lanes 7,8) was almost identical to the replication of the simple URR-containing plasmid (Figure 4b, lanes 1-6). Replication of the plasmid containing both the ori and TetO array was inhibited substantially when TetR-eGFP-nls was expressed (Figure 4b, lanes 9,10). Inhibition of replication by TetR-eGFP-nls was reversed by culturing cells in media containing doxycycline (dox) through the entire incubation period to reduce the affinity of TetR for the TetO array (Figure 4b, lanes 11,12). Observing replicated, 11-URR / 48-copy tetO array plasmid in situ by fluorescence microscopy would be impossible unless dox can be removed to allow the binding of GFP-TetR to tetO sites. However, restoring the affinity of TetR-eGFP-nls for TetO sites by rinsing dox out of cells proved unfeasible due to the tight binding of TetR to dox. Due to this limitation and drawback of TetR-eGFP-nls, a different strategy for fluorescently-tagging our tracking plasmids would be essential. 38 To image HPV ori-containing tracking plasmid DNA, I felt that rtTA, which only binds to TetO sites in the presence of dox, might be better-suited. An rtTA variant known as rtTA3 has lower-level background affinity for TetO sites in the absence of dox. I generated an rtTA3-eGFP-nls expression plasmid and tested its impact on replication of HPV ori plasmids in 293 cells (Figure 5). Compared with baseline replication levels (Figure 5 lanes 1,2), URR-only plasmid replication was unaffected by rtTA3-eGFP-nls expression either alone (Figure 5, lanes 3,4) or together with doxycycline (dox) treatment (Figure 5, lanes 5,6). Replication of HPV URR / 48-copy TetO array plasmid (Figure 5, lanes 7,8) was very similar to replication of URR-only plasmid (Figure 5, lanes 1-6). Expressing rtTA3-eGFP-nls did not impair replication of the 11-URR/48-copy TetO array plasmid (Figure 5, lanes 9,10), unless dox had also been added to cell culture media throughout the incubation period (Figure 5, lanes 11,12). No replication of the TetO array-only plasmid was detected (Figure 5, lanes 13-18). 39 Figure 5: The replication efficiency of HPV-11 URR-containing plasmids in the presence or absence of rtTA3-eGFP-nls and doxycycline. 293 cells were transiently transfected with HPV-11 URR-containing plasmid with (lanes 7-12) or without (lanes 1-6) an array of 48 copies of TetO sequence. Additional cells were transfected with 48-copy TetO array-only plasmid lacking the ori (lanes 13-18). In each experiment, the eBFP211-E1E2 expression plasmid was also co-transfected. As indicated (+ or -), in some experiments, the expression vector of rtTA3-eGFP-nls was co-transfected. Cells were cultured in DMEM in the presence or absence of 0.5 μg/ml doxycycline(dox), also indicated (+ or -). Low-molecular weight plasmid DNA was purified 60 hrs post-electroporation and analyzed by Southern blot hybridization. All purified plasmid was digested with ApaLI. HindIII-cut, gel-purified 11-URR/eBFP2-11-E1E2/48-copy tetO array plasmid was used as template to generate radiolabelled probe. A single asterisk points to the linearized DNA band that corresponds to the URR-only plasmid (lanes 1-6), the 48-copy TetO array-only plasmid (lanes 13-18) or the plasmid with both the URR and the TetO array (lanes 7-12). A double-asterisk marks the open-circle DNA band of these same plasmids due to incomplete digestion by the ApaLI. 40 Tracking plasmid detection sensitivity during mitosis requires optimized expression of the eBFP2-11-E1E2 operon and rtTA3-eGFP-nls To reduce the overall number of different plasmids that had to be electroporated for each experiment, I attempted to consolidate plasmids by driving the expression of either eBFP2-11-E1E2 or rtTA3-eGFP-nls from the HPV-11-URR / 48-copy TetO array plasmid. In theory, the 11-URR / eBFP2-11-E1E2 / 48-copy TetO array plasmid should have been able to function as an autonomously-replicating tracking plasmid. Also in theory, the 11-URR / rtTA3-eGFP-nls / 48-copy TetO array plasmid should have been able to function as a self-detecting tracking plasmid. Being able to use either of these strategies would have meant that a maximum of two plasmids would need to be electroporated for each experiment. This would be, in theory, much superior to utilizing up to three plasmids for each transfection. The two-plasmid approach proved, however, to be problematic. If my replicon were to be useful, I needed to be able to have eBFP2-11-E1E2 operon expression reach ideal levels at the same time that the rtTA3-eGFP-nls expression was optimal. Since the tracking plasmid is replicating exponentially, the gene copy number of the eBFP2-11-E1E2 operon is also rising exponentially. This causes a rapid spike in E1, E2, and E1^E4 expression levels during the middle and later parts of the experiment. At the same time, the plasmid copy number of the rtTA3 expression plasmids holds steady or is falling. This means that optimal expression levels of the HPV replication proteins and the GFP plasmid-tagging protein could not be attained simultaneously. The expression levels of the E1E2 proteins varied from one cell to another depending on the extent of plasmid replication, and cytotoxicity was observed frequently. When E1, 41 E2, and E1^E4 were being expressed, cells on average were expressing extremely-high levels of the proteins. Replication was extensive, but almost no mitotic cells could be found, probably because of the ability of the E1^E4 protein to sequester cyclin B/cdk1 in the cytoplasm (Knight et al. 2004; Knight et al. 2006). Thus, the replicon was unusable and expression of the E1E2 operon needed to be driven from a distinct plasmid at an optimized amount. I did not attempt to use rtTA3-eGFP-nls in a self-detecting tracking plasmid because I anticipated difficulty titrating rtTA3-eGFP-nls expression levels, based on my experience with TetR. Overall, this approach would lead to the production of too much of the GFP fusion protein. Just as carefully-controlled levels of eBFP2-11-E1E2 are central to minimizing cytotoxicity, carefully-titrated expression levels of rtTA3-eGFP-nls are vital for maximizing tracking plasmid detection sensitivity. When rtTA3-eGFP-nls was expressed from one dedicated plasmid and the eBFP2-11-E1E2 operon from another, I was able to achieve satisfactory titration of rtTA3-eGFP-nls, E1, E2, and E1^E4 expression plasmids for transfection. All imagery presented hereafter was acquired by using this approach. I observed less cell-to-cell variation in the amount of rtTA3 being expressed. During the titration experiments, cells expressing different amounts of the GFP fusion protein were imaged, but the best sensitivity and resolution were obtained in cells where a moderate amount of rtTA3-eGFP-nls protein was expressed. Regardless of how wellor poorly-titrated rtTA3-eGFP-nls expression was, big, bright green spots of replicated DNA were easily-detected in the nuclei of non-dividing cells. It was in mitotic cells that the benefit of carefully-titrating rtTA3-eGFP-nls expression became evident. During mi- 42 tosis, there are hundreds or thousands of dim, discrete spots. Over-expression of rtTA3eGFP-nls masks detection and often prevents resolution of the small, dim foci that each might represent as few as one copy of replicated HPV tracking plasmid. The only way to better-normalize cell-to-cell expression levels of rtTA3-eGFP-nls would be to generate stable cell lines. I have generated stable, clonal 293-rtTA3-eGFP-nls cell lines that express the fusion protein to different levels but have yet to explore their utility. (I thought you had to supplement the cell line with additional expression vector. I did with the pooled TetR-eGFP cells that were non-clonal, but with the clonally-selected rtTA-eGFPnls cell lines I did not have to) By optimizing peBFP2-11-E1E2 used in the transfection, the mitotic index was greatly-improved relative to early pilot experiments where E1 and E2 had been expressed from two separate plasmids, and I was able to image large numbers of mitotic cells with ease. As already stated, I attempted to use both TetR-GFPnls and rtTA-GFPnls for imaging purposes to follow the HPV tracking and control plasmids. Attempts to image plasmids with TetR-GFPnls were entirely unsuccessful. Due to the tightness with which TetR binds to dox, it was unfeasible to wash the drug out of treated cells. Even after changing the DMEM multiple times at long intervals, only dim spots in interphase cell nuclei were apparent, and no discernible foci were detectible in mitotic cells. TetR-GFP is unsuitable for use in experiments where URR-TetO plasmid needed to be visualized inside cells capable of supporting HPV DNA replication. 43 Clear visualization of HPV DNA localization is obtained when rtTA3-eGFP-nls binds to arrays of tetO sites embedded in tracking plasmids. The cellular distribution pattern of the 48-copy TetO array-containing plasmid Make it clear that all experiments were conducted by transfection or cotransfection of one or more plasmids. (as opposed to a cell line which already expresses rtTA3-eGFP-nls The TetO array negative control plasmid (pTetO) contains 48 tetO sites. When this plasmid was transfected into 293 cells together with the rtTA3-eGFP-nls expression plasmid, treatment with doxycycline (dox) for four hours prior to fixing the cells allows this control plasmid to be visualized as bright spots. Depending on spot intensities and sizes, each spot from experimental (Figure 8) and control cells (Figure 6) can be grouped into one of three categories. In dox-treated cells where TetO array-containing plasmid either with (Figure 8) or without (Figure 6) the HPV-11 URR has been transfected, spots are usually: 1) small & dim, 2) larger & brighter, or 3) humongous & extremely bright. The pTetO control plasmid was not generally seen as small, dim spots. When I fixed these control cells 36 hours post-electroporation, bright or dim small spots were not seen in the vast majority of the nuclei. Instead, almost all pTetO plasmids were detected in the cytoplasm as multiple clusters that were large and extremely-bright (Figure 6, interphase #2 panels). In interphase cells fixed 60 hours post-electroporation, there was at most one or two cytoplasmic cluster(s) of control plasmid (data not shown). In cells where the 11URR/TetO array plasmid (p11-URR-TetO) and the rtTA3-eGFP-nls expression plasmid were co-transfected without the eBFP2-11-E1E2 operon expression plasmid, the fate of 44 the plasmid was identical to the TetO array-only control plasmid (data not shown). The same is also true when the TetO array-only control plasmid cells were also co-transfected with rtTA3-eGFP-nls expression plasmid and the eBFP2-11-E1E2 operon plasmid (data not shown). We believe the vast majority of TetO array-containing plasmids in these control experiments have coalesced to form clusters due in part to recombination or that they were sequestered to certain domains in interphase cells where they are degraded or lost during subsequent rounds of the cell cycle. A portion of the cytoplasmic plasmid aggregate clusters were seen co-localizing with microtubules (Figure 6, interphase cell #2, column A). 45 Figure 6: The distribution pattern of the negative control, 48-copy TetO array plasmid in doxycycline (dox)-treated 293 cells. 293 cells were transiently transfected with rtTA3eGFP-nls expression plasmid and 48-copy TetO array plasmid. 36 hours post- electroporation, cells were fixed after being treated with 2 μg/mL dox. Fixed cells were stained and then imaged using z-stepping. β-tubulin (Columns B & D; pseudo-colored 46 red) and chromatin (Columns C &D; pseudo-colored blue) were stained using Cy3conjugated monoclonal antibody and DAPI, respectively. The TetO array plasmid is visible as green foci. To maximize the signal-to-noise ratio, background, non-specific signal was subtracted from the red, green, and blue image channels. rtTA3-eGFP-nls not associating with a TetO site on the negative control plasmid resulted in a haze in each image. The hazy rtTA3-eGFP-nls signal was subtracted from the green channel, while ensuring that all green foci remained intact. This maximized the visibility and contrast of green foci. The background-subtracted image file of each cell was then used to generate the maximum-intensity projections seen here in the figure panels. The TetO array plas- mid produces bright green foci (Columns A-D). Image analysis software calculated whether TetO array plasmid foci and β-tubulin did or did not co-localize. Partial or total colocalization of green foci with β-tubulin is displayed as magenta or white (Column A). Magenta represents the majority of plasmid-microtubule co-localization events while white indicates where the brightest plasmids are co-localized with microtubules. Each scale bar represents 5μm. 47 Plasmid does not associate with transiently-expressed rtTA3-eGFP-nls to form foci until cells are treated with doxycycline. Whenever rtTA3-eGFP-nls was expressed from transiently-transfected plasmid in 293 cells in the absence of TetO sites, no foci were observed in any cells, whether or not they were mitotic. The GFP fusion protein was evenly-diffused inside the nuclei of interphase cells and throughout mitotic cells, regardless of whether the cells were treated with doxycycline (dox) (data not shown). No foci were detected in rtTA3-eGFP-nlsexpressing, dox-treated 293 cells where the URR-only negative control and eBFP2-11E1E2 expression plasmids were co-transfected. Observations were virtually-identical when the TetO array-only negative control and eBFP2-11-E1E2 plasmids had been cotransfected and the cells were fixed without having been exposed to dox (data not shown). (you need to shown one such control image, no dox, no foci. The control cells I show in Figure 7 are cells with 11-URR-TetO plasmid that have not been exposed to dox. As this is actually a more-stringent control, is Figure 7 sufficient or should I also show that this is the case with the TetO-only plasmid?) Figure 7 shows cells transientlytransfected with the p11-URR-TetO, peBFP2-11-E1E2, and rtTA3-eGFP that have not been exposed to dox. For rtTA3-eGFP-nls to bind to any TetO array-containing plasmid and yield detectable green foci, dox was absolutely essential. For the purpose of presentation, I did not subtract all rtTA3-GFP haze from the control cells in Figure 7 because little or no GFP signal would remain to be shown. In Figures 6 and 8, rtTA3GFP haze was subtracted as thoroughly as possible. Aside from that, all imaging conditions, image processing and data analysis were standardized were held constant from control to ex- 48 perimental cells ensure that my interpretations of all results were completely unbiased. Each and every positive mitotic cell I encountered was imaged. When I combine the controls I have shown (Figure 7) and/or discussed in this section with controls shown (Figure 6) and/or discussed in the previous section, I can comfortably place a high degree of confidence in the experimental results to be discussed next. 49 Figure 7: The inability to detect replicated 11-URR / 48-copy tetO array plasmid in 293 cells not treated with doxycycline (dox). 11-URR/48-copy TetO array plasmid was electroporated into 293 cells together with the rtTA3-eGFP-nls and eBFP2-11-E1E2 expression plasmids. Cells were fixed 36 hours post-electroporation. Cells were stained for β- 50 tubulin (Columns B & D; pseudo-colored red) and chromatin (Columns C &D; pseudocolored blue) using Cy3-conjugated monoclonal antibody and DAPI, respectively. rtTA3-eGFP-nls is shown in green. Background signal in the red and blue channels was subtracted with the same stringency as with cells in figures 4 and 6. No foci were detectable in these cells untreated with dox. Rather, the rtTA3-eGFP-nls signal was diffused throughout the nuclei in interphase cells or the cell in mitotic cells in these negative control cells. This signal was not subtracted. Only extracellular rtTA3-eGFP-nls true background signal was subtracted. Image analysis software determined where green haze and β-tubulin did or did not co-localize. Partial or total co-localization of the green rtTA3eGFP-nls haze with β-tubulin is displayed as magenta (Column A). Scale bars in each image represent 5μm 51 The 11-URR/TetO array-containing tracking plasmid efficiently-targeted to microtubules during mitosis. The experimental 11-URR/48-copy TetO array tracking plasmid was cotransfected together with the peBFP2-11-E1E2 and prtTA3-eGFP-nls expression plasmids, and the 293 cells were treated with doxycycline(dox) prior to fixation. Very different results (Figure 8) were observed when compared to controls (Figure 6). In Figure 8, the vast majority of green foci were similar in intensity. Some tracking plasmid clusters or aggregates were observed, but they were smaller and dimmer than the ultra-bright, large aggregates seen in Figure 6. A breakdown of what I observed during different phases of the cell cycle follows. The p11-URR-TetO tracking plasmid showed a highly-dynamic localization pattern in mitotic cells. In interphase cells (Figure 8, row #1), p11-URR-TetO replicated efficiently. I observed multiple bright foci that varied in size from small to extremelylarge. Each focus was equivalent to multiple copies of tracking plasmids. I suggest that they are replication centers, as no or few such foci were observed in the control experiments described above. At the nuclear periphery, almost every replication centers showed co-localization with a single microtubule at a single, discrete point (Figure 8, Column A, row #1). At prophase (Figure 8, row #2), most of the large replication centers with dense amounts of p11-URR-TetO had dissociated to become very small and discrete spots with relatively uniform brightness. The couple larger clusters of plasmids may have represented replication centers that had yet to disperse. This same pattern was observed at pro-metaphase (Figure 8, Prometaphase cell, Columns A and B). 52 During metaphase, maximum co-localization of the 11-URR-TetO tracking plasmids with mitotic microtubules was observed (Figure 8, Metaphase cell, Columns A, B). Additionally, the large clusters of un-dispersed, replicated DNA visible at pro- and prometaphase had dissipated and only discrete, tiny green foci were observed. At anaphase, a substantial fraction of the tracking plasmids co-localized with the mitotic spindle fibers (Figure 8, Anaphase cell, Columns A and B). Most notably, the 11URR-TetO plasmids co-localized with the central spindle fibers as well as microtubules proximal to the centrosomes. Some of the plasmids co-localized with the central spindle fibers were slightly larger than green foci observed at metaphase. During telophase and into cytokinesis (Figure 8, telophase and telophase/cytokinesis cells, Columns A and B), much of the tracking plasmids still co-localized with microtubules. A greater number of medium-sized spots were observed than was the case during anaphase. At the completion of cytokinesis, the majority of the ori plasmids had successfully migrated back into the nuclei of each daughter cell. We observed equal partitioning of the 11-URR-TetO tracking plasmid in this and the other mitotic cells shown in Figure 8. Plasmids remaining at the cleavage furrow and midbody were likely associated with E2 (Dao et al., 2006) and might have eventually migrated back into the nucleus. Alternatively, they might be lost upon mitosis. When I began to study the localization pattern of p11URR-TetO (Figure 8, Column A, rows 1-8) and pTetO (Figure 6, Column A, rows 1-8) relative to microtubules in both interphase and mitotic cells, a technical challenge emerged that required utilization of more intensive image processing and analysis tools. Each image panel in figures 6-8 is a maximum intensity projection. With maximum-intensity projections, a simple overlay 53 of the red (β-tubulin) and green (rtTA3-eGFP-nls) yields yellow when there is colocalization, but it is challenging to distinguish red and various shades of yellow accurately (Column B, Figures 6 and 8). Another issue with maximum-intensity projections where only the tubulin (red) and rtTA3-eGFP-nls (green) are shown is even more serious (Figures 6-8, Column B). Yellow does not necessarily mean that the red and green are co-localized when looking at a particular X-Y coordinate in a maximum-intensity projection. As a hypothetical example of this, non-spindle-associated green foci of either pTetO (Figure 6) or p11URR-TetO sometimes show up as yellow if there is a microtubule in any separate z-plane that is either above or below that particular green spot. When the maximum-intensity projections of the red and green channels are generated, the x-y coordinate marking the plasmid will appear yellow even though the plasmid is not colocalized with a microtubule. To solve this problem, the co-localization analysis tool in SoftWoRx was utilized to identify co-localization and produce a visual rendering of whether microtubules associate with pTetO and p11URR-TetO (Column A, Figures 6 and 8 respectively). The co-localization data was assigned by me to the blue channel. When I overlaid blue co-localization image data with the green channel, I was left with colocalization spots that were different shades of turquoise. As with viewing different shades of yellow, an image displaying greens and turquoises would have been less than optimal. To circumvent this problem, I duplicated the blue co-localization channel inside the red image channel to yield magenta spots wherever green foci and β-tubulin were colocalized. Co-localization spots that are magenta are easy on the eyes and print extremely well. Overlaying the magenta atop the green image channel yields can yield an 54 image with mixtures of different-colored spots that are green (not co-localized with βtubulin), magenta (dim green foci co-localized with β-tubulin), and/or white (bright-green foci co-localized with β-tubulin). In control images where dox was not added, p11URR-TetO is invisible, and the free rtTA3-eGFP-nls is evenly dispersed throughout the nuclei of interphase cells and the entirety of mitotic cells (Figure 7). The magenta in these control cells does not depict plasmid co-localization (Figure 7, Column A). Rather, the magenta renders each instance when the free rtTA3-eGFP-nls is co-localized with the β-tubulin (Figure 7, Column A). In figures 6 and 8, non-plasmid-associated rtTA3-eGFP-nls signal was subtracted as thoroughly as possible. Residual green signal is overwhelmingly-specific and is comprised of the green plasmid foci, clusters, and aggregates plus whatever tiny bit of residual rtTA3-eGFP-nls remained after the stringent subtraction. By this approach, I was able to study potential co-localization of pTetO and p11URR-TetO accurately, and subsequently, I was able to present the co-localization data clearly for all viewers to see (Figures 6-8, Column A). 55 Figure 8: The distribution pattern of replicated 11-URR/48-copy TetO array tracking plasmid in doxycycline (dox)-treated 293 cells. 11-URR/48-copy TetO array plasmid 56 was electroporated into 293 cells together with the rtTA3-eGFP-nls and eBFP2-11-E1E2 expression plasmids. 36 hours post-electroporation, cells were fixed following a 4-hour treatment with 2 μg/mL dox. Fixed cells were stained and then imaged using z-stepping. β-tubulin (Columns B & D; pseudo-colored red) and chromatin (Columns C &D; pseudocolored blue) were stained using Cy3-conjugated monoclonal antibody and DAPI, respectively. The 11-URR/TetO array plasmid is visible as green foci. Image processing was as described in the legend to Figure 4. The 11-URR/TetO array plasmid produces distinct, green foci (Columns A-D). Shown in column A, image analysis software calculated whether TetO array plasmid foci and β-tubulin did or did not co-localize. Magenta (column A) represents the majority of plasmid-microtubule co-localization events while white indicates where the brightest plasmids are co-localized with microtubules. Each scale bar represents 5μm. 57 DISCUSSION I have developed a system to visualize replicated HPV-11 ori plasmids and have followed the plasmid distribution through the cell cycle. My conclusion was that the ori tracking plasmid colocalized with mitotic spindles, enabling equal partitioning into daughter cells. This conclusion is in agreement with previous reports from our lab (Van Tine et al. 2004; Dao et al. 2006). In those reports, either Gal4-eGFP and native HPV-11 E2 protein bound to the HPV-11 ori plasmid or GFP-HPV-11 E2 alone was used as reporter, not replicating tracking plasmids. From these three studies, it is also apparent that the binding of 11-E2 to p11URR-TetO (or any other URR-containing tracking plasmids I characterized) is necessary to allow segregation of the tracking plasmid during mitosis. We suggest that this association with mitotic spindles provides a mechanism by which viral DNA can persist in dividing cells. It should also be mentioned that there is evidence that the E2 proteins of other HPV types, primarily those of the beta HPVs, associate with mitotic chromosomes, rather than the mitotic spindles (Skiadopoulos and McBride 1998; McBride et al. 2004; You et al. 2004; Baxter et al. 2005; Brannon et al. 2005; McPhillips et al. 2005; Zheng et al. 2005; McPhillips et al. 2006; Oliveira et al. 2006; Parish et al. 2006; Yu et al. 2007; Cardenas-Mora et al. 2008) I note that the transient replication system used in my study is not comparable to viral DNA replication which occurs in the basal cells in organotypic raft cultures. Rather, it is equivalent to viral DNA amplification in post-mitotic differentiated cells. In the longliving basal cells where HPV DNA must partition properly to persist, there are somewhere between 50 and 200 copies of HPV genomic DNA per cells. The levels of viral E1 58 and E2 proteins expressed from the homologous promoter would be much reduced relative to the ectopically expressed viral protein in 293 cells. We theorize that the host protein to which E2 interact to effect the attachment of ori tracking plasmid to microtubules can be overwhelmed by the high levels of the E2 protein. Although ChlR1 and MKlp2 have both been shown to associate with E2 proteins of multiple PV types previously, it is unclear how many molecules of each of these proteins are present per mitotic cell. Their possible saturation by the over-expressed E2 may account for some or all of the nonmicrotubule-associated HPV tracking plasmids as well as those found in the cleavage furrow and midbody (Parish et al. 2006; Yu et al. 2007). These plasmids might be lost upon cytokinesis. Despite this caveat, our observations of the p11-URR-TetO tracking plasmid in the transient replication system do provide relevant insights (Figure 8). Our new tracking plasmid system is however not perfect. Exposure times for capturing the green channel were 5 seconds, and hence, I was not able to perform live cell imaging. My new tracking plasmid system is a significant improvement over previouslyreported systems that utilize TetR and Lac Repressor instead of rtTA3. In addition to being one of only a few people who were able to use this strategy to label DNA directly in mammalian cells, I have reduced the number of TetO sites required substantially. With this inducible, rtTA3-eGFP-nls-mediated DNA tagging system, the DNA that is to be observed can remain unbound and undisturbed by rtTA3-eGFP-nls until the addition of dox four hours prior to fixing the cells. As a result, negative impact to DNA replication or transcription can be minimized. While researchers can theoretically rinse dox off of TetR to restore its affinity for TetO sites near the endpoint of the experiment, my experience is that this strategy was impractical. The drug associates with TetR and rtTA3 too tightly to 59 be removed. With Lac Repressor, the protein always exhibits a high affinity for LacO sequences. As such, Lac Repressor could not be used for my experimental purposes as the bound Lac Repressor would have prevented the tracking plasmid from replication. I have optimized the rtTA3-eGFP-nls system to tag TetO sites a great deal. To my knowledge, my results show the first time that a TetO-containing DNA sequence has been fluorescently-tagged by rtTA3 following the addition of dox inducer. This is a novel, valuable approach whereby we can now study HPV persistence and DNA segregation. Additionally, I should be able to adapt this model system with relative ease in order to address biological questions in other model systems and organisms. My thesis work with the rtTA3-eGFP-nls and TetO constructs has advanced the ability that researchers have to study discrete DNA elements by fluorescence microscopy a great deal. Possibilities for future research Additional HPV tracking experiments of interest My new HPV ori tracking plasmid system is a powerful tool that will certainly yield more-detailed findings in the future. For instance, co-staining with an antibody to PML, a major component of ND10 bodies, can help localize the replication centers in interphase cells. Previously, our lab has reported that a fraction of the E2 protein colocalizes with ND10 bodies (Swindle et al. 1999). Using GFP-E1, the replication centers were found to be near PML-positive sites (J-H Yu, L.D. Dao, T.R. Broker and L.T. Chow, personal communication). The tracking plasmids will allow us to identify the replication centers relative to host nuclear domains and will yield new insights into how HPV plas- 60 mid DNA is dispersed at and around sites of active replication prior to, after, and during HPV DNA replication. Another line of investigation would be to introduce the tracking plasmid and the prtTA3-GeGFP-NLS into primary human keratinocytes that already harbor the HPV-18 genomic plasmid (Wang et al. 2009). Because of the highly conserved nature of the viral ori sequence, the HPV-18 E1 and E2 should be able to replicate the tracking plasmid which contains the HPV-11 ori. Additionally, the ability to introduce genomic HPV-11 DNA by the same strategy used with HPV-18 is being developed. The advantage of this experiment will be the low copy numbers of the replicating tracking plasmid supported by the low levels of viral replication proteins expressed from the HPV-18 plasmid. This way, the partitioning of the tracking plasmid will reflect the environment in the basal cells more closely than is possible in 293 cells where the tracking plasmid amplifies to extremely high copy numbers. Either in 293 cells or in HPV-18 containing PHKs, staining with additional mitotic marker proteins was performed previously with the Gal4DBD-GFP tracking plasmid system (data not shown). These host marker protein can be used for further characterization of the mitotic localization pattern of the new 11-URR-TetO tracking plasmid system. Markers I can already stain for include Bub1, AuroraB, AuroraA, CenpE, CenpF, Crm1, and multiple microtubule-associated motor proteins. Additionally, the sera from human Crest schleroderma patients (an autoimmune disease) can be used to stain kinetochores. Together, staining with these antibodies will hopefully produce a more-detailed understanding of the dynamic localization of HPV DNA during mitosis. 61 Since we now know that E2 must bind to 11-URR plasmids to allow colocalization of tracking plasmids with the mitotic apparatus, it would be useful to perform mutagenesis and truncation mapping of the URR to determine the precise region(s) and/or elements that are sufficient to facilitate association of HPV plasmids with the mitotic apparatus. I have generated clones incorporating a wide range of URR truncations and/or modifications while I was working with the Gal4DBD-GFP tracking plasmid system. These truncated and/or mutated URR fragments can easily be transferred into my new TetO tracking plasmid system to replace the wild-type 11-URR. These experiments would narrow down the elements within the URR critical for HPV DNA partitioning during mitosis. The use of transiently-transfected plasmid to express rtTA3-eGFP-nls has proved technically-challenging. A significant portion of cells either express too much or too little rtTA3-GFP, and these cells cannot yield useful imaging data. I have generated clonal cell lines that stably-express rtTA3-eGFP-nls at low and moderate levels. It remains to be seen whether these stable 293 cell lines will be more advantageous than transient expression of the rtTA3-eGFP protein. If one of the clonal cell lines expresses rtTA3- eGFP-nls at optimal levels, we might be able to image p11-URR-TetO in almost all cells positive for both the tracking plasmid and the E1E2 operon, and this would represent an enormous improvement over present and past strategies where two and three-plasmid cotransfections are required. An alternative to ChIP to pull down discrete chromatin sequence 62 The most commonly-practiced approach for studying protein-chromatin relies upon pull-downs with antibody specific to a protein of interest. There are inherent problems with this strategy though. First, proteins rarely associate with the locus of interest exclusively. Theoretically, antibody pull-downs will isolate chromatin sequences anywhere in the genome that the pull-down protein is bound. Especially while studying some rare events such as developmental gene switching and differentiation, relying on a native, cellular protein to pull down a specific chromatin fragment of interest can result in authentic background signal. That authentic background can drown out rare events to the point where drawing conclusions is impossible. In this instance, high-resolution chromatin mapping would be more appropriate, but currently, this is difficult and tedious. My work with TetR, rtTA3, and my new TetO-based tracking plasmid system has given me some ideas of how to isolate chromatin at a single locus efficiently and with a low amount background. The presence of other cellular proteins of interest on the pull-down chromatin can then be subsequently analyzed. By knocking TetO arrays in at genetic loci of interest via recombineering (Copeland et al. 2001; Court et al. 2002; Sawitzke et al. 2007; Bamps and Hope 2008), cross-linked, purified chromatin from a specific genomic locus should be able to be studied efficiently. Genomic DNA would be sonicated to yield chromatin fragments of the appropriate average size. Unlike in bacteria, TetR-GFP-nls can be expressed in mammalian cells without toxicity, and the fusion protein could be purified in its active form from crude cellular lysate. Whether it is best to express TetR alone, as a GFP, or as a GFP-nls fusion protein remains to be determined. GFP and the NLS might expendable if a good enough antibody to TetR is available. It might be advantageous to eliminate the NLS al- 63 together to reduce background affinity for the TetO sites. Alternatively, use of antibody raised against GFP might allow TetR more steric flexibility and enhance the ability of TetR to bind TetO sites. This non-enriched, TetR-eGFP-nls-containing lysate could then be added to sonicated chromatin isolated from the cells to be studied. Biotin-conjugated anti-GFP monoclonal antibody could be incubated with the TetR-GFP-nls-exposed chromatin. This mixture could then be incubated with Streptavidin resin to pull down all TetO array-containing chromatin fragments. In theory, the resin should only associate with TetO array-containing chromatin fragments, mediated by biotinylated anti-GFP monoclonal antibody. After appropriate washing to remove unbound chromatin and cellular proteins, the TetO array-containing chromatin fragments could be released from the resin through the addition of doxycyclin (dox). While better than standard ChIP, an ever more favorable strategy is described next. I propose a simpler approach whereby a TetR fusion protein could be expressed, bound to biotinylated microbeads, and used to affinity purify TetO array-containing chromatin fragments directly without any need for antibody. A fungal, avidin-like protein, tamavidin 2 was recently used for affinity purification purposes, but unlike streptavidin, high levels of this avidin homologue can be expressed and purified in soluble form from bacteria (Takakura et al.). Mammalian cell lines such as Cos7’s can express fusion proteins in abundance whenever bacterial expression would be toxic as is the case with TetR. It remains to be determined whether the fusion protein could be expressed and purified using insect cells. To re-state it again, it remains to be determined whether TetR should be expressed with or without being fused to GFP. I propose that TetR-Tamavidin fusion protein could be stably- or transiently-expresed in Cos7 cells. Crude Cos7 lysate 64 with soluble TetR-eGFP-nls-avidin could be incubated with biotinylated microbeads, and unbound cellular material could be washed from the resin. Next, unenriched, sonicated chromatin fragments with and without the TetO array-tagged locus of interest would be incubated with theTetR-eGFP-nls-tamavidin-bound biotinylated microbeads. The microbeads could in theory pull down all TetO array-containing chromatin fragments along with any proteins that were crosslinked to the chromatin. After washing, the only chromatin and protein remaining on the resin should be the TetO-containing chromatin fragments. Addition of dox should allow all chromatin and protein to be released into solution gently, and the TetR-eGFP-nls-tamavidin should all remain in the insoluble, beadcontaining fraction. This strategy should allow for the efficient, relatively-simple purification of specific chromatin fragments and any interacting cellular protein complexes. To my knowledge, this strategy for high-resolution chromatin purification or a similar technique has not yet been reported in the literature. The ability to purify and characterize affinity-enriched, chromatin fragments from genetic loci where a TetO array had been incorporated via homologous recombination would be a significant technological advance. Targeted knock-ins of modified DNA sequences without the need for antibiotic selection has become simpler and more efficient thanks to techniques like recombineering. The same biochemical toolset used to analyze chromatin isolated using standard ChIP protocols could be utilized, including coimmunopreciptations, quantitative PCR, blue-native gel electrophoresis, mass spectrometry, Westerns, and gel-shift assays. Since, this strategy requires no tagging and/or overexpressing of cellular and/or exogenous proteins, all cellular protein interactions and expression profiles can remain undisturbed and completely wild-type. While not replacing 65 the need for standard ChIP protocols, these novel approaches for purifying specific chromatin fragments would reduce background signal and enable researchers to study dynamic chromatin-protein interactions during gene activation, gene repression, and gene silencing. This could produce a better understanding of how expression of individual genes are regulated and enable better dissection of complex processes like cell maturation, differentiation, re-programming. Apart from the chromatin pull-downs, visualization of TetO-tagged loci might be feasible if the cells being studied can tolerate transient transfection and expression of the rtTA3-eGFP-nls plasmid. To have the ability to resolve this, it might be necessary to incorporate a TetO array on each side of the genetic locus to be studied. This sort of experiment could provide a visual readout of whether the chromatin of interest is hetero- or euchromatic. Euchromatin would result in either a single, bright spot or two smaller ones of differing intensities. Heterochromatin would appear as one or two spots smaller than spots produced by euchromatin at the same locus. The total GFP fluorescence of each spot(s) should remain constant, regardless of whether it is hetero- or euchromatin. While ChIP provides an average picture of what is occurring in a population of cells, direct in situ visulazation of DNA by fluorescence microscopy could permit observation of some dynamic processes in single cells. The ability to study specific loci by ChIP and fluorescence microscopy studies could provide new insights including the dynamics and regulation of single genes. 66 Super-resolution fixed- and live-cell imaging on an OMX microscope With more than two-fold greater effective resolution than the standard DeltaVision Spectris, re-imaging the specimen presented earlier in my dissertation on the OMX super-resolution microscope platform (Applied Precision, Inc.) should yield additional details and insight regarding how HPV genomes partition into daughter cells during mitosis. The OMX system employs structured illumination in the X/Y and Z axes to determinee the precise location of fluorophores in three dimensions (Gustafsson et al. 2008; Baddeley et al. 2010; Duleh and Welch 2010; Nakamura et al. 2010). The system is also capable of collecting images more rapidly and with greater sensitivity than is the case with the DeltaVision Spectris. This might make live-cell, time-lapse imaging a possibility with the HPV tracking plasmid system I developed. Ultimately, this work could result in substantial improvement to our present understanding of how HPV DNA partitions into daughter cells, and other more basic biological questions such as those described in the preceding section might also be better-studied by using this more-advanced, higherresolution microscope platform. Materials and Methods Plasmids The HPV-11 expression plasmids peGFP-11-E1E2, peGFP-11E1dm, pMT2-11E2 and the origin of replication-containing plasmid p7730-99 have been described previously (Kuo et al. 1994; Liu et al. 1995; Deng et al. 2003). 67 1. pTetR-eGFP-N1 (lacks SV40 Tag nls; TetR ORF semi-codon-optimized; GFP kozak and AUG deleted): the high-fidelity KOD Hot Start kit was used to generate this clone. Frag-A (770bp) and Frag-B (660bp) PCR products were generated where pCDNA6/TR and peGFP-N1-ΔkozakΔAUG were used as templates, respectively. The 3’ of Frag-A and 5’ of Frag-B include 21nt of overlapping, identical sequence that was used to prime a second round of PCR. The Frag-A and Frag-B PCR products were gel-purified (Qiagen Inc.) and used for a second round of PCR where the sense-strand from Frag-A annealed to the anti-sense strand from Frag-B to yield a 1.43 kbp fragment. The PCR reaction was paused, and the forward primer used to amplify Frag-A and reverse primer used to amplify Frag-B were added to the PCR reactions. The Frag-A forward and reverse primers (IDT Inc.) were ”FragA-FOR-pCDNA6TR-Nhe” (5’-ACC-GTT-GCT-AGC-GTT-TAAACT-TAA-GCT-TGG-TAC-CCG-3’) and “FragA-REV-pCDNA6TR” (5’-TGC-TCACCA-TGG-TGG-CGG-CCT-CGA-GCC-CTA-TAG-TGA-GTC-GTA-TTA-CAA-TTCTTT-GCC-3’), respectively. The Frag-A PCR product included an Nhe restriction site (GCT-AGC), the subsequent UTR-rabbit β-globin intron#2-UTR from the pCDNA6/TR vector, an XhoI cut site (CTC-GAG), a Kozak sequence (GCC-GCC-ACC), and the first ten base-pairs of the TetR open reading frame (ORF) (ATG-GTG-AGC-A). The third codon was changed from TCT to AGC to improve protein expression in human cells. The Frag-B forward and reverse primers were “For-XhoIkozakTetRGFP” (5’-CTCGAG-GCC-GCC-ACC-ATG-GTG-AGC-AGA-CTG-GAC-AAG-AGC-AAG-GTGATT-AAC-AGC-GCC-CTG-GAG-CTG-C-3’) and “FragB-Rev-TetRCTDHindBam” (5’GAC-CGG-TGG-ATC-CAA-GCT-TTC-GGA-CCC-GCT-TTC-ACA-TTT-CAGCTG-TTT-TTC-CAG-TCC-GCA-TAT-GAT), respectively. The portions of the final 68 1.43 kbp PCR product that were contributed by the Frag-B PCR product were the XhoI site (CTC-GAG), kozak sequence (GCC-GCC-ACC), TetR codon-optimization of the third (TCT to AGC), sixth (GAT to GAC), seventh (AAA to AAG), eighth (AGT to AGC), ninth (AAA to AAG), fourteenth (GCA to GCC), and fifteenth (TTA to CTG) TetR codons in the 5’-ORF of TetR. The 3’-portion of the Frag-B PCR product resulted in the codon-optimization of five additional triplets. With the sequence corresponding to the L-E-K-Q-L stretch of amino acids, the Leucine (TTA to CTG), Glutamine (CAA to CAG), and Leucine (CTT to CTG) codons were optimized. The two downstream Serine codons in the TetR ORF were also optimized (AGT to AGC and TCT to TCC, respectively). The TetR ORF was followed immediately by G and A nucleotides, a HindIII site (AAGCTT), a BamHI site (GGATCC), and lastly more of the spacer between the TetR and GFP ORFs (ACC-GGT-C). In all, seven codons of the 5’ TetR ORF were optimized (resulting in the first 17 codons being highly-optimized) and five codons in the 3’ portion of the TetR ORF were optimized (resulting in the final 15 codons in the TetR ORF being highly-optimized). Since I mutated the in the 5’-The gel-purified, full-length 1.43 kbp PCR product was cut with NheI and BamHI. This fragment was ligated into the backbone of peGFP-N1-ΔkozakΔAUG between these same sites to yield a plasmid with a final overall size of 6.1kb. 2. pTetR-eGFP-nls (N1) version (the TetR ORF is semi-codon-optimized; between the TetR and GFP ORFs, the kozak preceding GFP is wild-type; the GFP AUG is wild-type). The NheI/AgeI restriction digest fragment from pTetR-eGFP was purified and ligated into a modified version of the peGFP-N1 backbone where a linker (tcc-gga-ctc-aga-tct- 69 cga-gct) and SV40 T-ag nuclear localization signal (cca-aaa-aag-aag-aga-aag-gta-gtc) were fused to the carboxyl-terminus of GFP at an earlier time. The wild-type GFP kozak sequence and start codon are intact. 3. prtTA3-eGFP-nls (start codon of GFP changed to ATC): the sequence between the two XhoI sites in pTetR-eGFP-nls was excised and removed. The replacement fragment with the rtTA3 ORF and a mutated GFP start codon (ATG to ATC) was commerciallysynthesized (IDT DNA Inc.) based on the rtTA3 ORF (Open Biosystems/Thermo Scientific Inc.). This ensured that rtTA3-eGFP-nls (but not free GFP) protein was solelyexpressed. 4. pTetR-eGFP-nls-C1: The NheI-TetR-eGFP-nls-XbaI fragment was excised from pTetR-eGFP-nls and ligated between the same sites in the peGFP-C1 backbone. This yielded a plasmid with the peGFP-N1 multiple cloning site but the peGFP-C1 plasmid backbone so that I could insert a TetO array into the MluI site at a later time. 5. pTetR-eGFP-nls-TetO: Using TetO clones and a strategy that were previouslydescribed (Lau et al. 2003), I generated and inserted a 48-copy TetO array into pTetReGFP-nls-C1 at the MluI site. 6. p11-URR-TetR-eGFP: The full-length, wild-type HPV-11-URR (nt7070 to 7933/1 to 99) was amplified by PCR from the previously-described p11Rc (Deng et al. 2003) using recombinant Taq polymerase (Invitrogen) except that forward and reverse primers (IDT 70 DNA Inc) incorporated AseI/AclI sites at the 5’ and BssHII/AseI sites were added onto the 3’-end of the URR. The URR was inserted into the AseI site of pTetR-eGFP-nls-C1 in the forward orientation. The complete HPV-11 URR was PCR-amplified from p11Rc to yield a 5’-AseI-AclI-11-URR-BssHII-AseI-3’ product that was ligated into the AseI site of pTetR-eGFP-nls-C1. 7. p11-URR-TetR-eGFP-TetO (lacks an nls): An 5’-AseI-AclI-11-URR-BssHII-AseI-3’ fragment was used to insert the HPV-11-URR into pTetR-eGFP-C1. The 48-copy tetO array was amplified by PCR and inserted into the MluI site of p11-URR-TetR-eGFP. 8. peGFP-N1-ΔkozakΔAUG – PCR mutagenesis was used to delete the kozak sequence and start codon of GFP from the peGFP-N1 vector (Clontech). 9. p11-URR: The region between the BssHII and MluI sites was excised from p11URR-TetR-eGFP-TetO. The sticky ends were filled in and re-ligated to yield this plasmid. 10. pTetO: The URR, CMV promoter, TetR-GFP ORF and the other DNA between the AseI and NotI sites of p11-URR-TetR-eGFP-TetO was excised and discarded. The sticky ends that remained were filled in and re-ligated to yield this plasmid. 71 11. p11-URR-TetO: The region between the BssHII and NotI sites in p11-URR-TetReGFP-TetO was excised. The sticky ends were filled in and re-ligated to yield this plasmid. 12. peBFP2-11-E1E2: The AgeI-eBFP2-BglII fragment from the peBFP2-C1 (unpub- lished) replaced the AgeI-eGFP-BglII fragment I removed from peGFP-11-E1E2. 13. p11-URR-eBFP2-11-E1E2: Using peBFP2-11-E1E2 as a backbone, the 11-URR was inserted at the AseI site as done with other plasmids. 14. peBFP2-11E1E2-TetO: Using peBFP2-11-E1E2 as the backbone, the 48-copy TetO array was inserted at the MluI site. 15. p11-URR-eBFP2-11E1E2-TetO: Using p11-URR-eBFP-11-E1E2 as a backbone, the 48-copy TetO array was amplified by PCR and inserted at the MluI site. 16. pBabePuro-TetR-eGFP-nls: The NheI/XbaI fragment from pTetR-eGFP-nls-N1 was excised and the ends were filled in. This was ligated with the filled-in ends of BamHI/EcoRI-digested pBabePuro. 17. pBabePuro-rtTA3-eGFP-nls: The NheI/XbaI fragment from prtTA3-eGFP-nls was purified and the ends were filled in. This fragment was ligated with the filled-in ends of BamHI/EcoRI-digested pBabePuro. 72 18. TetO array plasmids: pLau02(3 TetO sites), pLau04(6 TetO sites), pLau06(12 TetO sites), and three plasmids with 24 TetO sites each (pLau10-1, pLau10-2, and pLau10-3) were all generously shared with us by David Sherratt through Pat Higgins. Longer TetO arrays were generated using these plasmids and a previously-published strategy (Lau et al. 2003). Cell culture for imaging 293 cells were maintained in phenol red-free Dulbeccos’s modified eagle medium (Mediatech cat#17-205-CV) with 10% fetal bovine serum and 2 mM L-glutamine (Mediatech cat#25-005). Transfections were conducted by electroporation as described previously (Chiang et al. 1992a). For imaging purposes, electroporated cells were seeded onto poly-L-lysine coated coverslips. Round, 40mm, #1 coverslips (Bioptechs Inc. cat#401313-0319-2) were acid-washed, rinsed, and stored in 70% EtOH (http://spectorlab.cshl.edu/acid_clean.html). Flamed coverslips were laid atop parallel lengths of 1 mL plastic pipettes that had been glued to the bottoms of 60mm tissue culture dishes. Poly-L-lysine stock solution (Sigma-Aldrich cat# P8920) was diluted 1:10 with distilled water to yield a 0.01% w/v solution, and approximately 1.3mL of diluted poly-L-lysine was pipetted onto each coverslip and allowed to stand there for 30 minutes. The solution was aspirated off and coverlips were dried in a warm SpeedVac. Dried, coated coverslips were double-UV-sterilized (BioRad GS Gene Linker) and stored until use. Each coverslip was seeded with one-third the cells from an electroporation. As such, each transfection could generate three identical coverslips. 73 Non-clonal pools of TetR-eGFP-nls-293 cells were generated by infecting parental 293 cells with pBabePuro-TetR-eGFP-nls retrovirus produced in AM-12 cells and selected with 1 μg/mL puromycin (Sigma-Aldrich). For clonal rtTA3-eGFP-nls-293 cell lines, pools of non-clonal cells were first generated using pBabePuro-rtTA3-eGFP-nls retrovirus. Clusters of fluorescing rtTA3-eGFP-nls-293 cells were marked individually and passaged using 6- and 8-mm cloning cylinders (Corning Inc cat#3166). Cell sorting was used to sort the different populations of cells and generate a variety of different clonal cell lines. Cell fixation, cell staining, and indirect immunofluorescence 293 cells seeded on coverslips were placed in 60mm, treated tissue culture dishes (Corning Inc.) and cultured for (36 - 60 hrs) at 370C in 5% CO2. Coverslips remained in the same 60mm that they were cultured in until it was time to apply the antibody. The cells were exposed to 2 μg/mL doxycycline for the four hours prior to harvest for imaging as follows. Since the salts in DMEM caused the dox to precipitate out of solution when dox stock solution was added directly to the culture dish and swirling the dishes to redissolve precipitated dox caused many cells to detach, I opted to prepare DMEM with 2μg/mL dox several hours in advance of it being used. To do this, I used 125-mL glass bottles to prepare either 50- or 100-mL batches of induction media. The dox-containing media was stored in 5% CO2 at 37C with the cap loose until the time when I added it to 74 the coverslip-containing dishes Media was vacuum-aspirated using long-tip Pasteur pi- pettes, and 4-mL of dox-containing media was added to the treated cells. Mock-treated cells received 4mL of fresh DMEM that lacked dox. All solutions used during fixation, permeabilization, and the rinses were applied using 60-mL syringes (BD Inc) with 18-Ga needles. Solutions were gently-applied to the edges of the 60mm tissue culture dishes to keep from disturbing or detaching cells excessively. Solutions were removed gently by pouring them off into a waste beaker. To preserve microtubules and spindle fibers, great care was taken to ensure that cells never cooled prior to fixation. For fixation, the DMEM was poured off and cells were fixed directly using freshly-prepared 4% paraformaldehyde in 1x-PBS (pH 7.2) that was maintained at 37C. I observed that fewer cells detached when I added a small amount of PFA to the media prior to pouring it off and putting the cells into pure, 4% PFA solution. After a few seconds and gentle mixing, the dilute DMEM-PFA mixture was poured off and 3-mL of warm, 4% PFA-PBS was added. Coverslip dishes were set atop a small stack of paper towels to allow the fixative to remain warm for longer. After eight minutes, the fixative was poured off and residual PFA was aspirated using a syringe and needle. For five minutes each, three rinses at roomtemperature with 1x-PBS pH7.2 followed. Cells were permeabilized for 10 minutes using a 4oC solution of 1x-PBS/0.5%-Triton-X. As with fixation, I found that fewer cells were lost due to detachment if I first added ~1-mL of the permeabilization buffer to the PBS, prior to pouring the PBS off and adding pure permeabilization solution. After a few seconds of gentle swirling, the dilute PBS-triton-X solution was poured off and replaced with the pure 1x-PBS-0.5% Triton-X permeabilization buffer for ten minutes. Three washes with 1x-PBS, five minutes each, at room-temperature followed. For the third 75 PBS rinse, I used the same basic strategy as I did during fixation and permeabilization. I poured all but about 0.5mL of the second PBS rinse buffer off, taking care to minimize exposure of the cells to air and then added ~4mL of fresh 1x-PBS. A few seconds later, I repeated this procedure and then let the cells sit for five minutes. To block and reduce background staining, 400μL of a 50% goat serum/1x-PBS pH 7.2 blocking solution was added to each dish (final [goat serum]=5% v/v). Immunostaining was performed using 60mm wide slides (Corning) where I had applied histology pen wax to all but a square, ~45mm portion of the slide surface. Coverslips were inverted and set atop a 400μL droplet of the staining antibody which I pre-diluted using 50% goat serum/PBS. Coverslips were removed from their 60mm dish filled with the blocking solution by squeezing the sides of the dish with my left hand. This allowed me to get underneath the coverslip with forceps using my right hand. I only gripped the outermost ~3-4mm of coverslips to minimize destruction of cells. Coverslips were lifted out of the dish gently. I held the coverslips with my left hand by the very edge and blotted excess blocking solution from the side of the coverslip that did not have cells growing on it. I grabbed the same 3-4mm surface of the coverslip with the forceps and gently laid it atop the 400μL bead of diluted antibody on the glass slide that was resting inside a standard slide storage box with wet paper towels lining the bottom. Cells were incubated with antibody overnight. The following morning, staining boxes were removed from the cold box. Again, I used 60-mL syringes to apply all wash solutions. Fresh 60-mm plates with 5mL of 1xPBS/0.1% tween-20 (Sigma-Aldrich) were set up in advance prior to the first wash. To remove the coverslips from the slides where they were being stained, I added ~1-1.5mL uniformly around the edge of each coverslip. This caused the coverslips to float up enough so that I 76 could easily-insert a pair of forceps and grasp ~3-4mmof the surface of the coverslip. The coverslips were gently lifted off of each slide, inverted, and promptly-added to one of the 60mm dishes that already contained the PBS/0.1% tween-20 solution for the first wash. Three 5-minute PBS/tween washes were then followed by three 5-minute 1x-PBS rinses. Care was taken to ensure that the third PBS rinse was done incrementally, as was done following permeabilization. DAPI was added to the third PBS rinse solution for 10 minutes to stain for chromatin. Two PBS rinses followed (each five minutes), and cells were set atop kimwipes (cell-side up) to dry before being mounted with ~40μL of antifade prepared per David Spector’s (Cold Spring Harbor Laboratories) protocol. The recipe is available at: http://spectorlab.cshl.edu/acid_clean.html. In my hands this anti-fade is better and easier to use than any commercially-available anti-fades I have tested. It should also be noted that pure p-phenylenediamine (Sigma-Aldrich) is white. Crystals usually arrive as a tan or brown color. I rinsed large batches of the crystals multiple times in non-pH-adjusted 10x-PBS buffer until all crystals were white and the supernatant was clear. It is my observation that 10x stocks of the pure crystals can be stored stably for several years at -80C. When 1x stock of the crystals is exhausted I simply thaw a 10x stock tube and dilute it 10-fold to refresh our supply of 1x-p-phenylenediamine solutions. 1x and 10x stocks of the crystals are both stored in non-pH-adjusted 10xPBS at 80C. Using David Spector’s anti-fade recipe, 1.5mL tubes of colorless anti-fade could be stored under nitrogen at -20C inside a tightly-capped bottle for several months without becoming noticeably-discolored. Discolored anti-fade (more than a pale yellow or orange color) will result in autofluorescent staining of chromatin. 77 For all antibody staining I used the blocking, staining, and wash conditions described in the previous paragraph, and antibody dilutions were as follows. For staining microtubules, Cy3-conjugated, anti-β-tubulin mouse monoclonal antibody (Sigma) was used at a 1:50 dilution. Chromatin was stained using DAPI. Slides with coverslips of cells were stored at -20oC until it was time to image. Though images were not presented in the body of my dissertation, I have successfully stained 293 cells using anti-CREST serum (Cortex Biochem) at 1:800, AIM-1 antibody (BD Transduction Labs.) at 1:100, aurora-A Kinase antibody (BD Transduction Labs) at 1:200, CENP-F (anti-Mitosin) antibody (BD Transduction Labs) at 1:100, CENP-E (abCam) at 1:250, and Crm1 antibody (Santacruz) at 1:100 dilutions with very low background noise. Transient replication assays For each electroporation, 5x106 293 cells re-suspended in 250μL DMEM that contained BES and 50μg of boiled, snap-cooled salmon-sperm DNA (Invitrogen). Wherever required the following amounts of plasmid were transfected: 500ng of p7730-99 HPV-11 ori, 1μg peGFP-11-E1dm, 5μg pMT2-11-E1, 5μg pMT2-11-E2, 5μg p11-URR, 5μg pTetO, 5μg p11-URR-TetO, 5μg peBFP2-11-E1E2, 6μg pTetR-eGFP-nls, and 6μg prtTA3-eGFP-nls plasmid were utilized. Cells were grown either in the presence or absence of 2 μg/mL doxycycline (Sigma-Aldrich cat#D9891) for the entire duration of the experiment. The harvest of the low molecular weight DNA and Southern blot hybrization were performed as previously described (Chiang et al. 1992a) DpnI was acquired from Promega. Apa I and HindIII were acquired from New England Biolabs Inc. 78 Derivation of 293 cell lines expressing TetR-eGPF or rtTA3-eGFP pBabePuro-TetR-eGFP-nls and pBabePuro-intron-rtTA3-eGFP-nls were used to generate retroviruses. The plasmids were electroporated into Bosc23 cells. The following day, sterile-filtered supernatant containing polybrene (Sigma) at a final concentration of 5.2 μg/mL was dripped over Am12 cells (at 50-60% confluency). Media was changed after 6 hours. 24 hours later, infected Am12 cells began 48-hour selection period utilizing 1 μg/mL puromycin. Upon removal of selection, Am12 cells were grown to nearconfluency and passaged. rtTA3-eGFP-nls and TetR-eGFP-nls virus-containing supernatant from the respective Am12 cells was sterile-filtered, mixed with polybrene, and dripped onto 10 cm plates of 293 cells that had been seeded with either 1- or 1.2E6 293 HEK cells the prior day. 36-48 hours later, one-week selection with 1 or 1.25 μg/mL puromycin was carried out. Media and antibiotic were changed every two days. Pools of non-clonal 293 cells infected with the TetR-eGFP-nls retrovirus did not continue expressing the GFP fusion protein. As a result of this observation I utilized cloning cylinders and FACS sorting to produce clonal 293 cell lines that stably-express rtTA3-eGFP-nls at low to moderate levels. 79 References: Abbate, E.A., Berger, J.M., and Botchan, M.R. 2004. The X-ray structure of the papillomavirus helicase in complex with its molecular matchmaker E2. Genes Dev 18(16): 1981-1996. Ai, H.-w., Shaner, N.C., Cheng, Z., Tsien, R.Y., and Campbell, R.E. 2007. Exploration of New Chromophore Structures Leads to the Identification of Improved Blue Fluorescent Proteins. Biochemistry 46(20): 5904-5910. Amin, A.A., Titolo, S., Pelletier, A., Fink, D., Cordingley, M.G., and Archambault, J. 2000. Identification of Domains of the HPV11 E1 Protein Required for DNA Replication in Vitro. Virology 272(1): 137-150. Androphy, E.J., Lowy, D.R., and Schiller, J.T. 1987. Bovine papillomavirus E2 transactivating gene product binds to specific sites in papillomavirus DNA. Nature 325(6099): 70-73. Aoki, Y. and Tosato, G. 2004. Neoplastic Conditions in the Context of HIV-1 Infection. Current HIV Research 2(4): 343-349. Baddeley, D., Chagin, V.O., Schermelleh, L., Martin, S., Pombo, A., Carlton, P.M., Gahl, A., Domaing, P., Birk, U., Leonhardt, H., Cremer, C., and Cardoso, M.C. 2010. Measurement of replication structures at the nanometer scale using superresolution light microscopy. Nucl Acids Res 38(2): e8-. Bamps, S. and Hope, I.A. 2008. Large-scale gene expression pattern analysis, in situ, in Caenorhabditis elegans. Briefings in Functional Genomics and Proteomics 7(3): 175-183. Banerjee, N.S., Genovese, N.J., Noya, F., Chien, W.-M., Broker, T.R., and Chow, L.T. 2006. Conditionally Activated E7 Proteins of High-Risk and Low-Risk Human Papillomaviruses Induce S Phase in Postmitotic, Differentiated Human Keratinocytes. J Virol 80(13): 6517-6524. Baxter, M.K., McPhillips, M.G., Ozato, K., and McBride, A.A. 2005. The Mitotic Chromosome Binding Activity of the Papillomavirus E2 Protein Correlates with Interaction with the Cellular Chromosomal Protein, Brd4. J Virol 79(8): 4806-4818. Bechtold, V., Beard, P., and Raj, K. 2003. Human Papillomavirus Type 16 E2 Protein Has No Effect on Transcription from Episomal Viral DNA. J Virol 77(3): 20212028. Bedell, M., Jones, K., Grossman, S., and Laimins, L. 1989. Identification of human papillomavirus type 18 transforming genes in immortalized and primary cells. Journal of Virology 63(3): 1247-1255. Bernard, H.-U. 2002. Gene expression of genital human papillomaviruses and considerations on potential antiviral approaches. Antivir Ther 7(4): 219-237. 80 Blitz, I. and Laimins, L. 1991. The 68-kilodalton E1 protein of bovine papillomavirus is a DNA binding phosphoprotein which associates with the E2 transcriptional activator in vitro. Journal of Virology 65(2): 649-656. Brannon, A.R., Maresca, J.A., Boeke, J.D., Basrai, M.A., and McBride, A.A. 2005. Reconstitution of papillomavirus E2-mediated plasmid maintenance in Saccharomyces cerevisiae by the Brd4 bromodomain protein. PNAS 102(8): 2998-3003. Bromberg-White, J.L. and Meyers, C. 2003. Comparison of the basal and glucocorticoidinducible activities of the upstream regulatory regions of HPV18 and HPV31 in multiple epithelial cell lines. Virology 306(2): 197-202. Brown, K., Wood, K., and Cleveland, D. 1996. The kinesin-like protein CENP-E is kinetochore-associated throughout poleward chromosome segregation during anaphase-A. J Cell Sci 109(5): 961-969. Burke, D.J. and Stukenberg, P.T. 2008. Linking Kinetochore-Microtubule Binding to the Spindle Checkpoint. Developmental Cell 14(4): 474-479. Cardenas-Mora, J., Spindler, J.E., Jang, M.K., and McBride, A.A. 2008. Dimerization of the Papillomavirus E2 Protein Is Required for Efficient Mitotic Chromosome Association and Brd4 Binding. J Virol 82(15): 7298-7305. Cheng, S., Schmidt-Grimminger, D., Murant, T., Broker, T., and Chow, L. 1995. Differentiation-dependent up-regulation of the human papillomavirus E7 gene reactivates cellular DNA replication in suprabasal differentiated keratinocytes. Genes Dev 9(19): 2335-2349. Chiang, C., Broker, T.R., and Chow, L.T. 1991. An E1M--E2C fusion protein encoded by human papillomavirus type 11 is a sequence-specific transcription repressor. J Virol 65(6): 3317-3329. Chiang, C., Ustav, M., Stenlund, A., Ho, T., Broker, T., and Chow, L. 1992a. Viral E1 and E2 Proteins Support Replication of Homologous and Heterologous Papillomaviral Origins. PNAS 89(13): 5799-5803. Chiang, C.-M., Broker, T.R., and Chow, L.T. 1992b. Properties of bovine papillomavirus E1 mutants. Virology 191(2): 964-967. Chiang, C.-M., Dong, G., Broker, T.R., and Chow, L.T. 1992c. Control of human papillomavirus type 11 origin of replication by the E2 family of transcription regulatory proteins. J Virol 66(9): 5224-5231. Chin, M., Broker, T.R., and Chow, L.T. 1989. Identification of a novel constitutive enhancer element and an associated binding protein: implications for human papillomavirus type 11 enhancer regulation. J Virol 63(7): 2967-2976. Chin, M., Hirochika, R., Hirochika, H., Broker, T.R., and Chow, L.T. 1988. Regulation of human papillomavirus type 11 enhancer and E6 promoter by activating and repressing proteins from the E2 open reading frame: functional and biochemical studies. J Virol 62(8): 2994-3002. Chow, L.T. and Broker, T.R. 1997. In vitro experimental systems for HPV: Epithelial raft cultures for investigations of viral reproduction and pathogenesis and for genetic analyses of viral proteins and regulatory sequences. Clin Derm 15(2): 217-227. Chow, L.T. and Broker, T.R. 2006. Mechanisms and Regulation of Papillomavirus DNA Replication. in Papillomavirus research: from natural history to vaccines and beyond (ed. M.S. Campo), pp. 53-72. Caister Academic Press, Wynmondham, England. 81 Chow, L.T., Duffy, A.A., Wang, H.-K., and Broker, T.R. 2009. A highly efficient system to produce infectious human papillomavirus: Elucidation of natural virus-host interactions. Cell Cycle 8(9): 1319-1323. Conger, K.L., Liu, J.-S., Kuo, S.-R., Chow, L.T., and Wang, T.S.-F. 1999. Human Papillomavirus DNA Replication. Interactions Between the Viral E1 Protein and Two Subunits of Human DNA Polymerase alpha/Primase. J Biol Chem 274(5): 26962705. Conrad, M., Bubb, V., and Schlegel, R. 1993. The human papillomavirus type 6 and 16 E5 proteins are membrane- associated proteins which associate with the 16kilodalton pore-forming protein. J Virol 67(10): 6170-6178. Copeland, N.G., Jenkins, N.A., and Court, D.L. 2001. Recombineering: a powerful new tool for mouse functional genomics. Nat Rev Genet 2(10): 769-779. Court, D.L., Sawitzke, J.A., and Thomason, L.C. 2002. Genetic engineering using homoogous recombination. Annual Review of Genetics 36(1): 361-388. Crusius, K., Kaszkin, M., Kinzel, V., and Alonso, A. 1999. The human papillomavirus type 16 E5 protein modulates phospholipase C--1 activity and phosphatidyl inositol turnover in mouse fibroblasts. Oncogene 18(48): 6714-6718. Crusius, K., Rodriguez, I., and Alonso, A. 2000. The Human Papillomavirus Type 16 E5 Protein Modulates ERK1/2 and p38 MAP Kinase Activation by an EGFRIndependent Process in Stressed Human Keratinocytes. Virus Genes 20(1): 69. Dao, L.D., Duffy, A., Van Tine, B.A., Wu, S.-Y., Chiang, C.-M., Broker, T.R., and Chow, L.T. 2006. Dynamic Localization of the Human Papillomavirus Type 11 Origin Binding Protein E2 through Mitosis While in Association with the Spindle Apparatus. J Virol 80(10): 4792-4800. Das, A.T., Zhou, X., Vink, M., Klaver, B., Verhoef, K., Marzio, G., and Berkhout, B. 2004. Viral Evolution as a Tool to Improve the Tetracycline-regulated Gene Expression System. Journal of Biological Chemistry 279(18): 18776-18782. Davy, C.E., Jackson, D.J., Wang, Q., Raj, K., Masterson, P.J., Fenner, N.F., Southern, S., Cuthill, S., Millar, J.B.A., and Doorbar, J. 2002. Identification of a G2 Arrest Domain in the E1{wedge}E4 Protein of Human Papillomavirus Type 16. J Virol 76(19): 9806-9818. de Villiers, E.-M., Fauquet, C., Broker, T.R., Bernard, H.-U., and zur Hausen, H. 2004. Classification of papillomaviruses. Virology 324(1): 17-27. Decordier, I., Cundari, E., and Kirsch-Volders, M. 2008. Mitotic checkpoints and the maintenance of the chromosome karyotype. Mutation Research/Genetic Toxicology and Environmental Mutagenesis 651(1-2): 3-13. DeFilippis, R.A., Goodwin, E.C., Wu, L., and DiMaio, D. 2003. Endogenous Human Papillomavirus E6 and E7 Proteins Differentially Regulate Proliferation, Senescence, and Apoptosis in HeLa Cervical Carcinoma Cells. J Virol 77(2): 15511563. Deng, W., Jin, G., Lin, B.-Y., Van Tine, B.A., Broker, T.R., and Chow, L.T. 2003. mRNA Splicing Regulates Human Papillomavirus Type 11 E1 Protein Production and DNA Replication. J Virol 77(19): 10213-10226. Disbrow, G.L., Sunitha, I., Baker, C.C., Hanover, J., and Schlegel, R. 2003. Codon optimization of the HPV-16 E5 gene enhances protein expression. Virology 311(1): 105-114. 82 Dollard, S., Wilson, J., Demeter, L., Bonnez, W., Reichman, R., Broker, T., and Chow, L. 1992. Production of human papillomavirus and modulation of the infectious program in epithelial raft cultures. OFF. Genes Dev 6(7): 1131-1142. Dong, G., Broker, T.R., and Chow, L.T. 1994. Human papillomavirus type 11 E2 proteins repress the homologous E6 promoter by interfering with the binding of host transcription factors to adjacent elements. J Virol 68(2): 1115-1127. Doorbar, J. 2005. The papillomavirus life cycle. Journal of Clinical Virology 32(Supplement 1): 7-15. Doorbar, J., Ely, S., Sterling, J., McClean, C., and Crawford, L. 1991. Specific interactions between HPV-16 E1-E4 and cytokeratins results in collapse of the epithelial cell intermediate filament network. Nature 352: 824-827. Dowhanick, J., McBride, A., and Howley, P. 1995. Suppression of cellular proliferation by the papillomavirus E2 protein. J Virol 69(12): 7791-7799. Duleh, S.N. and Welch, M.D. 2010. WASH and the Arp2/3 complex regulate endosome shape and trafficking. Cytoskeleton 67(3): 193-206. Ebersbach, G., Sherratt, D.J., and Gerdes, K. 2005. Partition-associated incompatibility caused by random assortment of pure plasmid clusters. Molecular Microbiology 56(6): 1430-1440. Ellenson, L.H. and Wu, T.C. 2004. Focus on endometrial and cervical cancer. Cancer Cell 5(6): 538. Fiebig, A., Keren, K., and Theriot, J.A. 2006. Fine-scale time-lapse analysis of the biphasic, dynamic behaviour of the two Vibrio cholerae chromosomes. Molecular Microbiology 60(5): 1164-1178. Fouts, E.T., Yu, X., Egelman, E.H., and Botchan, M.R. 1999. Biochemical and Electron Microscopic Image Analysis of the Hexameric E1 Helicase. J Biol Chem 274(7): 4447-4458. Francis, D.A., Schmid, S.I., and Howley, P.M. 2000. Repression of the Integrated Papillomavirus E6/E7 Promoter Is Required for Growth Suppression of Cervical Cancer Cells. J Virol 74(6): 2679-2686. Gage, J., Meyers, C., and Wettstein, F. 1990. The E7 proteins of the nononcogenic human papillomavirus type 6b (HPV-6b) and of the oncogenic HPV-16 differ in retinoblastoma protein binding and other properties. Journal of Virology 64(2): 723-730. Genovese, N.J., Banerjee, N.S., Broker, T.R., and Chow, L.T. 2008. Casein Kinase II Motif-Dependent Phosphorylation of Human Papillomavirus E7 Protein Promotes p130 Degradation and S-Phase Induction in Differentiated Human Keratinocytes. J Virol 82(10): 4862-4873. Gewin, L., Myers, H., Kiyono, T., and Galloway, D.A. 2004. Identification of a novel telomerase repressor that interacts with the human papillomavirus type-16 E6/E6AP complex. Genes Dev 18(18): 2269-2282. Giri, I. and Yaniv, M. 1988. Structural and mutational analysis of E2 trans-activating proteins of papillomaviruses reveals three distinct functional domains. EMBO J 7(9): 2823-2829. Gius, D., Grossman, S., Bedell, M., and Laimins, L. 1988. Inducible and constitutive enhancer domains in the noncoding region of human papillomavirus type 18. Journal of Virology 62(3): 665-672. 83 Gossen, M. and Bujard, H. 1992. Tight control of gene expression in mammalian cells by tetracycline-responsive promoters. PNAS 89(12): 5547-5551. Grm, H.S. and Banks, L. 2004. Degradation of hDlg and MAGIs by human papillomavirus E6 is E6-AP-independent. J Gen Virol 85(10): 2815-2819. Gustafsson, M.G.L., Shao, L., Carlton, P.M., Wang, C.J.R., Golubovskaya, I.N., Cande, W.Z., Agard, D.A., and Sedat, J.W. 2008. Three-Dimensional Resolution Doubling in Wide-Field Fluorescence Microscopy by Structured Illumination. Biophysical Journal 94(12): 4957-4970. Halbert, C.L., Demers, G.W., and Galloway, D.A. 1992. The E6 and E7 genes of human papillomavirus type 6 have weak immortalizing activity in human epithelial cells. J Virol 66(4): 2125-2134. Ham, J., Dostatni, N., Gauthier, J.-M., and Yaniv, M. 1991. The papillomavirus E2 protein: a factor with many talents. Trends in Biochemical Sciences 16: 440-444. Han, Y., Loo, Y.-M., Militello, K.T., and Melendy, T. 1999. Interactions of the Papovavirus DNA Replication Initiator Proteins, Bovine Papillomavirus Type 1 E1 and Simian Virus 40 Large T Antigen, with Human Replication Protein A. J Virol 73(6): 4899-4907. Handa, K., Yugawa, T., Narisawa-Saito, M., Ohno, S.-i., Fujita, M., and Kiyono, T. 2007. E6AP-Dependent Degradation of DLG4/PSD95 by High-Risk Human Papillomavirus Type 18 E6 Protein. J Virol 81(3): 1379-1389. Hengstermann, A., D'silva, M.A., Kuballa, P., Butz, K., Hoppe-Seyler, F., and Scheffner, M. 2005. Growth Suppression Induced by Downregulation of E6-AP Expression in Human Papillomavirus-Positive Cancer Cell Lines Depends on p53. J Virol 79(14): 9296-9300. Higuchi, T. and Uhlmann, F. 2005. Stabilization of microtubule dynamics at anaphase onset promotes chromosome segregation. Nature 433(7022): 171-176. Hirochika, H., Broker, T.R., and Chow, L.T. 1987. Enhancers and trans-acting E2 transcriptional factors of papillomaviruses. J Virol 61(8): 2599-2606. Hirochika, H., Hirochika, R., Broker, T., and Chow, L. 1988. Functional mapping of the human papillomavirus type 11 transcriptional enhancer and its interaction with the trans-acting E2 proteins [published erratum appears in Genes Dev 1988 Apr;2(4):490]. Genes Dev 2(1): 54-67. Horner, S.M., DeFilippis, R.A., Manuelidis, L., and DiMaio, D. 2004. Repression of the Human Papillomavirus E6 Gene Initiates p53-Dependent, TelomeraseIndependent Senescence and Apoptosis in HeLa Cervical Carcinoma Cells. J Virol 78(8): 4063-4073. Hou, S.Y., Wu, S.-Y., Zhou, T., Thomas, M.C., and Chiang, C.-M. 2000. Alleviation of Human Papillomavirus E2-Mediated Transcriptional Repression via Formation of a TATA Binding Protein (or TFIID)-TFIIB-RNA Polymerase II-TFIIF Preinitiation Complex. Mol Cell Biol 20(1): 113-125. Howley, P.M. 1996. Papillomaviridae: the viruses and their replication. in Fields Virology (ed. B.N. Fields, D.M. Knipe, and P.M. Howley), pp. 2045-2076. LippincottRaven, Philadelphia. Hummel, M., Hudson, J., and LAimins, L. 1992. Differentiation-induced and constitutive transcription of human papillomavirus type 31b in cell lines containing viral episomes. Journal of Virology 66(10): 6070-6080. 84 Janicki, S.M., Tsukamoto, T., Salghetti, S.E., Tansey, W.P., Sachidanandam, R., Prasanth, K.V., Ried, T., Shav-Tal, Y., Bertrand, E., Singer, R.H., and Spector, D.L. 2004. From Silencing to Gene Expression: Real-Time Analysis in Single Cells. Cell 116(5): 683-698. Jeong Seo, E., Jung Kim, H., Jae Lee, C., Tae Kang, H., and Seong Hwang, E. 2004. The role of HPV oncoproteins and cellular factors in maintenance of hTERT expression in cervical carcinoma cells. Gynecologic Oncology 94(1): 40-47. Jha, S., Vande Pol, S., Banerjee, N.S., Dutta, A.B., Chow, L.T., and Dutta, A. 2010. Destabilization of TIP60 by Human Papillomavirus E6 Results in Attenuation of TIP60-Dependent Transcriptional Regulation and Apoptotic Pathway. Molecular Cell 38(5): 700-711. Kastan, M., Canman, C., and Leonard, C. 1995. P53, cell cycle control and apoptosis: implications for cancer. Cancer and Metastasis Reviews 14: 3-15. Kirnbauer, R., Booy, F., Cheng, N., Lowy, D., and Schiller, J. 1992. Papillomavirus L1 Major Capsid Protein Self-Assembles into Virus-Like Particles that are Highly Immunogenic. PNAS 89(24): 12180-12184. Kiyono, T., Foster, S.A., Koop, J.I., McDougall, J.K., Galloway, D.A., and Klingelhutz, A.J. 1998. Both Rb/p16INK4a inactivation and telomerase activity are required to immortalize human epithelial cells. Nature 396(6706): 84-88. Knight, G.L., Grainger, J.R., Gallimore, P.H., and Roberts, S. 2004. Cooperation between Different Forms of the Human Papillomavirus Type 1 E4 Protein To Block Cell Cycle Progression and Cellular DNA Synthesis. J Virol 78(24): 13920-13933. Knight, G.L., Turnell, A.S., and Roberts, S. 2006. Role for Wee1 in Inhibition of G2-toM Transition through the Cooperation of Distinct Human Papillomavirus Type 1 E4 Proteins. J Virol 80(15): 7416-7426. Krawczyk, E., Suprynowicz, F.A., Liu, X., Dai, Y., Hartmann, D.P., Hanover, J., and Schlegel, R. 2008. Koilocytosis: A Cooperative Interaction between the Human Papillomavirus E5 and E6 Oncoproteins. Am J Pathol 173(3): 682-688. Kuo, S., Liu, J., Broker, T., and Chow, L. 1994. Cell-free replication of the human papillomavirus DNA with homologous viral E1 and E2 proteins and human cell extracts. J Biol Chem 269(39): 24058-24065. Lau, I.F., Filipe, S.R., Søballe, B., Økstad, O.-A., Barre, F.-X., and Sherratt, D.J. 2003. Spatial and temporal organization of replicating <i>Escherichia coli</i> chromosomes. Molecular Microbiology 49(3): 731-743. Lechner, M. and Laimins, L. 1994. Inhibition of p53 DNA binding by human papillomavirus E6 proteins. Journal of Virology 68(7): 4262-4273. Lee, C.J., Suh, E.J., Kang, H.T., Im, J.S., Um, S.J., Park, J.S., and Hwang, E.S. 2002. Induction of Senescence-like State and Suppression of Telomerase Activity through Inhibition of HPV E6/E7 Gene Expression in Cells Immortalized by HPV16 DNA. Experimental Cell Research 277(2): 173-182. Leykauf, K., Salek, M., Schluter, H., Lehmann, W.-D., and Alonso, A. 2004. Identification of membrane proteins differentially expressed in human papillomavirus type 16 E5-transfected human keratinocytes by nanoelectrospray ionization mass spectrometry. J Gen Virol 85(6): 1427-1431. 85 Lin, B.Y., Makhov, A.M., Griffith, J.D., Broker, T.R., and Chow, L.T. 2002. Chaperone Proteins Abrogate Inhibition of the Human Papillomavirus (HPV) E1 Replicative Helicase by the HPV E2 Protein. Mol Cell Biol 22(18): 6592-6604. Liu, J.-S., Kuo, S.-R., Broker, T.R., and Chow, L.T. 1995. The Functions of Human Papillomavirus Type 11 E1, E2, and E2C Proteins in Cell-free DNA Replication. J Biol Chem 270(45): 27283-27291. Liu, J.-S., Kuo, S.-R., Makhov, A.M., Cyr, D.M., Griffith, J.D., Broker, T.R., and Chow, L.T. 1998. Human Hsp70 and Hsp40 Chaperone Proteins Facilitate Human Papillomavirus-11 E1 Protein Binding to the Origin and Stimulate Cell-free DNA Replication. J Biol Chem 273(46): 30704-30712. Longworth, M.S. and Laimins, L.A. 2004. Pathogenesis of Human Papillomaviruses in Differentiating Epithelia. Microbiol Mol Biol Rev 68(2): 362-372. Lusky, M., Hurwitz, J., and Seo, Y. 1993. Cooperative assembly of the bovine papilloma virus E1 and E2 proteins on the replication origin requires an intact E2 binding site. J Biol Chem 268(21): 15795-15803. MacPherson, P., Thorner, L., Parker, L.M., and Botchan, M. 1994. The Bovine Papilloma Virus E1 Protein Has ATPase Activity Essential to Viral DNA Replication and Efficient Transformation in Cells. Virology 204(1): 408. Mao, Y., Abrieu, A., and Cleveland, D.W. 2003. Activating and Silencing the Mitotic Checkpoint through CENP-E-Dependent Activation/Inactivation of BubR1. Cell 114(1): 98. Mao, Y., Desai, A., and Cleveland, D.W. 2005. Microtubule capture by CENP-E silences BubR1-dependent mitotic checkpoint signaling. J Cell Biol 170(6): 873-880. Masterson, P.J., Stanley, M.A., Lewis, A.P., and Romanos, M.A. 1998. A C-Terminal Helicase Domain of the Human Papillomavirus E1 Protein Binds E2 and the DNA Polymerase alpha -Primase p68 Subunit. J Virol 72(9): 7407-7419. Matsumoto, Y., Nakagawa, S., Yano, T., Takizawa, S., Nagasaka, K., Nakagawa, K., Minaguchi, T., Wada, O., Ooishi, H., Matsumoto, K., Yasugi, T., Kanda, T., Huibregtse, J.M., and Taketani, Y. 2006. Involvement of a cellular ubiquitin-protein ligase E6AP in the ubiquitin-mediated degradation of extensive substrates of high-risk human papillomavirus E6. Journal of Medical Virology 78(4): 501-507. McBride, A.A., McPhillips, M.G., and Oliveira, J.G. 2004. Brd4: tethering, segregation and beyond. Trends in Microbiology 12(12): 527-529. McLaughlin-Drubin, M.E., Christensen, N.D., and Meyers, C. 2004. Propagation, infection, and neutralization of authentic HPV16 virus. Virology 322(2): 213-219. McLaughlin-Drubin, M.E. and Meyers, C. 2004. Evidence for the coexistence of two genital HPV types within the same host cell in vitro. Virology 321(2): 173-180. McPhillips, M.G., Oliveira, J.G., Spindler, J.E., Mitra, R., and McBride, A.A. 2006. Brd4 Is Required for E2-Mediated Transcriptional Activation but Not Genome Partitioning of All Papillomaviruses. J Virol 80(19): 9530-9543. McPhillips, M.G., Ozato, K., and McBride, A.A. 2005. Interaction of Bovine Papillomavirus E2 Protein with Brd4 Stabilizes Its Association with Chromatin. J Virol 79(14): 8920-8932. Michaelis, C., Ciosk, R., and Nasmyth, K. 1997. Cohesins: Chromosomal Proteins that Prevent Premature Separation of Sister Chromatids. Cell 91(1): 35-45. 86 Miller, J.H. and Reznikoff, W.A. 1980. The Operon. Cold Spring Harbor Laboratory Press, Woodbury, NY. Morrison, E. 2007. Action and interactions at microtubule ends. Cellular and Molecular Life Sciences (CMLS) 64(3): 307-317. Münger, K. and Howley, P.M. 2002. Human papillomavirus immortalization and transformation functions. Virus Research 89(2): 213-228. Musacchio, A. and Salmon, E.D. 2007. The spindle-assembly checkpoint in space and time. Nat Rev Mol Cell Biol 8(5): 379-393. Nakagawa, S. and Huibregtse, J.M. 2000. Human Scribble (Vartul) Is Targeted for Ubiquitin-Mediated Degradation by the High-Risk Papillomavirus E6 Proteins and the E6AP Ubiquitin-Protein Ligase. Mol Cell Biol 20(21): 8244-8253. Nakamura, A., VA Rao, V., Pommier, Y., and Bonner, W. 2010. The complexity of phosphorylated H2AX foci formation and DNA repair assembly at DNA doublestrand breaks. Cell Cycle 9(2): 389-397. Nasseri, M., Hirochika, R., Broker, T.R., and Chow, L.T. 1987. A human papilloma virus type 11 transcript encoding an E1E4 protein. Virology 159(2): 433-439. Nishimura, A., Ono, T., Ishimoto, A., Dowhanick, J., Frizell, M.A., Howley, P.M., and Sakai, H. 2000. Mechanisms of Human Papillomavirus E2-Mediated Repression of Viral Oncogene Expression and Cervical Cancer Cell Growth Inhibition. J Virol 74(8): 3752-3760. Oh, S.T., Kyo, S., and Laimins, L.A. 2001. Telomerase Activation by Human Papillomavirus Type 16 E6 Protein: Induction of Human Telomerase Reverse Transcriptase Expression through Myc and GC-Rich Sp1 Binding Sites. J Virol 75(12): 5559-5566. Oliveira, J.G., Colf, L.A., and McBride, A.A. 2006. Variations in the association of papillomavirus E2 proteins with mitotic chromosomes. PNAS: 0507624103. Ono, T., Fang, Y., Spector, D.L., and Hirano, T. 2004. Spatial and Temporal Regulation of Condensins I and II in Mitotic Chromosome Assembly in Human Cells. Mol Biol Cell: E04-03-0242. Ozbun, M.A. and Meyers, C. 1998. Temporal Usage of Multiple Promoters during the Life Cycle of Human Papillomavirus Type 31b. J Virol 72(4): 2715-2722. Parish, J.L., Bean, A.M., Park, R.B., and Androphy, E.J. 2006. ChlR1 Is Required for Loading Papillomavirus E2 onto Mitotic Chromosomes and Viral Genome Maintenance. Molecular Cell 24(6): 867-876. Park, P., Copeland, W., Yang, L., Wang, T., Botchan, M., and Mohr, I. 1994. The Cellular DNA Polymerase {alpha}-Primase is Required for Papillomavirus DNA Replication and Associates with the Viral E1 Helicase. PNAS 91(18): 8700-8704. Parker, J., Zhao, W., Askins, K., Broker, T., and Chow, L. 1997. Mutational analyses of differentiation-dependent human papillomavirus type 18 enhancer elements in epithelial raft cultures of neonatal foreskin keratinocytes. Cell Growth Differ 8(7): 751-762. Pim, D., Thomas, M., Javier, R., Gardiol, D., and Banks, L. 2000. HPV E6 targeted degradation of the discs large protein: evidence for the involvement of a novel ubiquitin ligase. Oncogene 19(6): 719-725. 87 Plug-DeMaggio, A.W., Sundsvold, T., Wurscher, M.A., Koop, J.I., Klingelhutz, A.J., and McDougall, J.K. 2004. Telomere erosion and chromosomal instability in cells expressing the HPV oncogene 16E6. Oncogene 23(20): 3561-3571. Pray, T.R. and Laimins, L.A. 1995. Differentiation-dependent expression of E1?E4 proteins in cell lines maintaining episomes of human papillomavirus type 31b. Virology 206(1): 685. Raj, K., Berguerand, S., Southern, S., Doorbar, J., and Beard, P. 2004. E1{wedge}E4 Protein of Human Papillomavirus Type 16 Associates with Mitochondria. J Virol 78(13): 7199-7207. Raj, K. and Stanley, M. 1995. The ATP-binding and ATPase activities of human papillomavirus type 16 E1 are significantly weakened by the absence of prolines in its ATP- binding domain. J Gen Virol 76(12): 2949-2956. Robinett, C., Straight, A., Li, G., Willhelm, C., Sudlow, G., Murray, A., and Belmont, A. 1996. In vivo localization of DNA sequences and visualization of large-scale chromatin organization using lac operator/repressor recognition. J Cell Biol 135(6): 1685-1700. Sanders, C.M. and Stenlund, A. 1998. Recruitment and loading of the E1 initiator protein: an ATP-dependent process catalysed by a transcription factor. EMBO J 17(23): 7044-7055. Sawitzke, J.A., Thomason, L.C., Costantino, N., Bubunenko, M., Datta, S., Court, D.L., Kelly, T.H., and Stanley, R.M. 2007. Recombineering: In Vivo Genetic Engineering in E. coli, S. enterica, and Beyond. in Methods in Enzymology, pp. 171-199. Academic Press. Scheffner, M., Huibregtse, J.M., Vierstra, R.D., and Howley, P.M. 1993. The HPV-16 E6 and E6-AP complex functions as a ubiquitin-protein ligase in the ubiquitination of p53. Cell 75(3): 495-505. Schiller, J.T. and Davies, P. 2004. Delivering on the Promise: HPV Vaccines and Cervical Cancer. Nat Rev Micro 2(4): 343-347. Schneider-Gadicke, A. and Schwarz, E. 1986. Different human cervical carcinoma cell lines show similar transcription patterns of human papillomavirus type 18 early genes. EMBO J 5(9): 2285-2292. Sdek, P., Zhang, Z.Y., Cao, J., Pan, H.Y., Chen, W.T., and Zheng, J.W. 2006. Alteration of cell-cycle regulatory proteins in human oral epithelial cells immortalized by HPV16 E6 and E7. International Journal of Oral and Maxillofacial Surgery 35(7): 653-657. Sedman, J. and Stenlund, A. 1998. The Papillomavirus E1 Protein Forms a DNADependent Hexameric Complex with ATPase and DNA Helicase Activities. J Virol 72(8): 6893-6897. Sen, E., Alam, S., and Meyers, C. 2004. Genetic and Biochemical Analysis of cis Regulatory Elements within the Keratinocyte Enhancer Region of the Human Papillomavirus Type 31 Upstream Regulatory Region during Different Stages of the Viral Life Cycle. J Virol 78(2): 612-629. Sen, E., Bromberg-White, J.L., and Meyers, C. 2002. Genetic Analysis of cis Regulatory Elements within the 5' Region of the Human Papillomavirus Type 31 Upstream Regulatory Region during Different Stages of the Viral Life Cycle. J Virol 76(10): 4798-4809. 88 Seo, Y., Muller, F., Lusky, M., Gibbs, E., Kim, H., Phillips, B., and Hurwitz, J. 1993. Bovine Papilloma Virus (BPV)-Encoded E2 Protein Enhances Binding of E1 Protein to the BPV Replication Origin. PNAS 90(7): 2865-2869. Shai, A., Brake, T., Somoza, C., and Lambert, P.F. 2007. The Human Papillomavirus E6 Oncogene Dysregulates the Cell Cycle and Contributes to Cervical Carcinogenesis through Two Independent Activities. Cancer Research 67(4): 1626-1635. Skiadopoulos, M.H. and McBride, A.A. 1998. Bovine Papillomavirus Type 1 Genomes and the E2 Transactivator Protein Are Closely Associated with Mitotic Chromatin. J Virol 72(3): 2079-2088. Spalholz, B., Yang, Y.C., and Howley, P.M. 1985. Transactivation of a bovine papilloma virus transcriptional regulatory element by the E2 gene product. Cell 42(1): 183191. Sparkowski, J., Anders, J., and Schlegel, R. 1995. E5 oncoprotein retained in the endoplasmic reticulum/cis Golgi still induces PDGF receptor autophosphorylation but does not transform cells. EMBO J 14(13): 3055-3063. Steenbergen, R.D.M., Walboomers, J.M.M., Meijer, C.J.L.M., van der Raaij-Helmer, E., Parker, J.N., Chow, L.T., and Snijders, P.J.F. 1996. Transition of human papillomavirus type 16 and 18 transfected human foreskin keratinocytes towards immortality: activation of telomerase and allele losses at 3p, 10p, 11q and/or 18q. Oncogene 13(6): 1249-1257. Stubenrauch, F., Colbert, A.M.E., and Laimins, L.A. 1998a. Transactivation by the E2 Protein of Oncogenic Human Papillomavirus Type 31 Is Not Essential for Early and Late Viral Functions. J Virol 72(10): 8115-8123. Stubenrauch, F., Lim, H.B., and Laimins, L.A. 1998b. Differential Requirements for Conserved E2 Binding Sites in the Life Cycle of Oncogenic Human Papillomavirus Type 31. J Virol 72(2): 1071-1077. Sun, S., Thorner, L., Lentz, M., MacPherson, P., and Botchan, M. 1990. Identification of a 68-kilodalton nuclear ATP-binding phosphoprotein encoded by bovine papillomavirus type 1. Journal of Virology 64(10): 5093-5105. Suprynowicz, F.A., Disbrow, G.L., Simic, V., and Schlegel, R. 2005. Are transforming properties of the bovine papillomavirus E5 protein shared by E5 from high-risk human papillomavirus type 16? Virology 332(1): 102-113. Swindle, C.S., Zou, N., Van Tine, B.A., Shaw, G.M., Engler, J.A., and Chow, L.T. 1999. Human Papillomavirus DNA Replication Compartments in a Transient DNA Replication System. J Virol 73(2): 1001-1009. Takakura, Y., Oka, N., Kajiwara, H., Tsunashima, M., Usami, S., Tsukamoto, H., Ishida, Y., and Yamamoto, T. 2009. Tamavidin, a versatile affinity tag for protein purification and immobilization. Journal of Biotechnology 145(4): 317-322. Thomas, M., Massimi, P., Jenkins, J., and Banks, L. 1995. HPV-18 E6 mediated inhibition of p53 DNA binding activity is independent of E6 induced degradation. Oncogene 10: 261-268. Thomas, M., Pim, D., and Banks, L. 1999. The role of the E6-p53 interaction in the molecular pathogenesis of HPV. Oncogene 18(53): 7690-7700. Thomas, M.C. and Chiang, C.-M. 2005. E6 Oncoprotein Represses p53-Dependent Gene Activation via Inhibition of Protein Acetylation Independently of Inducing p53 Degradation. Molecular Cell 17(2): 251-264. 89 Titolo, S., Pelletier, A., Pulichino, A.-M., Brault, K., Wardrop, E., White, P.W., Cordingley, M.G., and Archambault, J. 2000. Identification of Domains of the Human Papillomavirus Type 11 E1 Helicase Involved in Oligomerization and Binding to the Viral Origin. J Virol 74(16): 7349-7361. Tsukamoto, T., Hashiguchi, N., Janicki, S.M., Tumbar, T., Belmont, A.S., and Spector, D.L. 2000. Visualization of gene activity in living cells. Nat Cell Biol 2(12): 871878. Tsunokawa, Y., Takebe, N., Nozawa, S., Kasamatsu, T., Gissmann, L., and zur Hausen, H. 1986. Presence of human papillomavirus type-16 and type-18 DNA sequences and their expression in cervical cancers and cell lines from Japanese patients. Int J Canc 37: 499-503. Ustav, M. and Stenlund, A. 1991. Transient replication of BPV-1 requires two viral polypeptides encoded by the E1 and E2 open reading frames. EMBO J 10(2): 449-457. Van Tine, B.A., Dao, L.D., Wu, S.-Y., Sonbuchner, T.M., Lin, B.Y., Zou, N., Chiang, C.M., Broker, T.R., and Chow, L.T. 2004. Human papillomavirus (HPV) originbinding protein associates with mitotic spindles to enable viral DNA partitioning. PNAS 101(12): 4030-4035. Wang, H.-K., Duffy, A.A., Broker, T.R., and Chow, L.T. 2009. Robust production and passaging of infectious HPV in squamous epithelium of primary human keratinocytes. Genes & Development 23(2): 181-194. Wang, Q., Griffin, H., Southern, S., Jackson, D., Martin, A., McIntosh, P., Davy, C., Masterson, P.J., Walker, P.A., Laskey, P., Omary, M.B., and Doorbar, J. 2004. Functional Analysis of the Human Papillomavirus Type 16 E1{wedge}E4 Protein Provides a Mechanism for In Vivo and In Vitro Keratin Filament Reorganization. J Virol 78(2): 821-833. Wang, X., Reyes-Lamothe, R., and Sherratt, D.J. 2008a. Modulation of Escherichia coli sister chromosome cohesion by topoisomerase IV. Genes Dev 22(17): 2426-2433. Wang, X., Reyes-lamothe, R., and Sherratt, D.J. 2008b. Visualizing genetic loci and molecular machines in living bacteria. Biochemical Society Transactions 36(4): 749753. Weaver, B.A.A. and Cleveland, D.W. 2006. Does aneuploidy cause cancer? Current Opinion in Cell Biology 18(6): 658-667. Wells, S., Francis, D., Karpova, A., Dowhanick, J., Benson, J., and Howley, P. 2000. Papillomavirus E2 induces senescence in HPV-positive cells via pRB- and p21(CIP)-dependent pathways. EMBO Journal 19(21): 5762-5771. Wells, S.I., Aronow, B.J., Wise, T.M., Williams, S.S., Couget, J.A., and Howley, P.M. 2003. Transcriptome signature of irreversible senescence in human papillomavirus-positive cervical cancer cells. PNAS 100(12): 7093-7098. Woodworth, C., Cheng, S., Simpson, S., Hamacher, L., Chow, L.T., Broker, T.R., and DiPaolo, J. 1992. Recombinant retroviruses encoding human papillomavirus type 18 E6 and E7 genes stimulate proliferation and delay differentiation of human keratinocytes early after infection. Oncogene 7(4): 619-626. Yang, L., Li, R., Mohr, I., Clark, R., and Botchan, M. 1991. Activation of BPV-1 replication in vitro by the transcription factor E2. Nature 353: 628-632. 90 Yang, L., Mohr, I., Fouts, E., Lim, D., Nohaile, M., and Botchan, M. 1993. The E1 Protein of Bovine Papilloma Virus 1 is an ATP-Dependent DNA Helicase. PNAS 90(11): 5086-5090. Yao, X., Abrieu, A., Zheng, Y., Sullivan, K.F., and Cleveland, D.W. 2000. CENP-E forms a link between attachment of spindle microtubules to kinetochores and the mitotic checkpoint. Nat Cell Biol 2(8): 491. You, J., Croyle, J.L., Nishimura, A., Ozato, K., and Howley, P.M. 2004. Interaction of the Bovine Papillomavirus E2 Protein with Brd4 Tethers the Viral DNA to Host Mitotic Chromosomes. Cell 117(3): 349-360. Yu, J., Xiao, J., Ren, X., Lao, K., and Xie, X.S. 2006. Probing Gene Expression in Live Cells, One Protein Molecule at a Time. Science 311(5767): 1600-1603. Yu, T., Peng, Y.-C., and Androphy, E.J. 2007. Mitotic Kinesin-Like Protein 2 Binds and Colocalizes with Papillomavirus E2 during Mitosis. J Virol 81(4): 1736-1745. Zhao, W., Chow, L., and Broker, T. 1997. Transcription activities of human papillomavirus type 11 E6 promoter- proximal elements in raft and submerged cultures of foreskin keratinocytes. J Virol 71(11): 8832-8840. Zhao, W., Chow, L.T., and Broker, T.R. 1999a. A Distal Element in the HPV-11 Upstream Regulatory Region Contributes to Promoter Repression in Basal Keratinocytes in Squamous Epithelium. Virology 253(2): 219-229. Zhao, W., Noya, F., Chen, W.Y., Townes, T.M., Chow, L.T., and Broker, T.R. 1999b. Trichostatin A Up-Regulates Human Papillomavirus Type 11 Upstream Regulatory Region-E6 Promoter Activity in Undifferentiated Primary Human Keratinocytes. J Virol 73(6): 5026-5033. Zheng, P.-S., Brokaw, J., and McBride, A.A. 2005. Conditional Mutations in the Mitotic Chromosome Binding Function of the Bovine Papillomavirus Type 1 E2 Protein. J Virol 79(3): 1500-1509. zur Hausen, H. 2000. Papillomaviruses Causing Cancer: Evasion From Host-Cell Control in Early Events in Carcinogenesis. J Natl Cancer Inst 92(9): 690-698. -. 2002. Papillomaviruses and Cancer: from Basic Studies to Clinical Application. Nature Reviews Cancer 2(5): 342-350.