You are on page 1of 17

Clinical Science (2006) 110, 525–541 (Printed in Great Britain) doi:10.

1042/CS20050369 525

R E V I E W

Molecular biology of human papillomavirus


infection and cervical cancer

John DOORBAR
Division of Virology, National Institute for Medical Research, The Ridgeway, Mill Hill, London NW7 1AA, U.K.

A B S T R A C T

HPVs (human papillomaviruses) infect epithelial cells and cause a variety of lesions ranging from
common warts/verrucas to cervical neoplasia and cancer. Over 100 different HPV types have
been identified so far, with a subset of these being classified as high risk. High-risk HPV DNA is
found in almost all cervical cancers (> 99.7 %), with HPV16 being the most prevalent type in both
low-grade disease and cervical neoplasia. Productive infection by high-risk HPV types is manifest
as cervical flat warts or condyloma that shed infectious virions from their surface. Viral genomes
are maintained as episomes in the basal layer, with viral gene expression being tightly controlled
as the infected cells move towards the epithelial surface. The pattern of viral gene expression in
low-grade cervical lesions resembles that seen in productive warts caused by other HPV types.
High-grade neoplasia represents an abortive infection in which viral gene expression becomes
deregulated, and the normal life cycle of the virus cannot be completed. Most cervical cancers
arise within the cervical transformation zone at the squamous/columnar junction, and it has
been suggested that this is a site where productive infection may be inefficiently supported. The
high-risk E6 and E7 proteins drive cell proliferation through their association with PDZ domain
proteins and Rb (retinoblastoma), and contribute to neoplastic progression, whereas E6-mediated
p53 degradation prevents the normal repair of chance mutations in the cellular genome. Cancers
usually arise in individuals who fail to resolve their infection and who retain oncogene expression
for years or decades. In most individuals, immune regression eventually leads to clearance of the
virus, or to its maintenance in a latent or asymptomatic state in the basal cells.

HPVs (HUMAN PAPILLOMAVIRUSES) genera (Figure 1), and the lesions they cause have different
AND DISEASE characteristics.
The two main HPV genera are the Alpha and Beta
HPVs cause a diverse range of epithelial lesions. Over 100 papillomaviruses, with approx. 90 % of currently char-
different HPV types have been identified based on DNA acterized HPVs belonging to one or other of these
sequence analysis [1], with each being associated with groups. Beta papillomaviruses are typically associated
infection at specific epithelial sites [2]. At an evolutionary with inapparent cutaneous infections in humans but,
level, HPVs fall into a number of distinct groups or in immunocompromised individuals and in patients

Key words: cervical cancer, epithelial cell, human papillomavirus, immunity, infection, neoplasia.
Abbreviations: ARF, ADP-ribosylation factor; BPV, bovine papillomavirus; CIN, cervical intraepithelial neoplasia; COPV, canine
oral papillomavirus; DAPI, 4,6-diamidino-2-phenylindole; Dlg, the Drosophila discs large protein; E6AP, E6-associated protein;
EV, epidermodysplasia cerruciformis; HPV, human papillomavirus; HSIL, high-grade squamous intraepithelial neoplasia lesions;
Hsp, heat-shock protein; hTERT, human telomerase reverse transcriptase; LSIL, low-grade squamous intraepithelial neoplasia
lesions; MCM, multicopy maintenance; ORF, open reading frame; PCNA, proliferating-cell nuclear antigen; PML, promyelocytic
leukaemia; Rb, retinoblastoma; pRb, Rb protein; ROPV, rabbit oral papillomavirus; SCC, squamous cell carcinoma; URR, upstream
regulatory region.
Correspondence: Dr John Doorbar (email jdoorba@nimr.mrc.ac.uk).


C 2006 The Biochemical Society
526 J. Doorbar

melanoma skin cancer [3,4]. EV patients carry mutations


in their TMC6 (previously known as EVER1) or TMC8
(previously known as EVER2) genes, which renders them
susceptible to these viruses [5,6].
The largest group of HPVs comprise the Alpha papi-
llomaviruses, and it is this group that contains the genital/
mucosal HPV types (Figure 1). The Alpha papilloma-
viruses also include cutaneous viruses such as HPV2,
which cause common warts, and which are only very
rarely associated with cancers [7]. More than 30 different
HPV types are known to infect cervical epithelium, with
a subset of these being associated with lesions that can
progress to cancer. These cancer-associated HPVs are
classified as high-risk HPV types. HPV16 is the most
prevalent high-risk HPV in the general population, and
is responsible for approx. 50 % of all cervical cancers. The
remaining mucosal types are classified as intermediate or
low risk depending on the frequency with which they are
found in cancers. Low-risk HPV types, such as HPV11,
are associated with cervical cancer only very rarely,
but are still medically important because they cause
genital warts. Genital warts are a major sexually trans-
mitted disease in many countries and can affect 1–2 % of
young adults [8].
The remaining HPVs come from three genera
(Gamma, Mu and Nu) and generally cause cutaneous
papillomas and verrucas that do not progress to cancer
(but see [9,10]). The relationship between the different
genera of HPVs, and in particular the high-risk HPV
types associated with cervical cancer, is shown in Figure 1.

NORMAL PRODUCTIVE INFECTION


Many HPV types produce only productive lesions fol-
lowing infection and are not associated with human
cancers. In such lesions, the expression of viral gene prod-
ucts is carefully regulated, with viral proteins being
produced at defined times and at regulated levels as the
Figure 1 HPVs and their association with cervical disease infected cell migrates towards the epithelial surface.
(A) HPVs are contained within five evolutionary groups. HPV types that infect The events that lead to virus synthesis in the upper
the cervix come from the Alpha group which contains over 60 members. HPV epithelial layers appear common to both the low- and
types from the Beta, Gamma, Mu and Nu groups primarily infect cutaneous sites. high-risk HPV types, with protein expression patterns
(B) The Alpha papillomaviruses can be subdivided into three categories (high risk, in low-grade CIN (cervical intraepithelial neoplasia) 1
low risk and cutaneous), depending on their prevalence in the general population resembling the patterns found in benign warts caused
and on the frequency with which they cause cervical cancer (shown in the right- by other papillomaviruses. Productive infection can be
most column). High-risk types come from the Alpha 5, 6, 7, 9 and 11 groups. The divided into distinct phases, with the different viral
frequency with which the different HPV types are found in cervical cancers (squam- proteins playing specific roles.
ous cell carcinoma and adenocarcinoma/adenosquamous carcinoma) is shown in
the central columns (based on information contained in [170]). Where no percent- Establishment of HPV infection
age is listed, the HPV type is not generally associated with cervical cancer. All papillomaviruses share a number of characteristics
and contain double-stranded circular DNA within an
icosahedral capsid (Figure 2). Although the viral genome
suffering from the inherited disease EV (epidermodys- can vary slightly in size between different HPV types,
plasia verruciformis), these viruses can spread unchecked it typically contains around 8000 bp [7904 bp for HPV16
and become associated with the development of non- (GenBank® accession number NC 001526)], and encodes


C 2006 The Biochemical Society
Papillomavirus molecular biology 527

eight or nine ORFs (open reading frames) (Figure 2). The to mitotic chromosomes is critical for correct segregation
virus shell is made up of two coat proteins. The L1 protein [30] and, in some papillomavirus types, involves the
is the primary structural element, with infectious virions cellular Brd4 protein, which directly associates through
containing 360 copies of the protein organized into 72 its C-terminus with the viral E2 protein [30,31]. For
capsomeres [11]. L2 is a minor virion component, and the high-risk HPV types that are associated with human
it is thought that a single L2 molecule may be present cancers, association appears to be via the spindle, with
in the centre of the pentavalent capsomeres at the virion additional cellular proteins being involved. In addition
vertices [11,12]. Both proteins play an important role in to its role in replication and genome segregation, E2 can
mediating efficient virus infectivity. also act as a transcription factor and can regulate the viral
Infection by papillomaviruses requires that virus part- early promoter (p97 in HPV16; p99 in HPV31) and con-
icles gain access to the epithelial basal layer and enter the trol expression of the viral oncogenes (E6 and E7). At
dividing basal cells. At present there is some controversy low levels, E2 acts as a transcriptional activator, whereas
as to the precise nature of the receptor for virus entry at high levels E2 represses oncogene expression by dis-
[13], but it is thought that heparan sulphate proteoglycans placing SP1 transcriptional activator from a site adjacent
may play a role in initial binding and/or virus uptake to the early promoter (Figure 2). It is unclear whether
[13–15]. As with other viruses [16,17], it seems that HPV E2-mediated transcriptional regulation is important in
infection requires the presence of secondary receptors for basal epithelial cells, where the nucleosomal structure
efficient infection, and it has been suggested that this role and/or methylation status of the viral DNA may not
may be played by the α6 integrin [18–20]. Papillomavirus be compatible with efficient activation of the early pro-
particles are taken into the cell relatively slowly following moter [32,33]. The analysis of genome maintenance in
binding [21] and, for HPV16, this occurs by clathrin- BPV (bovine papillomavirus)-transformed keratinocytes
coated endocytosis [22]. This mode of entry may not be and cell lines derived from cervical lesions [34,35] has
conserved among all HPV types, however, and it has been suggested that viral episomes may be maintained at 10–
suggested that HPV31 may gain entry via caveolae [23]. 200 copies in basal cells, although, as yet, this has not been
Papillomavirus particles disassemble in late endosomes adequately demonstrated in lesions. Knockout mutants in
and/or lysosomes, with the transfer of viral DNA to the viral genome have suggested that both E1 and E2 are
the nucleus being facilitated by the minor capsid protein required for viral genome maintenance in the basal layer
L2 [22,24]. In experimental systems, viral transcripts can [36,37].
be detected as early as 12 h post-infection, with mRNA
levels increasing over the course of several days [24]. Stimulation of cell proliferation
Infection leads to the establishment of the viral genome A number of model systems have been used to examine
as a stable episome (without integration into the host cell the papillomavirus productive cycle during in vivo infec-
genome) in cells of the basal layer, and this is thought to tion. Following experimental inoculation of mucosal
require expression of the viral replication proteins, E1 and epithelial tissue by ROPV (rabbit oral papillomavirus)
E2. The E2 protein plays several roles during productive or COPV (canine oral papillomavirus), an increase in
infection and, in basal cells, is required for the initiation cell proliferation in the basal and suprabasal cellular
of viral DNA replication and genome segregation. E2 is compartments is readily apparent, with mature warts
a DNA-binding protein that recognizes a palindromic becoming visible by 4 weeks post-infection [38,39]. The
motif [AACCg(N4 )cGGTT] in the non-coding region of time between initial infection and the appearance of pro-
the viral genome ([25], and Figure 2). HPV16 has four ductive papillomas can vary depending on the titre of
such motifs, one of which lies adjacent to the viral origin the infecting virus and possibly also on the nature of the
of replication. E2 binding is necessary for the recruit- infecting papillomavirus type, and it has been suggested
ment of the E1 helicase to the viral origin, which binds that latency may result when inoculating titres are low
in turn to cellular proteins necessary for DNA repli- [40,41]. In cervical lesions caused by HPVs, the increased
cation, including RPA (replication protein A) and DNA proliferation of suprabasal epithelial cells is attributed
polymerase α primase [26–29]. The E2 protein sub- to the expression of the viral oncogenes, E6 and E7.
sequently dissociates from the viral origin, allowing the During natural infection, the activity of these genes
assembly of E1 into a double hexameric ring that func- allows the small number of infected cells to expand, in-
tionally resembles the hexameric ring structures that creasing the number of cells that subsequently go on to
assemble at cellular replication origins and which are produce infectious virions. The ability of E6 and E7
built from the cellular MCM (multicopy maintenance) to drive cells into S-phase is also necessary, along with
proteins. E1 and E2, for the replication of viral episomes above
In the basal cells, it appears that the viral genome the basal layer. Suprabasal cells normally exit the cell
replicates with the cellular DNA during S-phase, with cycle and begin the process of terminal differentiation in
the replicated genomes being partitioned equally during order to produce the protective barrier that is normally
cell division. The role of E2 in anchoring viral episomes provided by the skin [42]. In HPV-infected keratinocytes,


C 2006 The Biochemical Society
528 J. Doorbar

Figure 2 Organization of the HPV genome and the virus life cycle

(A) The HPV16 genome (7904 bp) is shown as a black circle with the early (p97) and late (p670) promoters marked by arrows. The six early ORFs [E1, E2, E4 and
E5 (in green) and E6 and E7 (in red)] are expressed from either p97 or p670 at different stages during epithelial cell differentiation. The late ORFs [L1 and L2
(in yellow)] are also expressed from p670, following a change in splicing patterns, and a shift in polyadenylation site usage [from early polyadenylation site (PAE)
to late polyadenylation site (PAL)]. All the viral genes are encoded on one strand of the double-stranded circular DNA genome. The long control region (LCR from
7156–7184) is enlarged to allow visualization of the E2-binding sites and the TATA element of the p97 promoter. The location of the E1- and SP1-binding sites is
also shown. (B) The key events that occur following infection are shown diagrammatically on the left. The epidermis is shown in colour with the underlying dermis
being shown in grey. The different cell layers present in the epithelium are indicated on the left. Cells in the epidermis expressing cell cycle markers are shown with
red nuclei. The appearance of such cells above the basal layer is a consequence of virus infection, and in particular, the expression of the viral oncogenes, E6 and
E7. The expression of viral proteins necessary for genome replication occurs in cells expressing E6 and E7 following activation of p670 in the upper epithelial layers
(cells shown in green with red nuclei). The L1 and L2 genes (yellow) are expressed in a subset of the cells that contain amplified viral DNA in the upper epithelial


C 2006 The Biochemical Society
Papillomavirus molecular biology 529

however, the restraint on cell-cycle progression is lost and cell proliferation independently of E7 through its C-
normal terminal differentiation does not occur [43]. The terminal PDZ ligand domain [the name PDZ is derived
basic mechanism by which papillomaviruses stimulate from the first three proteins in which these domains
cell-cycle progression is well known and is similar to the were found: PSD-95 (a 95 kDa protein involved in
way that other tumour viruses deregulate cell growth. signalling), Dlg (the Drosophila discs large protein), and
E7 associates with pRb [Rb (retinoblastoma) protein] ZO1 (the zonula occludens 1 protein which is involved in
and other members of the pocket protein family and maintaining epithelial cell polarity)] [51]. E6 PDZ binding
disrupts the association between pRb and the E2F family can mediate suprabasal cell proliferation [52,53] and may
of transcription factors, irrespective of the presence of contribute to the development of metastatic tumours by
external growth factors (Figure 3). E2F subsequently disrupting normal cell adhesion. In productive lesions,
transactivates cellular proteins required for viral DNA cells are driven into cycle only in the lower epithelial
replication such as cyclins A and E. E7 also associates with layers and extend towards the epithelial surface to varying
other proteins involved in cell proliferation, including extents depending on lesion grade and the nature of the
histone deacetylases [44], components of the AP1 tran- infecting HPV type [54]. In benign warts caused by
scription complex [45] and the cyclin-dependent kinase HPV1, proliferating cells are restricted to the epithelial
inhibitors p21 and p27 [46]. During natural infection, basal layer, with genome amplification beginning as soon
however, the ability of E7 to drive cell proliferation is as the infected cell enters the suprabasal cell layers. HPVs,
inhibited in some cells, depending on the levels of the p21 such as HPV6 and HPV16 (Alpha papillomaviruses),
and p27 cyclin-dependent kinase inhibitors (Figure 3). typically have a region that may be many cell layers thick,
High levels of p21 and p27 in differentiating keratinocytes where cells are retained in cycle prior to the onset of
can lead to the formation of inactive complexes with E7 vegetative viral genome amplification [54]. In the case
and cyclinE within the cell. It appears that the ability of of the high-risk HPV types that are associated with
E7 to drive cells through mitosis in differentiating epi- cervical cancer, the relative thickness of these layers
thelium may be limited to those cells which express p21 increases with the grade of neoplasia, while the extent
and p27 at low level or which express sufficient E7 to of epithelial differentiation decreases.
overcome the block to cell-cycle progression [47]. This is
important given that deregulated expression of the viral
oncogenes is a predisposing factor in the development Genome amplification
of HPV-associated cancers (Figure 3). The function of Although cell proliferation is required for lesion form-
the viral E6 protein complements that of E7 and, in the ation and the maintenance of viral episomes, all papi-
high-risk HPV types, the two proteins are expressed llomaviruses must eventually amplify and package their
together from a single polycistronic mRNA species [48]. genomes if infectious virions are to be produced. What
A primary role of E6 is its association with p53 which, in triggers the onset of late events is not yet fully understood,
the case of the high-risk HPV types, mediates p53 ubi- but appears to depend in part on changes in the cellular
quitination and degradation. This is thought to prevent environment as the infected cell moves towards the
growth arrest or apoptosis in response to E7-mediated epithelial surface. Critical for this is the up-regulation
cell-cycle entry in the upper epithelial layers, which might of the differentiation-dependent promoter, which for
otherwise occur through activation of the ARF (ADP- many HPV types is contained within the E7 ORF (P670
ribosylation factor) pathway (see Figure 3). The general in HPV16; p742 in HPV31) [55,56]. The activation of the
role of E6 as an anti-apoptotic protein is emphasized differentiation-dependent promoter depends on changes
further by the finding that it also associates with Bak [49] in cellular signalling, rather than on genome ampli-
and Bax [50]. This role of E6 is of key significance in the fication [55,56], and leads to an increase in the level of
development of cervical cancers (discussed in more detail viral proteins necessary for replication (i.e. E1, E2, E4
below), as it compromises the effectiveness of the cellular and E5). Cells supporting productive infection can be
DNA damage response and allows the accumulation of visualized using antibodies to E4 (Figure 4), whereas
secondary mutations to go unchecked. The E6 protein those supporting genome amplification also contain E7
of the high-risk HPV types also plays a role in mediating [57].

layers. Cells containing infectious particles are eventually shed from the epithelial surface (cells shown in green with yellow nuclei). In cutaneous tissue, this follows
nuclear degeneration and the formation of flattened squames. The timing and extent of expression of the various viral proteins are summarized using arrows at the
right of the Figure. The consequence of expressing viral gene products in this ordered way is shown on the far right. The expression of E6 and E7 in the presence of
low levels of E1, E2, E4 and E5 allows maintenance of the viral genome (genome maintenance/cell proliferation). Elevation in the levels of these replication proteins
facilitates viral genome amplification. The first appearance of L2 allows genome packaging to begin, with the expression of L1 allowing the formation of infectious
virions (virus assembly). The accumulation of E4 close to the epithelial surface may improve the efficiency of virus release.


C 2006 The Biochemical Society
530 J. Doorbar

Figure 3 Stimulation of cell-cycle progression by high-risk HPV types


HPV infection leads to deregulation of the cell cycle. Regulation of protein expression in uninfected epithelium is shown in (A). In the presence of high-risk HPV
(B), the regulation of proteins necessary for cell proliferation is altered, allowing HPV to stimulate S-phase entry in the upper epithelial layers. (A) Uninfected epi-
thelium. The expression of proteins necessary for cell-cycle progression is controlled by pRB, which in non-cycling cells associates with members of the E2F transcription
factor family (centre). In the presence of growth factors, cyclinD/Cdk4/6 is activated, which leads to Rb phosphorylation and the release of the transcription factor
E2F, which drives the expression of proteins involved in S-phase progression. p16 regulates the levels of active cyclinD/Cdk in the cell, providing a feedback mechanism
that regulates the levels of MCM, PCNA (proliferating-cell nuclear antigen) and cyclinE. p14Arf , whose expression is directly linked to that of p16, regulates the activity
of the MDM (murine double minute) ubiquitin ligase, which maintains p53 at a level below that required for cell cycle arrest and/or apoptosis. (B) HPV-infected
epithelium. In cervical epithelium infected by high-risk HPV types, progression through the cell cycle is not dependent on external growth factors, but is stimulated
by the E7 protein, which binds and degrades pRB and facilitates E2F-mediated expression of cellular proteins necessary for S-phase entry. Although p16 levels rise,
normal feedback is by-passed, as HPV-mediated cell proliferation is not dependent on cyclinD/Cdk4/6. The rise in the level of p14Arf , which occurs in the absence
of p16-mediated feedback, leads to the inhibition of MDM function and an increase in the level of p53. This is countered by E6, which associates with the E6AP
ubiquitin ligase in order to stimulate the degradation of the p53 protein and prevent growth arrest and/or apoptosis. In low-grade cervical disease, where E7 levels
are carefully regulated, it is thought that E7-mediated cell proliferation can sometimes be inhibited as a result of association with p21 and cyclin E/Cdk. The high
levels of E7 found in cervical cancer cells are thought to overcome this block by binding and inactivating the Cdk inhibitor p21.

The role of E1 and E2 in viral genome amplification in particular Hsp40 and Hsp70, and this contributes to
is well established. E1 is highly conserved among papi- the formation of E1 dihexamers [61]. E1 and E2 also act
llomaviruses and has a weak affinity for a consensus to regulate the viral early promoter (p97 in HPV16 and
motif (AACNAT) repeated 6 times in the viral origin. p99 in HPV31), with high levels of E2 acting to down-
During natural infection, E1 is expressed at very low regulate the expression of E6 and E7 in experimental
levels and requires the presence of E2 in order to be effi- systems. The ability of E2 to either repress or activate
ciently targeted to its binding sites. E2 associates with early viral gene expression according to its abundance
E1 primarily through its N-terminus and binds to DNA is thought to result from differences in the affinity of
as a dimer through its C-terminus [58,59] (Figure 2). E2 for its various binding sites [62]. In HPV16, it is
The formation of an E1–E2 complex at the origin of thought that binding site 4 is the primary site that is occu-
replication induces localized distortion in the viral DNA, pied when E2 is present at low levels and that binding to
which facilitates the recruitment of additional E1 mol- this site and to binding site 3 (Figure 2) leads to pro-
ecules and the eventual displacement of E2 [60]. E1 can moter activation [63]. As E2 increases in abundance,
also associate with cellular Hsps (heat-shock proteins), occupancy of the remaining sites leads to the displacement


C 2006 The Biochemical Society
Papillomavirus molecular biology 531

expression that is important in stimulating viral genome


amplification will lead eventually to the down-regulation
of E6/E7 expression and to the eventual loss of the
replicative environment necessary for viral DNA syn-
thesis. This curious link between replication and tran-
scription provides a mechanism by which the virus can
limit the timing and duration of genome amplification. In
cervical lesions caused by HPV16, cells supporting viral
genome amplification are often scarce and vary greatly
in their location within the lesion [57]. By contrast, in
verrucas caused by HPV1, cells supporting genome
amplification are relatively prevalent and are consistently
found immediately above the basal layer.
Although E1 and E2 are key players in viral genome
amplification, the viral E4 and E5 proteins also contribute
[65–69]. E5 mutant genomes of HPV16 and 31 exhibit
lower levels of genome amplification than wild type,
and it is thought that the ability of E5 to modulate cell
signalling is responsible for this. HPV E5 is a trans-
membrane protein that resides predominantly in the ER
(endoplasmic reticulum), but which can associate with
the vacuolar proton ATPase and delay the process of
endosomal acidification [70–72]. It is thought that this
affects the recycling of growth factor receptors on the cell
surface, leading to an increase in EGF (epidermal growth
factor)-mediated receptor signalling and the maintenance
of a replication competent environment in the upper
epithelial layers [73]. By contrast, the role of E4 in
genome amplification is not yet fully established. E4
accumulates in the cell at the time of viral genome ampli-
fication, and its loss has been shown to disrupt late events
in a number of experimental systems [67–69]. Part of
the effect may be related to the ability of certain HPV
types to associate with cyclinB/Cdk2 and to relocate
the complex to the cytoplasm [74,75]. This prevents the
nuclear accumulation of cyclinB/Cdk2, which is neces-
sary for progression through mitosis. The ability of E4
to cause cell-cycle arrest in G2 and to antagonize E7-
Figure 4 Characterization of events during productive mediated cell proliferation is a common feature of the
papillomavirus infection by immunostaining E4 proteins encoded by several HPV types, including
The pattern of viral protein expression is conserved in productive lesions caused by HPV16, HPV11 [75], HPV18 [76] and HPV1 [77]. Inter-
papillomaviruses of diverse origins. Typical immunofluorescence staining patterns estingly, recent work has suggested that the E4 proteins
are shown. (A) Productive papilloma showing the typical expression profile of PCNA of HPV16 and HPV18 can also associate with E2, which
[a surrogate marker of E6/E7 gene expression (in red)] and E4 (in green), which suggests an additional mechanism by which E4 may act
is thought to reflect the sites of expression of E1 and E2. Nuclei are counter- [78].
stained with DAPI (4,6-diamidino-2-phenylindole; in blue). A small area of un-
infected tissue is apparent on either side of the papilloma. (B) Comparison of Virus assembly and release
the sites of viral genome amplification [in red; detected by FISH (fluorescence The final stage in the papillomavirus productive cycle
in situ hybridization)] and E4 expression (in green) reveal that the two events cor- requires that the replicated genomes are packaged into
relate very closely. Nuclei are counterstained with DAPI (blue). (C) The viral capsid infectious particles. Capsid proteins (L1 and L2) accumu-
protein (L1; in red) is expressed in a subset of the cells that express E4 (in green) late after the onset of genome amplification, with L2
in the upper epithelial layers. Nuclei are counterstained with DAPI (in blue). expression preceding the expression of L1 [79,80]. The
events that link genome amplification to the synthesis
of basal transcription factors, such as Sp1 and TBP of the capsid proteins are not yet fully understood, but
(TATA-box-binding protein), that are necessary for pro- are dependent on changes in mRNA splicing and on
moter activation [64]. It appears that the increase in E2 the generation of transcripts that terminate at the late


C 2006 The Biochemical Society
532 J. Doorbar

(rather than the early) polyadenylation site (Figure 2). detect infection. Ultimately, virus release requires effi-
During epithelial cell differentiation, the timing of capsid cient escape from the cornified envelope at the cell surface,
synthesis is regulated both at the level of RNA processing which may be facilitated by the E4 protein. E4 can disrupt
and at the level of protein synthesis [81–83]. Negative the keratin network [105,106] and can affect the integrity
regulatory elements that control RNA stability are of the cornified envelope [104,107].
present in the coding regions and in the late untranslated
region of HPV16 [84,85], whereas a splicing silencer in
the HPV16 L1 gene leads to the preferential synthesis ABORTIVE INFECTION AS A PRECURSOR
of early transcripts in proliferating cells [86]. In addition TO CANCER
to this, the pattern of codon usage within the HPV16
L1 and L2 genes is distinct from the pattern usually Virus-induced cancers often arise at sites where pro-
found in mammalian cells, which contributes further to ductive infection cannot be properly supported. CRPV
the inhibition of capsid expression in the lower epithelial (cottontail rabbit papillomavirus) induces productive
layers [82,87,88]. Similar regulatory mechanisms are also papillomas in its natural host (the cottontail rabbit),
thought to exist in HPV1 [89], HPV31 [90] and BPV [91] but gives rise to lesions that cannot support virus syn-
and are likely to extend to other papillomavirus types. thesis when inoculated into domestic rabbits. Such abor-
The assembly of infectious virions in the upper epi- tive infections progress to cancer more frequently than
thelial layers is thought to require E2 in addition to do infections in cottontails. A similar situation is seen
the capsid proteins L1 and L2 [92,93], and it has been following the inoculation of SV40 (simian virus 40) and
suggested that E2 may improve the efficiency of genome adenovirus type 5 (which are also considered to be
encapsidation during natural infection [93,94]. L2 local- DNA tumour viruses) into inappropriate hosts, such
izes to the nucleus by virtue of nuclear localization as hamsters and rats, which support early, but not late,
signals located at its N- and C-termini and, once there, it viral gene expression. The papillomavirus types that
associates with PML (promyelocytic leukaemia) bodies. usually cause benign cutaneous warts in humans and are
Although some papillomavirus L2 proteins can associate considered low risk (e.g. HPV2 and 4) can be associated
directly with DNA [95], the specific recruitment of occasionally with cancers at mucosal sites. The restricted
viral genomes to PML bodies is thought to require E2, tissue tropism of the different HPV types, coupled with
which can associate with viral DNA through its specific their ability to infect non-optimal sites, may provide a
recognition sites. L1 assembles into capsomeres in the partial explanation as to why otherwise benign HPV
cytoplasm prior to nuclear relocation and is recruited into types are occasionally associated with human cancers.
PML bodies only after L2 has bound and has displaced The high-risk papillomavirus types that are associated
the PML component sp100 [79]. The assembly of virus- with human cancers more frequently come predomin-
like particles in experimental systems requires the pres- antly from the Alpha 9 and Alpha 7 groups (see Figure 1),
ence of the cellular protein Hsp70 [96], which is also with HPV16 and HPV18 being the most prevalent
involved in recycling the viral E1 protein during repli- types [108]. These viruses are found in women with no
cation [97], but does not appear to be strictly dependent cytological abnormalities, as well as in women with LSIL
on PML localization [98]. Although papillomavirus part- and HSIL (low- and high-grade squamous intraepithelial
icles can assemble in the absence of L2, its presence con- neoplasia lesions respectively) and/or cancer [109–111].
tributes to efficient packaging [99] and enhances virus By PCR, HPV16 DNA is apparent in approx. 26 %
infectivity [100]. Loss of L2 in the context of the HPV31 of LSIL [112], but can be seen in as many as 63 % of
genome results in a 10-fold reduction in packaging effi- SCCs (squamous cell carcinomas) of the cervix [113].
ciency and a 100-fold reduction in virus infectivity when HPV18 is the second most common HPV type associated
compared with wild-type HPV31 [101]. L2 associates with this disease (causing 10–14 % of all cases of
with L1 through a hydrophobic region near the C-ter- cervical SCCs), but is the primary HPV type associated
minus of the protein that is thought to insert into the with cervical adenocarcinoma, causing 37–41 % (HPV16
central hole in the pentavalent L1 capsomeres [102]. causes 26–37 % of all such cases [113]). The prevalence of
The interaction between capsomeres requires the C- high-risk HPV infection amongst young women (mean
terminus of the L1 protein [11] with virus maturation and age, 25 years) is typically approx. 20–40 % depending
stabilization occurring as the infected cells approach the on geographical location [111,114], with the incidence
epithelial surface as a result of disulphide cross-linking. declining with age as infections are either resolved or
Although papillomaviruses are resistant to desiccation brought under control by the host immune system. HPV
[103], it is thought that their survival may be enhanced infections are very common in this age group, with
by being shed from the epithelial surface as a cornified cumulative incidence of infection over 5 years being as
squame [104]. The retention of papillomavirus antigens high as 60 % in some populations [110]. Most women
until the infected cell reaches the epithelial surface is (80 %) infected with a specific HPV type will, however,
thought to limit the ability of the immune system to show no evidence of that type after 18 months, and it


C 2006 The Biochemical Society
Papillomavirus molecular biology 533

Figure 5 Changes in expression patterns that accompany progression to cervical cancer


During cancer progression, the pattern of viral gene expression changes. In CIN1 (LSIL), the order of events is generally similar to that seen in productive lesions
(shown diagrammatically on the left). In CIN2 and CIN3, however, the onset of late events is retarded, and although the order of events remains the same, the
production of infectious virions becomes restricted to smaller and smaller areas close to the epithelial surface. Integration of HPV sequences into the host cell genome
can accompany these changes and can lead to further deregulation in the expression of E7 (and the loss of the E1 and E2 replication proteins). In cervical cancer
(shown on the right), the productive stages of the virus life cycle are no longer supported and viral episomes are usually lost.

Figure 6 Detection of viral proteins in cervical intraepithelial neoplasia


(A) The expression of surrogate markers of E7 (in this case MCM; in red) and E4 (in green) in HPV16-infected cervix resembles the pattern of expression seen in
productive papillomas caused by other papillomavirus types (see also Figure 4). Nuclei are counterstained with DAPI (in blue) and uninfected tissue is apparent either
side of the area of infection. (B) During cancer progression, the expression of surrogate markers of E7 (in this case MCM; in red) extends towards the epithelial surface
and the expression of E4 is restricted to isolated pockets close to the upper epithelial layers. CIN3 is shown on the left, whereas the pattern seen in many productive
CIN1 lesions is shown on the right. Nuclei are counterstained with DAPI (in blue).

is generally thought that re-infection by the same HPV in the absence of screening (0.018–0.044 %)] with most
type is uncommon [115]. The development of high- infections being successfully resolved [117].
grade cervical neoplasia arises in women who cannot Low-grade cervical lesions (i.e. LSIL or grade 1 CIN)
resolve their infection and who maintain persistent caused by high- and low-risk HPV types are similar to
active infection for years or decades following initial productive infections in the pattern of viral gene expres-
exposure. In such women, the continuous stimulation sion, and viral coat proteins can usually be detected in
of S-phase entry and cell proliferation by E7, coupled cells at the epithelial surface [57] (Figure 5). High-grade
with the loss of p53-mediated DNA repair pathways as lesions (HSIL or grade 2 or 3 CIN) have a more extensive
a result of E6 expression, allows the accumulation of proliferative phase, with the productive stages of the virus
secondary point mutations in the cellular genome that life cycle being supported only poorly (Figures 5 and 6).
eventually lead to cancer. Previous work from a num- It has been estimated that approx. 20 % of CIN1 will
ber of laboratories has suggested that not all cervical progress to CIN2, and that approx. 30 % of these lesions
cancers develop through this route, however, and that, in will progress to more severe neoplasia if left untreated.
some instances rapid onset of HSIL can arise after initial Approx. 40 % of CIN3 lesions can progress to cancer
infection [110,116]. Given the prevalence of genital HPV [118], with cervical neoplasia generally arising within
types in the general population and the high life-time risk the cervical transformation zone where the columnar
of infection (estimated at approx. 80 %), the incidence cells of the endocervix meet the stratified squamous
of cervical cancers is very low [typically approx. 0.03 % epithelial cells of the ectocervix. The extent of columnar


C 2006 The Biochemical Society
534 J. Doorbar

and stratified epithelium in this region changes markedly chromosome is a critical event in the development of most
during a woman’s life, and it is in this region of change that cervical cancers and, although this can occur randomly
most cervical neoplasias develop. It has been suggested throughout the genome, several studies have indicated
that the transformation zone may be a sub-optimal site a preference for integration at common fragile sites
for completion of the productive life cycle of high-risk and have suggested that changes in the expression of
HPV types such as HPV16. Interestingly, high-risk HPV genes at or near the integration site may participate
infections can be found at many sites in the anogenital in cancer development [125]. It is clear, however, that
tract, including the vagina, vulva and penis, but, despite integration (which often results in the loss of E2) can
this distribution, the incidence of cancer at these sites is lead to the deregulation of E6/E7 expression, and that
generally low (0.001 %). Only at the anus, in men who this is critical for the enhanced growth characteristics
have sex with men, does the incidence of HPV-associated of cervical cancer cells. The requirement of these genes
cancer rise to levels that are similar to those found at the for the maintenance of the cancer phenotype is shown
cervix (0.035 %) [119], and it is interesting that at both most dramatically by studies that have aimed to inhibit
these susceptible sites there is a transformation zone. In viral oncogene activity in cervical cancer cells. Cell lines
the cervix, the reserve cells of the transformation zone such as HeLa and SiHa, which have been grown in the
eventually form the basal cells of a stratified squamous laboratory for many decades, will undergo apoptotic
epithelium, and the higher frequency of cancers at such cell death in the presence of molecules that inhibit
sites may reflect the fact that the virus can access these E6 function [126–128]. Similar results are obtained
basal cells more readily than those that are protected following the reintroduction of E2 into such cell lines,
from infection by a permanent stratified epithelial layer. which suppresses the expression of the viral oncogenes
Despite this possibility, it appears that the transformation by binding to the URR (upstream regulatory region),
zone may in fact be an epithelial site where high-risk HPV preventing continued cell proliferation [129,130]. The
types cannot properly regulate their productive cycle, ubiquitous retention of the E6/E7 region of the viral
and that variation both in the level and in the timing of genome following integration into the host chromosome
expression of viral proteins may underlie the development is usually accompanied by the loss or disruption of viral
of cancers at these sites. sequences encoding most of E1, as well as E2 and E4. The
viral E2 protein has a negative effect on cell proliferation
by regulating the viral URR and can also cause cell-cycle
MOLECULAR EVENTS DURING CANCER arrest at the G2 phase of the cell cycle [131,132]. Similarly,
PROGRESSION the E4 protein of HPV16 is able to inhibit mitosis
by preventing the nuclear localization of cyclinB/Cdk1,
The identification of cervical lesions as flat condyloma, and it is not surprising therefore that the expression of
LSIL, HSIL or invasive cervical cancer reflects molecular these proteins is abrogated in HPV-associated cancers. In
changes in the normal programme of epithelial cell differ- addition to the loss of regulatory proteins, integration
entiation that occur following infection. Virus production also leads to the loss of sequences at the 3 -end of the
at the epithelial surface depends on the ordered and timely viral early transcripts that can suppress the production
expression of viral gene products (Figure 2) [54,57], of viral mRNA species encoding E6 and E7 [133–
with the timing of such events becoming progressively 135], contributing further to the deregulation of viral
disturbed during neoplastic progression (Figures 5 and oncogene expression. Although integrated HPV DNA
6). In HSIL, viral genome amplification occurs closer is found in the vast majority of cervical cancers, other
to the epithelial surface than in condylomas and LSIL, factors are likely to influence the development of the
and the expression of viral coat proteins is retarded precancerous changes seen in squamous intraepithelial
[57] (Figure 5). The molecular bases for these changes neoplasia. These include exposure to glucocorticoids and
are not fully understood, but may in some instances progesterones, which can affect viral oncogene expression
reflect changes in the levels of E6 and E7 expression [136–138], and the regulation of viral gene expression by
that occur following integration of the viral genome DNA methylation and chromatin organization [139].
into the host cell chromosome. Integrated HPV DNA Both factors can also regulate expression from integrated
is found in most invasive cancers and in a subset of sequences and, in instances where integration occurs as
high-grade lesions [120,121], but can also be found in concatamers, it is often only one copy of the viral genome
some CIN1 lesions, and it has been suggested that that is transcriptionally active [140].
integration may be an early event in cancer progression Although many HPV types can infect the cervix, only
[122]. Indeed, p16INK4A expression, which is considered the high-risk HPV types are consistently associated with
a marker of elevated E7 expression, can be detected cervical cancers because of the specific activity of their
in some CIN1 lesions, as well as in CIN2 and CIN3 oncogenes. HPV16 gene expression facilitates integration
lesions that show evidence of integration [123,124]. of foreign DNA into the host cell chromosome, which
Integration of the HPV genome into the host cell may be a consequence of the increased level of host


C 2006 The Biochemical Society
Papillomavirus molecular biology 535

genome instability in these cells [141]. The high-risk E7 activation is not known, it is clear that such an ac-
proteins, but not the E7 proteins encoded by the low- tivity may predispose to long-term infection and the
risk HPV types, induce centrosomal abnormalities in cell development of cancer. Interestingly, recent studies have
culture and in transgenic animals [142,143], and it has shown that the viral oncogenes E6 and E7 can antagonize
been suggested that high-risk E7 may act as a mitotic BRCA-mediated inhibition of the hTERT promoter
mutator, which acts to increase the chance of errors [156].
during each round of cell division [144]. Although the Although aberrant expression of high-risk oncogenes
molecular basis for E7-mediated genome instability is not can predispose to the development of cervical cancer,
fully understood, it appears in part to be independent of their expression alone is not considered sufficient, and
the well-characterized association of E7 with the pocket the viral proteins cannot fully transform human keratino-
proteins Rb, p107 and p130. The association of E7 with cytes in culture [157]. It is generally accepted that papi-
these proteins does, however, contribute to the ability of llomavirus-mediated oncogenesis requires the accumu-
E7 to stimulate cell proliferation, with the high-risk E7 lation of additional genetic changes that occur over time
proteins binding Rb more efficiently than the E7 protein following initial infection. The average age of women
of the low-risk HPV types [145,146]. The high-risk E7 with invasive cervical cancer is approx. 50 years, whereas
proteins are also capable of mediating Rb degradation the mean age of women with HSIL is approx. 28 years,
through a proteosome-dependent mechanism [147,148], which suggests in most cases a long precancerous
which is important for E7-mediated cell transformation state that allows the accumulation of secondary genetic
[149]. As with E7, the E6 protein also differs in its changes. Although secondary genetic changes may occur
function between high- and low-risk HPV types. One randomly, the presence of tobacco metabolites in cervical
of the most important of these with regard to cancer secretions is considered a risk factor in the develop-
progression is the ability of high-risk E6 proteins to form ment of cervical cancer [158], with certain smoking car-
a tripartite complex with p53 and the cellular ubiquitin cinogens being found at significantly higher levels in the
ligase E6AP (E6-associated protein), which leads to cervical secretions of cigarette-smoking women [159].
proteosome-mediated p53 degradation [150]. Low-risk Multiparity and the long-term use of oral contraceptives
E6 proteins bind p53 with a lower affinity than the high- are also associated with increased risk [160,161].
risk types and have no significant ability to bind E6AP
and to stimulate p53 degradation [150,151]. The loss of the
p53-mediated DNA damage response in cells expressing REGRESSION, LATENCY AND
high-risk HPV E6 predisposes to the accumulation of ASYMPTOMATIC INFECTION
secondary changes in the host cell chromosome that
eventually lead to cancer. The high-risk E6 proteins also In contrast with our understanding of HPV-associated
differ from those encoded by the low-risk HPV types in cancers, our knowledge of the molecular events that
having a C-terminal PDZ-binding domain. High-risk E6 regulate the development of latent and/or asymptomatic
proteins bind and stimulate the degradation of several infections is only poorly developed. Experimental infec-
cellular targets that contain PDZ motifs, such as hDlg tions using animal papillomaviruses such as ROPV or
(human homologue of Dlg) and hSrib (human homologue COPV lead to the development of lesions that persist
of the Drosophila scribble protein), which are thought for months rather than years [38,162]. Lymphocyte
to be involved in the regulation of cell growth and infiltration and lesion regression take place between 8
attachment [152]. PDZ binding is a conserved feature of and 12 weeks post-infection and, by 16 weeks, infected
all high-risk E6 proteins, and it has been suggested that sites regain the appearance of uninfected epithelium [162].
the loss of cell–cell contacts mediated by tight junctions A similar pattern of events occurs in cattle [163], and
may contribute to the loss of cell polarity seen in HPV- regression of HPV-induced lesions in humans is also
associated cervical cancers [153]. PDZ binding, which is thought to follow this path [164]. The immune system
distinct from the ability of E6 to bind and degrade p53, is clearly important in controlling viral persistence, and
is important for cell transformation [154] and has been patients with immune defects can develop widespread
shown to be necessary for the stimulation of epithelial lesions that are refractory to treatment. Although the
hyperplasia in transgenic animals [53]. Another important immune responses involved in clearance of HPV in-
function for high-risk E6 is its ability to activate the fections are considered in depth elsewhere [165,166],
catalytic subunit of telomerase [hTERT (human telo- it is worth noting that Alpha papillomaviruses (and
merase reverse transcriptase)], which adds hexamer in particular those associated with cervical cancer) can
repeats to the telomeric ends of chromosomes [155]. cause persistent infections and have evolved a number
Telomerase activity is usually absent in somatic cells, of mechanisms to limit the chance of detection by the
leading to the shortening of telomeres with successive cell immune system. Among these are the regulation of E-
divisions and, eventually, to cell senescence. Although cadherin expression and Langerhans cell density by E6
the precise mechanism by which E6 mediates hTERT [167,168], the interference with MHC presentation by


C 2006 The Biochemical Society
536 J. Doorbar

E5 [167] and interference with the function of inter- developed, little is known as to the influence of the
feron response factor 3 by E7 [169]. Perhaps equally infected cell type and the consequence of different cellular
important, however, is the fact that papillomaviruses environments on viral gene expression. The factors that
do not cause a lytic infection, with infectious particles regulate viral persistence and the events that lead to
being produced only in the upper epithelial layers in latency are other areas that are only poorly understood,
cells that are eventually lost from the epithelial surface but which are very important in fully understanding the
at the end of their life span. The viral early proteins consequences of infection in humans. Such topics are
that mediate cell proliferation in the lower epithelial difficult to research, but will need to be addressed if our
layers are thought to be expressed at levels below those understanding of papillomavirus molecular biology is to
required to reliably trigger an effective host immune be advanced further.
response. When it occurs, the stimulation of a cell-medi-
ated immune response that can clear infection appears
to depend on cross-priming of dendritic cells by viral ACKNOWLEDGMENTS
antigens expressed in keratinocytes. The frequent de-
J. D. is a programme leader in the Division of Virology
tection of HPV DNA in cervical lesions in the absence
at the MRC National Institute for Medical Research and
of any obvious disease may represent latent infection in
is supported by the U.K. Medical Research Council. The
which viral genomes are maintained in the basal layer
leadership and vision provided by the present Director,
in the absence of detectable productive infection. Immune
Sir John Skehel FRS, is gratefully acknowledged, as
surveillance is thought to be important for this, as lesions
is the continued support and advice provided by
proliferate following immune supression, but this may
Dr Jonathan Stoye, Head of Virology. The commitment
not be the only way in which viral latency is maintained.
and enthusiasm of all members of the Papillomavirus
Recent studies have suggested that the generation of
laboratory in carrying out their work and in contributing
an asymptomatic infection in which viral DNA can
to our understanding of papillomavirus biology is greatly
be detected in the absence of any abnormal pathology
appreciated.
may be a normal consequence of infection under some
circumstances, and may result from the silencing of viral
gene expression by methylation [139]. Latent infection is REFERENCES
thought to require the expression of the viral E1 and E2
proteins necessary for genome maintenance in the basal 1 Bernard, H. U. (2005) The clinical importance of the
layer, with the E6 and E7 genes not being required [40]. nomenclature, evolution and taxonomy of human
papillomaviruses. J. Clin. Virol. 32 (Suppl. 1), S1–S6
2 Villiers de, E. M. (2001) Taxonomic classification of
papillomaviruses. Papillomavirus Rep. 12, 57–63
3 Harwood, C. A. and Proby, C. M. (2002) Human
CONCLUSIONS papillomaviruses and non-melanoma skin cancer.
Curr. Opin. Infect. Dis. 15, 101–114
The consequence of papillomavirus infection depends 4 Pfister, H. (2003) Human papillomavirus and skin cancer.
J. Natl. Cancer. Inst. Monogr. 31, 52–56
on the infecting HPV type and site of infection, as 5 Ramoz, N., Rueda, L. A., Bouadjar, B., Montoya, L. S.,
well as on host factors that regulate virus persistence, Orth, G. and Favre, M. (2002) Mutations in two adjacent
novel genes are associated with epidermodysplasia
regression and latency. The characteristics of different verruciformis. Nat. Genet. 32, 579–581
HPV types have been studied extensively and it is now 6 Tate, G., Suzuki, T., Kishimoto, K. and Mitsuya, T. (2004)
well known that high-risk types, such as HPV16, en- Novel mutations of EVER1/TMC6 gene in a Japanese
patient with epidermodysplasia verruciformis.
code genes that can contribute to cancer progression J. Hum. Genet. 49, 223–225
when aberrantly expressed. During productive infection, 7 de Villiers, E. M., Fauquet, C., Broker, T. R., Bernard,
H. U. and zur Hausen, H. (2004) Classification of
however, these genes are carefully regulated and play papillomaviruses. Virology 324, 17–27
important roles in virus synthesis and in avoiding 8 Persson, G., Andersson, K. and Krantz, I. (1996)
Symptomatic genital papillomavirus infection in a
detection by the host immune system. It seems that community. Incidence and clinical picture. Acta Obstet.
papillomaviruses, like many other DNA tumour viruses, Gynecol. Scand. 75, 287–290
cause cancers when their regulated pattern of gene 9 Shamanin, V., Glover, M., Rausch, C. et al. (1994) Specific
types of human papillomavirus found in benign
expression is disturbed. Viral oncoproteins such as E6 proliferations and carcinomas of the skin in
and E7, which are involved in cell proliferation and immunosuppressed patients. Cancer Res. 54, 4610–4613
10 Pfister, H. (1992) Human papillomaviruses and skin
the regulation of cell death, can be extremely dangerous cancer. Semin. Cancer Biol. 3, 263–271
when inappropriately expressed. This can happen at the 11 Modis, Y., Trus, B. L. and Harrison, S. C. (2002) Atomic
cervical transformation zone, where most cervical cancers model of the papillomavirus capsid. EMBO J. 21,
4754–4762
originate, and it has been suggested that this may be 12 Trus, B. L., Roden, R. B., Greenstone, H. L., Vrhel, M.,
an unstable site for productive infection by high-risk Schiller, J. T. and Booy, F. P. (1997) Novel structural
features of bovine papillomavirus capsid revealed by a
HPVs. Although our understanding of HPV-associated three dimensional reconstruction to 9A resolution.
cancer progression and productive infection is now well Nat. Struct. Biol. 4, 413–420


C 2006 The Biochemical Society
Papillomavirus molecular biology 537

13 Patterson, N. A., Smith, J. L. and Ozbun, M. A. (2005) 32 Bechtold, V., Beard, P. and Raj, K. (2003) Human
Human papillomavirus type 31b infection of human papillomavirus type 16 E2 protein has no effect on
keratinocytes does not require heparan sulfate. J. Virol. transcription from episomal viral DNA. J. Virol. 77,
79, 6838–6847 2021–2028
14 Shafti-Keramat, S., Handisurya, A., Kriehuber, E., 33 Kim, K., Garner-Hamrick, P. A., Fisher, C., Lee, D.
Meneguzzi, G., Slupetzky, K. and Kirnbauer, R. (2003) and Lambert, P. F. (2003) Methylation patterns of
Different heparan sulfate proteoglycans serve as cellular papillomavirus DNA, its influence on E2 function
receptors for human papillomaviruses. J. Virol. 77, and implications in viral infection. J. Virol. 77,
13125–13135 12450–12459
15 Joyce, J. G., Tung, J. S., Przysiecki, C. T. et al. (1999) The 34 Stanley, M. A., Browne, H. M., Appleby, M. and Minson,
L1 major capsid protein of human papillomavirus type 11 H. C. (1989) Properties of a nontumorigenic human
recombinant virus-like particles interacts with heparin cervical keratinocyte cell line. Int. J. Cancer 43, 672–676
and cell-surface glycosaminoglycans on human 35 De Geest, K., Turyk, M. E., Hosken, M. I., Hudson, J. B.,
keratinocytes. J. Biol. Chem. 274, 5810–5822 Laimins, L. A. and Wilbanks, G. D. (1993) Growth and
16 Chung, C. S., Hsiao, J. C., Chang, Y. S. and Chang, W. differentiation of human papillomavirus type 31b positive
(1998) A27L protein mediates vaccinia virus interaction human cervical cell lines. Gynecol. Oncol. 49, 303–310
with cell surface heparan sulfate. J. Virol. 72, 1577–1585 36 Stubenrauch, F., Lim, H. B. and Laimins, L. A. (1998)
17 Summerford, C., Bartlett, J. S. and Samulski, R. J. (1999) Differential requirements for conserved E2 binding sites
αVβ5 integrin: a co-receptor for adeno-associated virus in the life cycle of oncogenic human papillomavirus
type 2 infection. Nat. Med. 5, 78–82 type 31. J. Virol. 72, 1071–1077
18 McMillan, N. A., Payne, E., Frazer, I. H. and Evander, M. 37 Frattini, M. G., Lim, H. B. and Laimins, L. A. (1996)
(1999) Expression of the α6 integrin confers In-vitro synthesis of oncogenic human papillomaviruses
papillomavirus binding upon receptor-negative B-cells. requires episomal genomes for differentiation-dependent
Virology 261, 271–279 late gene expression. Proc .Natl. Acad. Sci. U.S.A. 93,
19 Evander, M., Frazer, I. H., Payne, E., Mei Qi, Y., 3062–3067
Hengst, K. and McMillan, N. A. J. (1997) Identification 38 Christensen, N. D., Cladel, N. M., Reed, C. A. and
of the α6 integrin as a candidate receptor for Han, R. (2000) Rabbit oral papillomavirus complete
papillomaviruses. J. Virol. 71, 2449–2456 genome sequence and immunity following genital
20 Bossis, I., Roden, R. B., Gambhira, R. et al. (2005) infection. Virology 269, 451–461
Interaction of tSNARE syntaxin 18 with the 39 Nicholls, P., Klaunberg, B., Moore, R. A. et al. (1999)
papillomavirus minor capsid protein mediates infection. Naturally occurring, nonregressing canine oral
J. Virol. 79, 6723–6731 papillomavirus infection: host immunity, virus
21 Culp, T. D. and Christensen, N. D. (2004) Kinetics of in characterization, and experimental infection.
vitro adsorption and entry of papillomavirus virions. Virology 265, 365–374
Virology 319, 152–161 40 Zhang, P., Nouri, M., Brandsma, J. L., Iftner, T. and
22 Day, P. M., Lowy, D. R. and Schiller, J. T. (2003) Steinberg, B. M. (1999) Induction of E6/E7 expression in
Papillomaviruses infect cells via a clathrin-dependent cottontail rabbit papillomavirus latency following UV
pathway. Virology 307, 1–11 activation. Virology 263, 388–394
23 Bousarghin, L., Touze, A., Sizaret, P. Y. and Coursaget, P. 41 Campo, M. S. (1995) Infection by bovine papillomavirus
(2003) Human papillomavirus types 16, 31, and 58 use and prospects for vaccination. Trends Microbiol. 3,
different endocytosis pathways to enter cells. J. Virol. 77, 92–97
3846–3850 42 Madison, K. C. (2003) Barrier function of the skin: ‘la
24 Day, P. M., Baker, C. C., Lowy, D. R. and Schiller, J. T. raison d’etre’ of the epidermis. J. Invest. Dermatol. 121,
(2004) Establishment of papillomavirus infection is 231–241
enhanced by promyelocytic leukemia protein (PML) 43 Sherman, L., Jackman, A., Itzhaki, H., Stoppler, M. C.,
expression. Proc. Natl. Acad. Sci. U.S.A. 101, Koval, D. and Schlegel, R. (1997) Inhibition of serum-
14252–14257 and calcium-induced differentiation of human
25 Dell, G., Wilkinson, K. W., Tranter, R., Parish, J., keratinocytes by HPV16 E6 oncoprotein: role of p53
Leo Brady, R. and Gaston, K. (2003) Comparison of the inactivation. Virology 237, 296–306
structure and DNA-binding properties of the E2 proteins 44 Brehm, A., Nielsen, S. J., Miska, E. A. et al. (1999) The E7
from an oncogenic and a non-oncogenic human oncoprotein associates with Mi2 and histone deacetylase
papillomavirus. J. Mol. Biol. 334, 979–991 activity to promote cell growth. EMBO J. 18, 2449–2458
26 Loo, Y. M. and Melendy, T. (2004) Recruitment of 45 Antinore, M. J., Birrer, M. J., Patel, D., Nader, L. and
replication protein A by the papillomavirus E1 protein McCance, D. J. (1996) The human papillomavirus type 16
and modulation by single-stranded DNA. J. Virol. 78, E7 gene product interacts with and trans-activates the
1605–1615 AP1 family of transcription factors. EMBO J. 15,
27 Masterson, P. J., Stanley, M. A., Lewis, A. P. and Romanos, 1950–1960
M. A. (1998) A C-terminal helicase domain of the human 46 Funk, J. O., Waga, S., Harry, J. B., Espling, E., Stillman, B.
papillomavirus E1 protein binds E2 and the DNA and Galloway, D. A. (1997) Inhibition of CDK
polymerase α-primase p68 subunit. J. Virol. 72, 7407–7419 activity and PCNA-dependent DNA replication by p21 is
28 Conger, K. L., Liu, J. S., Kuo, S. R., Chow, L. T. and blocked by interaction with the HPV16 E7 oncoprotein.
Wang, T. S. (1999) Human papillomavirus DNA Genes Dev. 11, 2090–2100
replication. Interactions between the viral E1 protein and 47 Noya, F., Chien, W.-M., Broker, T. R. and Chow, L. T.
two subunits of human DNA polymerase α/primase. (2001) p21cip1 degradation in differentiated keratinocytes
J. Biol. Chem. 274, 2696–2705 is abrogated by costabilization with cyclin E induced by
29 Han, Y., Loo, Y. M., Militello, K. T. and Melendy, T. human papillomavirus E7. J. Virol. 75, 6121–6134
(1999) Interactions of the papovavirus DNA replication 48 Stacey, S. N., Jordan, D., Williamson, A. J., Brown, M.,
initiator proteins, bovine papillomavirus type 1 E1 and Coote, J. H. and Arrand, J. R. (2000) Leaky scanning is
simian virus 40 large T antigen, with human replication the predominant mechanism for translation of human
protein A. J. Virol. 73, 4899–4907 papillomavirus type 16 E7 oncoprotein from E6/E7
30 You, J., Croyle, J. L., Nishimura, A., Ozato, K. and bicistronic mRNA. J Virol. 74, 7284–7297
Howley, P. M. (2004) Interaction of the bovine 49 Thomas, M. and Banks, L. (1998) Inhibition of
papillomavirus E2 protein with Brd4 tethers the viral Bak-induced apoptosis by HPV-18 E6. Oncogene 17,
DNA to host mitotic chromosomes. Cell 117, 349–360 2943–2954
31 McPhillips, M. G., Ozato, K. and McBride, A. A. (2005) 50 Li, B. and Dou, Q. P. (2000) Bax degradation by the
Interaction of bovine papillomavirus E2 protein with ubiquitin/proteasome-dependent pathway: involvement
Brd4 stabilizes its association with chromatin. J. Virol. 79, in tumor survival and progression. Proc. Natl.
8920–8932 Acad. Sci. U.S.A. 97, 3850–3855


C 2006 The Biochemical Society
538 J. Doorbar

51 Thomas, M., Laura, R., Hepner, K. et al. (2002) 69 Peh, W. L., Brandsma, J. L., Christensen, N. D., Cladel,
Oncogenic human papillomavirus E6 proteins target the N. M., Wu, X. and Doorbar, J. (2004) The Viral E4
MAGI-2 and MAGI-3 proteins for degradation. protein is required for the completion of the cottontail
Oncogene 21, 5088–5096 rabbit papillomavirus (CRPV) productive cycle in vivo.
52 Nguyen, M. L., Nguyen, M. M., Lee, D., Griep, A. E. and J. Virol. 15, 2142–2151
Lambert, P. F. (2003) The PDZ ligand domain of the 70 Hwang, E. S., Nottoli, T. and Dimaio, D. (1995) The
human papillomavirus type 16 E6 protein is required for HPV16 E5 protein: expression, detection and stable
E6’s induction of epithelial hyperplasia in vivo. J. Virol. complex formation with transmembrane proteins in COS
77, 6957–6964 cells. Virology 211, 227–233
53 Nguyen, M. M., Nguyen, M. L., Caruana, G., 71 Disbrow, G. L., Hanover, J. A. and Schlegel, R. (2005)
Bernstein, A., Lambert, P. F. and Griep, A. E. (2003) Endoplasmic reticulum-localized human papillomavirus
Requirement of PDZ-containing proteins for cell cycle type 16 E5 protein alters endosomal pH but not
trans-Golgi pH. J. Virol. 79, 5839–5846
regulation and differentiation in the mouse lens
72 Straight, S. W., Herman, B. and McCance, D. J. (1995)
epithelium. Mol. Cell. Biol. 23, 8970–8981
The E5 oncoprotein of HPV16 inhibits the acidification
54 Peh, W. L., Middleton, K., Christensen, N. et al. (2002) of endosomes in human keratinocytes. J. Virol. 69,
Life cycle heterogeneity in animal models of human 3185–3192
papillomavirus-associated disease. J. Virol. 76, 73 Crusius, K., Rodriguez, I. and Alonso, A. (2000) The
10401–10416 human papillomavirus type 16 E5 protein modulates
55 Bodily, J. M. and Meyers, C. (2005) Genetic analysis of the ERK1/2 and p38 MAP kinase activation by an
human papillomavirus type 31 differentiation-dependent EGFR-independent process in stressed human
late promoter. J. Virol. 79, 3309–3321 keratinocytes. Virus Genes 20, 65–69
56 Spink, K. M. and Laimins, L. A. (2005) Induction of the 74 Davy, C. E., Jackson, D. J., Raj, K. et al. (2005) Human
human papillomavirus type 31 late promoter requires Papillomavirus type 16 E1ˆE4-induced G2 arrest is
differentiation but not DNA amplification. J. Virol. 79, associated with cytoplasmic retention of active
4918–4926 Cdk1/cyclin B1 complexes. J. Virol. 79, 3998–4011
57 Middleton, K., Peh, W., Southern, S. A. et al. (2003) 75 Davy, C. E., Jackson, D. J., Wang, Q. et al. (2002)
Organisation of the human papillomavirus productive Identification of a G2 arrest domain in the E1ˆE4 protein
cycle during neoplastic progression provides a basis for of human papillomavirus type 16. J. Virol. 76, 9806–9818
the selection of diagnostic markers. J. Virol. 77, 76 Nakahara, T., Nishimura, A., Tanaka, M., Ueno, T.,
10186–10201 Ishimoto, A. and Sakai, H. (2002) Modulation of the cell
58 Sarafi, T. R. and McBride, A. A. (1995) Domains of the division cycle by human papillomavirus type 18 E4.
BPV-1 E1 replication protein required for origin-specific J. Virol. 76, 10914–10920
DNA binding and interaction with the E2 transactivator. 77 Knight, G. L., Grainger, J. R., Gallimore, P. H. and
Virology 211, 385–396 Roberts, S. (2004) Cooperation between different forms
59 Moscufo, N., Sverdrup, F., Breiding, D. E. and Androphy, of the human papillomavirus type 1 E4 protein to block
E. J. (1999) Two distinct regions of the BPV1 E1 cell cycle progression and cellular DNA synthesis.
replication protein interact with the activation domain of J. Virol. 78, 13920–13933
78 Sorathia, R., Davy, C., Sterlinko, H. et al. (2004) The E4
E2. Virus Res. 65, 141–154
protein of HPV16 Interacts with E2 and regulates the
60 Titolo, S., Pelletier, A., Sauve, F. et al. (1999) Role of the
trans-activation and replication functions of E2. 21st
ATP-binding domain of the human papillomavirus International Papillomavirus Conference, Mexico City,
type 11 E1 helicase in E2-dependent binding to the origin. Mexico, p318
J. Virol. 73, 5282–5293 79 Florin, L., Sapp, C., Streeck, R. E. and Sapp, M. (2002)
61 Lee, D., Sohn, H., Kalpana, G. V. and Choe, J. (1999) Assembly and translocation of papillomavirus capsid
Interaction of E1 and hSNF5 proteins stimulates proteins. J. Virol. 76, 10009–10014
replication of human papillomavirus DNA. Nature 80 Doorbar, J. and Gallimore, P. H. (1987) Identification of
(London) 399, 487–491 proteins encoded by the L1 and L2 open reading frames
62 Hines, C. S., Meghoo, C., Shetty, S., Biburger, M., of human papillomavirus type 1a. J. Virol. 61, 2793–2799
Brenowitz, M. and Hegde, R. S. (1998) DNA structure 81 Schwartz, S. (2000) Regulation of human papillomavirus
and flexibility in the sequence-specific binding of late gene expression. Ups. J. Med. Sci. 105, 171–192
papillomavirus E2 proteins. J. Mol. Biol. 276, 82 Zhou, J., Liu, W. J., Peng, S. W., Sun, X. Y. and Frazer, I.
809–818 (1999) Papillomavirus capsid protein expression level
63 Demeret, C., Goyat, S., Yaniv, M. and Thierry, F. (1998) depends on the match between codon usage and tRNA
The human papillomavirus type 18 (HPV18) replication availability. J. Virol. 73, 4972–4982
protein E1 is a transcriptional activator when interacting 83 Zhao, K. N., Liu, W. J. and Frazer, I. H. (2003) Codon
with HPV18 E2. Virology 242, 378–386 usage bias and A + T content variation in human
64 Steger, G. and Corbach, S. (1997) Dose-dependent papillomavirus genomes. Virus Res. 98, 95–104
regulation of the early promoter of human 84 McPhillips, M. G., Veerapraditsin, T., Cumming, S. A.
papillomavirus type 18 by the viral E2 protein. et al. (2004) SF2/ASF binds the human papillomavirus
J. Virol. 71, 50–58 type 16 late RNA control element and is regulated during
65 Fehrmann, F., Klumpp, D. J. and Laimins, L. A. (2003) differentiation of virus-infected epithelial cells. J. Virol.
Human papillomavirus type 31 E5 protein supports cell 78, 10598–10605
cycle progression and activates late viral functions upon 85 Cumming, S. A., McPhillips, M. G., Veerapraditsin, T.,
Milligan, S. G. and Graham, S. V. (2003) Activity of the
epithelial differentiation. J. Virol. 77, 2819–2831
human papillomavirus type 16 late negative regulatory
66 Genther, S. M., Sterling, S., Duensing, S., Munger, K.,
element is partly due to four weak consensus 5 splice
Sattler, C. and Lambert, P. F. (2003) Quantitative role of sites that bind a U1 snRNP-like complex. J. Virol. 77,
the human papillomavirus type 16 E5 gene during the 5167–5177
productive stage of the viral life cycle. J. Virol. 77, 86 Zhao, X., Rush, M. and Schwartz, S. (2004) Identification
2832–2842 of an hnRNP A1-dependent splicing silencer in the
67 Wilson, R., Fehrmann, F. and Laimins, L. A. (2005) Role human papillomavirus type 16 L1 coding region that
of the E1–E4 protein in the differentiation-dependent life prevents premature expression of the late L1 gene.
cycle of human papillomavirus type 31. J. Virol. 79, J. Virol. 78, 10888–10905
6732–6740 87 Leder, C., Kleinschmidt, J. A., Wiethe, C. and Muller, M.
68 Nakahara, T., Peh, W. L., Doorbar, J., Lee, D. and (2001) Enhancement of capsid gene expression: preparing
Lambert, P. F. (2005) Human papillomavirus type 16 the human papillomavirus type 16 major structural gene
E1ˆE4 contributes to multiple facets of the papillomavirus L1 for DNA vaccination purposes. J. Virol. 75,
life cycle. J. Virol. 79, 13150–13165 9201–9209


C 2006 The Biochemical Society
Papillomavirus molecular biology 539

88 Zhao, K. N., Gu, W., Fang, N. X., Saunders, N. A. and 107 Lehr, E., Hohl, D., Huber, M. and Brown, D. (2004)
Frazer, I. H. (2005) Gene codon composition determines Infection with Human Papillomavirus alters expression of
differentiation-dependent expression of a viral capsid the small proline rich proteins 2 and 3. J. Med. Virol. 72,
gene in keratinocytes in vitro and in vivo. Mol. Cell. Biol. 478–483
25, 8643–8655 108 Lorincz, A. T., Reid, R., Jenson, A. B., Greenberg, M. D.,
89 Tan, W. and Schwartz, S. (1995) The Rev protein of human Lancaster, W. D. and Kurman, R. J. (1992) Human
immunodeficiency virus type 1 counteracts the effect of papillomavirus infection of the cervix: relative risk
an AU-rich negative element in the human papillomavirus associations of 15 common anogenital types.
type 1 late 3 untranslated region. J Virol. 69, 2932–2945 Obstet. Gynecol. 79, 328–337
90 Cumming, S. A., Repellin, C. E., McPhillips, M., Radford, 109 Clifford, G. M., Gallus, S., Herrero, R. et al. (2005)
J. C., Clements, J. B. and Graham, S. V. (2002) The human Worldwide distribution of human papillomavirus types in
papillomavirus type 31 late 3 untranslated region cytologically normal women in the International Agency
contains a complex bipartite negative regulatory element. for Research on Cancer HPV prevalence surveys:
J Virol. 76, 5993–6003 a pooled analysis. Lancet 366, 991–998
91 Furth, P. A., Choe, W. T., Rex, J. H., Byrne, J. C. and 110 Woodman, C. B., Collins, S., Winter, H. et al. (2001)
Baker, C. C. (1994) Sequences homologous to 5 splice Natural history of cervical human papillomavirus
sites are required for the inhibitory activity of infection in young women: a longitudinal cohort study.
papillomavirus late 3 untranslated regions. Lancet 357, 1831–1836
Mol. Cell. Biol. 14, 5278–5289 111 Richardson, H., Kelsall, G., Tellier, P. et al. (2003) The
92 Day, P. M., Roden, R. B. S., Lowy, D. R. and Schiller, J. T. natural history of type-specific human papillomavirus
(1998) The papillomavirus minor capsid protein, L2, infections in female university students.
induces localization of the major capsid protein, L1 and Cancer Epidemiol. Biomarkers Prev. 12, 485–490
the viral transcription/replication protein, E2, to PML 112 Clifford, G. M., Rana, R. K., Franceschi, S., Smith, J. S.,
oncogenic domains. J. Virol. 72, 142–150 Gough, G. and Pimenta, J. M. (2005) Human
93 Zhao, K. N., Hengst, K., Liu, W. J. et al. (2000) BPV1 E2 papillomavirus genotype distribution in low-grade
protein enhances packaging of full-length plasmid DNA cervical lesions: comparison by geographic region and
in BPV1 pseudovirions. Virology 272, 382–393 with cervical cancer. Cancer Epidemiol. Biomarkers Prev.
94 Buck, C. B., Pastrana, D. V., Lowy, D. R. and Schiller, J. T. 14, 1157–1164
(2004) Efficient intracellular assembly of papillomaviral 113 Clifford, G. M., Smith, J. S., Plummer, M., Munoz, N. and
vectors. J. Virol. 78, 751–757 Franceschi, S. (2003) Human papillomavirus types in
95 Fay, A., Yutzy, W. H. T., Roden, R. B. and Moroianu, J. invasive cervical cancer worldwide: a meta-analysis.
(2004) The positively charged termini of L2 minor capsid Br. J. Cancer. 88, 63–73
protein required for bovine papillomavirus infection 114 Cuschieri, K. S., Cubie, H. A., Whitley, M. W. et al. (2004)
function separately in nuclear import and DNA binding. Multiple high risk HPV infections are common in cervical
J. Virol. 78, 13447–13454 neoplasia and young women in a cervical screening
96 Florin, L., Becker, K. A., Sapp, C. et al. (2004) Nuclear population. J. Clin. Pathol. 57, 68–72
translocation of papillomavirus minor capsid protein L2 115 Xi, L. F., Demers, G. W., Koutsky, L. A. et al. (1995)
requires Hsc70. J. Virol. 78, 5546–5553 Analysis of human papillomavirus type 16 variants
97 Lin, B. Y., Makhov, A. M., Griffith, J. D., Broker, T. R. indicates establishment of persistent infection.
and Chow, L. T. (2002) Chaperone proteins abrogate J. Infect. Dis. 172, 747–755
inhibition of the human papillomavirus (HPV) E1 116 Koutsky, L. A., Holmes, K. K., Critchlow, C. W. et al.
replicative helicase by the HPV E2 protein. (1992) A cohort study of the risk of cervical intraepithelial
Mol. Cell. Biol. 22, 6592–6604 neoplasia grade 2 or 3 in relation to papillomavirus
98 Becker, K. A., Florin, L., Sapp, C., Maul, G. G. and infection. N. Engl. J. Med. 327, 1272–1278
Sapp, M. (2004) Nuclear localization but not PML 117 Parkin, D. M., Bray, F., Ferlay, J. and Pisani, P. (2005)
protein is required for incorporation of the Global cancer statistics, 2002. CA Cancer J. Clin. 55,
papillomavirus minor capsid protein L2 into virus-like 74–108
particles. J. Virol. 78, 1121–1128 118 Peto, J., Gilham, C., Fletcher, O. and Matthews, F. E.
99 Stauffer, Y., Raj, K., Masternak, K. and Beard, P. (1998) (2004) The cervical cancer epidemic that screening has
Infectious human papillomavirus type 18 pseudovirions. prevented in the UK. Lancet 364, 249–256
J. Mol. Biol. 283, 529–536 119 Bosch, F. X. and de Sanjose, S. (2003) Human
100 Roden, R. B., Day, P. M., Bronzo, B. K. et al. (2001) papillomavirus and cervical cancer: burden and assessment
Positively charged termini of the L2 minor capsid protein of causality. J. Natl. Cancer Inst. Monogr. 31, 3–13
are necessary for papillomavirus infection. J. Virol. 75, 120 Klaes, R., Woerner, S. M., Ridder, R. et al. (1999)
10493–10497 Detection of high-risk cervical intraepithelial neoplasia
101 Holmgren, S. C., Patterson, N. A., Ozbun, M. A. and and cervical cancer by amplification of transcripts derived
Lambert, P. F. (2005) The minor capsid protein L2 from integrated papillomavirus oncogenes. Cancer Res.
contributes to two steps in the human papillomavirus 59, 6132–6136
type 31 life cycle. J. Virol. 79, 3938–3948 121 Fujii, T., Masumoto, N., Saito, M. et al. (2005)
102 Finnen, R. L., Erickson, K. D., Chen, X. S. and Garcea, Comparison between in situ hybridization and real-time
R. L. (2003) Interactions between papillomavirus L1 and PCR technique as a means of detecting the integrated
L2 capsid proteins. J. Virol. 77, 4818–4826 form of human papillomavirus 16 in cervical neoplasia.
103 Roden, R. B., Lowy, D. R. and Schiller, J. T. (1997) Diagn. Mol. Pathol. 14, 103–108
Papillomavirus is resistant to desiccation. J. Infect. Dis. 122 Peitsaro, P., Johansson, B. and Syrjanen, S. (2002)
176, 1076–1079 Integrated human papillomavirus type 16 is frequently
104 Bryan, J. T. and Brown, D. R. (2000) Association of the found in cervical cancer precursors as demonstrated by
human papillomavirus type 11 E1()E4 protein with a novel quantitative real-time PCR technique.
cornified cell envelopes derived from infected genital J. Clin. Microbiol. 40, 886–891
epithelium. Virology 277, 262–269 123 Kalof, A. N., Evans, M. F., Simmons-Arnold, L.,
105 Doorbar, J., Ely, S., Sterling, J., McLean, C. and Beatty, B. G. and Cooper, K. (2005) p16INK4A
Crawford, L. (1991) Specific interaction between HPV-16 immunoexpression and HPV in situ hybridization signal
E1-E4 and cytokeratins results in collapse of the epithelial patterns: potential markers of high-grade cervical
cell intermediate filament network. Nature (London) 352, intraepithelial neoplasia. Am. J. Surg. Pathol. 29, 674–679
824–827 124 Dray, M., Russell, P., Dalrymple, C. et al. (2005)
106 Wang, Q., Griffin, H., Southern, S. et al. (2004) Functional p16(INK4a) as a complementary marker of high-grade
Analysis of the human papillomavirus type 16 E1ˆE4 intraepithelial lesions of the uterine cervix. I: Experience
protein provides a mechanism for in vivo and in vitro with squamous lesions in 189 consecutive cervical
keratin filament re-organisation. J. Virol. 78, 821–833 biopsies. Pathology 37, 112–124


C 2006 The Biochemical Society
540 J. Doorbar

125 Yu, T., Ferber, M. J., Cheung, T. H., Chung, T. K., 143 Schaeffer, A. J., Nguyen, M., Liem, A. et al. (2004) E6 and
Wong, Y. F. and Smith, D. I. (2005) The role of viral E7 oncoproteins induce distinct patterns of chromosomal
integration in the development of cervical cancer. aneuploidy in skin tumors from transgenic mice.
Cancer Genet. Cytogenet. 158, 27–34 Cancer Res. 64, 538–546
126 Butz, K., Ristriani, T., Hengstermann, A., Denk, C., 144 Duensing, S. and Munger, K. (2004) Mechanisms of
Scheffner, M. and Hoppe-Seyler, F. (2003) siRNA genomic instability in human cancer: insights from studies
targeting of the viral E6 oncogene efficiently kills human with human papillomavirus oncoproteins. Int. J. Cancer
papillomavirus-positive cancer cells. Oncogene 22, 109, 157–162
5938–5945 145 Gage, J. R., Meyers, C. and Wettstein, F. O. (1990) The E7
127 Jiang, M. and Milner, J. (2002) Selective silencing of viral proteins of the nononcogenic human papillomavirus
gene expression in HPV-positive human cervical type 6b (HPV-6b) and of the oncogenic HPV-16 differ in
carcinoma cells treated with siRNA, a primer of RNA retinoblastoma protein binding and other properties.
interference. Oncogene 21, 6041–6048 J. Virol. 64, 723–730
128 Griffin, H., Elston, R., Jackson, D. et al. (2006) Inhibition 146 Munger, K., Phelps, W. C., Bubb, V., Howley, P. M. and
of papillomavirus protein function in cervical cancer cells Schlegel, R. (1989) The E6 and E7 genes of the human
by intrabody targeting. J. Mol. Biol. 355, 360–378 papillomavirus type 16 together are necessary and
129 Goodwin, E. C. and DiMaio, D. (2000) Repression of sufficient for transformation of primary human
human papillomavirus oncogenes in HeLa cervical keratinocytes. J. Virol. 63, 4417–4421
carcinoma cells causes the orderly reactivation of dormant 147 Berezutskaya, E., Yu, B., Morozov, A., Raychaudhuri, P.
tumor suppressor pathways. Proc. Natl. Acad. Sci. U.S.A. and Bagchi, S. (1997) Differential regulation of the pocket
97, 12513–12518 domains of the retinoblastoma family proteins by the
130 Nishimura, A., Ono, T., Ishimoto, A. et al. (2000) HPV16 E7 oncoprotein. Cell Growth Differ. 8,
Mechanisms of human papillomavirus E2-mediated 1277–1286
repression of viral oncogene expression and cervical 148 Boyer, S. N., Wazer, D. E. and Band, V. (1996) E7 protein
cancer cell growth inhibition. J. Virol. 74, 3752–3760 of human papilloma virus-16 induces degradation of
131 Fournier, N., Raj, K., Saudan, P. et al. (1999) Expression retinoblastoma protein through the ubiquitin-proteasome
of human papillomavirus 16 E2 protein in pathway. Cancer Res. 56, 4620–4624
Schizosaccharomyces pombe delays the initiation of 149 Helt, A. M. and Galloway, D. A. (2003) Mechanisms by
mitosis. Oncogene 18, 4015–4021 which DNA tumor virus oncoproteins target the Rb
132 Frattini, M. G., Hurst, S. D., Lim, H. B., Swaminathan, S. family of pocket proteins. Carcinogenesis 24, 159–169
and Laimins, L. (1997) Abrogation of a mitotic 150 Huibregtse, J. M., Scheffner, M. and Howley, P. M. (1993)
checkpoint by E2 proteins from oncogenic human Localization of the E6AP regions that direct human
papillomaviruses correlates with increased turnover of the papillomavirus E6 binding, association with p53 and
p53 tumour suppressor protein. EMBO J. 16, 318–331 ubiquitination of associated proteins. Mol. Cell. Biol. 13,
133 Vinther, J., Rosenstierne, M. W., Kristiansen, K. and 4918–4927
Norrild, B. (2005) The 3 region of human papillomavirus 151 Storey, A., Thomas, M., Kalita, A. et al. (1998) Role of a
type 16 early mRNAs decrease expression. p53 polymorphism in the development of human
BMC Infect. Dis. 5, 83 papillomavirus-associated cancer. Nature (London) 393,
134 Jeon, S. and Lambert, P. F. (1995) Integration of HPV16 229–234
DNA into the human genome leads to increased stability 152 Zeitler, J., Hsu, C. P., Dionne, H. and Bilder, D. (2004)
of E6/E7 mRNAs:implications for cervical carcinogenesis. Domains controlling cell polarity and proliferation in the
Proc. Natl. Acad. Sci. U.S.A. 92, 1654–1658 Drosophila tumor suppressor Scribble. J. Cell Biol. 167,
135 Jeon, S., Allen-Hoffmann, B. L. and Lambert, P. F. (1995) 1137–1146
Integration of human papillomavirus type 16 into the 153 Nakagawa, S. and Huibregtse, J. M. (2000) Human
human genome correlates with a selective growth scribble (Vartul) is targeted for ubiquitin-mediated
advantage of cells. J. Virol. 69, 2989–2997 degradation by the high-risk papillomavirus E6 proteins
136 Pater, M. M., Hughes, G. A., Hyslop, D. E., Nakshatri, H. and the E6AP ubiquitin-protein ligase. Mol. Cell. Biol.
and Pater, A. (1988) Glucocorticoid-dependent oncogenic 20, 8244–8253
transformation by type 16 but not type 11 human 154 Kiyono, T., Hiraiwa, A., Fujita, M., Hayashi, Y.,
papilloma virus DNA. Nature (London) 335, 832–835 Akiyama, T. and Ishibashi, M. (1997) Binding of high-risk
137 Chan, W. K., Klock, G. and Bernard, H. U. (1989) human papillomavirus E6 oncoproteins to the human
Progesterone and glucocorticoid response elements occur homologue of the Drosophila discs large tumor
in the long control regions of several human suppressor protein. Proc. Natl. Acad. Sci. U.S.A. 94,
papillomaviruses involved in anogenital neoplasia. 11612–11616
J. Virol. 63, 3261–3269 155 Klingelhutz, A. J., Foster, S. A. and McDougall, J. K.
138 Arbeit, J. M., Howley, P. M. and Hanahan, D. (1996) (1996) Telomerase activation by the E6 gene product of
Chronic estrogen-induced cervical and vaginal squamous human papillomavirus type 16. Nature (London) 380,
carcinogenesis in human papillomavirus type 16 79–82
transgenic mice. Proc. Natl. Acad. Sci. U.S.A. 93, 156 Zhang, Y., Fan, S., Meng, Q. et al. (2005) BRCA1
2930–2935 interaction with human papillomavirus oncoproteins.
139 Kalantari, M., Calleja-Macias, I. E., Tewari, D. et al. J. Biol. Chem. 280, 33165–33177
(2004) Conserved methylation patterns of human 157 Hawley-Nelson, P., Vousden, K. H., Hubbert, N. L.,
papillomavirus type 16 DNA in asymptomatic infection Lowy, D. R. and Schiller, J. T. (1989) HPV16 E6 and E7
and cervical neoplasia. J. Virol. 78, 12762–12772 proteins cooperate to immortalize human foreskin
140 Van Tine, B. A., Kappes, J. C., Banerjee, N. S. et al. (2004) keratinocytes. EMBO J. 8, 3905–3910
Clonal selection for transcriptionally active viral 158 Hellberg, D. and Stendahl, U. (2005) The biological role
oncogenes during progression to cancer. J. Virol. 78, of smoking, oral contraceptive use and endogenous
11172–11186 sexual steroid hormones in invasive squamous
141 Duensing, S. and Munger, K. (2002) The human epithelial cervical cancer. Anticancer Res. 25,
papillomavirus type 16 E6 and E7 oncoproteins 3041–3046
independently induce numerical and structural 159 Haverkos, H. W. (2004) Viruses, chemicals and
chromosome instability. Cancer Res. 62, 7075–7082 co-carcinogenesis. Oncogene 23, 6492–6499
142 Duensing, S., Duensing, A., Flores, E. R., Do, A., 160 Castle, P. E. (2004) Beyond human papillomavirus: the
Lambert, P. F. and Munger, K. (2001) Centrosome cervix, exogenous secondary factors and the development
abnormalities and genomic instability by episomal of cervical precancer and cancer. J. Low. Genit. Tract Dis.
expression of human papillomavirus type 16 in raft 8, 224–230
cultures of human keratinocytes. J. Virol. 75, 161 Moodley, J. (2004) Combined oral contraceptives and
7712–7716 cervical cancer. Curr. Opin. Obstet. Gynecol. 16, 27–29


C 2006 The Biochemical Society
Papillomavirus molecular biology 541

162 Nicholls, P. K., Moore, P. F., Anderson, D. M. et al. (2001) 167 Zhang, B., Li, P., Wang, E. et al. (2003) The E5 protein of
Regression of canine oral papillomas is associated with human papillomavirus type 16 perturbs MHC class II
infiltration of CD4+ and CD8+ lymphocytes. Virology antigen maturation in human foreskin keratinocytes
283, 31–39 treated with interferon-gamma. Virology 310,
163 Knowles, G., O’Neil, B. W. and Campo, M. S. (1996) 100–108
Phenotypical characterization of lymphocytes infiltrating 168 Matthews, K., Leong, C. M., Baxter, L. et al. (2003)
regressing papillomas. J. Virol. 70, 8451–8458 Depletion of Langerhans cells in human papillomavirus
164 Coleman, N., Birley, H. D. L., Renton, A. M. et al. (1994) type 16-infected skin is associated with E6-mediated
Immunological events in regressing genital warts. down regulation of E-cadherin. J. Virol. 77, 8378–8385
Am. J. Clin. Pathol. 102, 768–774 169 Barnard, P., Payne, E. and McMillan, N. A. (2000) The
165 Stanley, M. (2005) Immune responses to human human papillomavirus E7 protein is able to inhibit
papillomavirus. Vaccine, doi:10.1016/ the antiviral and anti-growth functions of interferon-α.
j.vaccine.2005.09.002 Virology 277, 411–419
166 Stern, P. L. (2005) Immune control of human 170 Munoz, N., Bosch, F. X., Castellsague, X. et al. (2004)
papillomavirus (HPV) associated anogenital disease and Against which human papillomavirus types shall we
potential for vaccination. J. Clin. Virol. 32 (Suppl. 1), vaccinate and screen? The international perspective.
S72–S81 Int. J. Cancer. 111, 278–285

Received 15 December 2005/24 January 2006; accepted 27 January 2006


Published on the Internet 11 April 2006, doi:10.1042/CS20050369


C 2006 The Biochemical Society

You might also like