ArticlePDF Available

Transcriptional Activity among High and Low Risk Human Papillomavirus E2 Proteins Correlates with E2 DNA Binding

Authors:

Abstract and Figures

The full-length E2 protein, encoded by human papillomaviruses (HPVs), is a sequence-specific transcription factor found in all HPVs, including cancer-causing high risk HPV types 16 and 18 and wart-inducing low risk HPV types 6 and 11. To investigate whether E2 proteins encoded by high risk HPVs may function differentially from E2 proteins encoded by low risk HPVs and animal papillomaviruses, we conducted comparative DNA-binding and transcription studies using electrophoretic mobility shift assays and cell-free transcription systems reconstituted with purified general transcription factors, cofactor, RNA polymerase II, and with E2 proteins encoded by HPV-16, HPV-18, HPV-11, and bovine papillomavirus type 1 (BPV-1). We found that although different types of E2 proteins all exhibited transactivation and repression activities, depending on the sequence context of the E2-binding sites, HPV-16 E2 shows stronger transcription activity and greater DNA-binding affinity than those displayed by the other E2 proteins. Surprisingly, HPV-18 E2 behaves more similarly to BPV-1 E2 than HPV-16 E2 in its functional properties. Our studies thus categorize HPV-18 E2 and BPV-1 E2 in the same protein family, a finding consistent with the available E2 structural data that separate the closely related HPV-16 and HPV-18 E2 proteins but classify together the more divergent BPV-1 and HPV-18 E2 proteins.
Content may be subject to copyright.
Transcriptional Activity among High and Low Risk Human
Papillomavirus E2 Proteins Correlates with E2 DNA Binding*
Received for publication, July 9, 2002, and in revised form, August 28, 2002
Published, JBC Papers in Press, September 17, 2002, DOI 10.1074/jbc.M206829200
Samuel Y. Hou, Shwu-Yuan Wu, and Cheng-Ming Chiang‡
From the Department of Biochemistry, Case Western Reserve University School of Medicine, Cleveland, Ohio 44106-4935
The full-length E2 protein, encoded by human papil-
lomaviruses (HPVs), is a sequence-specific transcription
factor found in all HPVs, including cancer-causing high
risk HPV types 16 and 18 and wart-inducing low risk
HPV types 6 and 11. To investigate whether E2 proteins
encoded by high risk HPVs may function differentially
from E2 proteins encoded by low risk HPVs and animal
papillomaviruses, we conducted comparative DNA-
binding and transcription studies using electrophoretic
mobility shift assays and cell-free transcription systems
reconstituted with purified general transcription fac-
tors, cofactor, RNA polymerase II, and with E2 proteins
encoded by HPV-16, HPV-18, HPV-11, and bovine papil-
lomavirus type 1 (BPV-1). We found that although differ-
ent types of E2 proteins all exhibited transactivation
and repression activities, depending on the sequence
context of the E2-binding sites, HPV-16 E2 shows stron-
ger transcription activity and greater DNA-binding af-
finity than those displayed by the other E2 proteins.
Surprisingly, HPV-18 E2 behaves more similarly to
BPV-1 E2 than HPV-16 E2 in its functional properties.
Our studies thus categorize HPV-18 E2 and BPV-1 E2 in
the same protein family, a finding consistent with the
available E2 structural data that separate the closely
related HPV-16 and HPV-18 E2 proteins but classify to-
gether the more divergent BPV-1 and HPV-18 E2
proteins.
Human papillomaviruses (HPVs)
1
are a family of small DNA
viruses that cause a wide variety of human diseases ranging
from benign epithelial lesions, such as warts, to invasive can-
cers, such as cervical carcinoma. So far, more than 100 HPV
types have been identified and fully sequenced, whereas more
than 120 putative novel types have been partially character-
ized (1, 2). HPV types frequently found in invasive cancers
include HPV-16, -18, -31, -33, -35, -39, -45, -51, -52, -58, and -59.
These are classified as high risk HPVs. In contrast, HPV types
that are rarely found in cancers but are associated with genital
warts, such as HPV-6 and HPV-11, are considered low risk
HPVs (1–3). Because the genomic structures of HPVs are
highly conserved, it is important to determine the functional
differences among individual HPV gene products which lead to
etiologically high and low risk phenotypes. Previous compara-
tive studies have primarily focused on HPV-encoded E6- and
E7-transforming proteins. These studies found that E6 and E7,
from high risk HPVs, lead to cellular transformation much
more readily than low risk E6 and E7 proteins (for review, see
Refs. 4 and 5). In both high and low risk HPVs, expression of E6
and E7 is transcriptionally regulated via the E6 promoter by
many cellular and viral proteins. The full-length viral E2
protein is a sequence-specific transcription factor that func-
tions as an activator or repressor to regulate tightly the E6
promoter through four consensus E2-binding sites (E2-BSs),
ACCGN
4
CGGT (6, 7), whose locations within the upstream
regulatory region (URR) are highly conserved among genital
HPVs. Efficient activation of the E6 promoter requires binding
of E2 protein to the promoter-distal E2-BSs in conjunction with
binding of cellular factors to the enhancer elements also located
within the URR (8 –11). Activation directly mediated through
E2 protein may occur via different mechanisms. First, E2 may
recruit the general transcription machinery to the promoter
through direct interactions with TATA-binding protein (TBP),
TBP-associated factors in TFIID, TFIIB, as well as RNA po-
lymerase II (12–17). Second, E2 interacts with nuclear factors
Sp1, Gps2/AMF-1, or TopBP1, which may then serve as coregu-
lators for transcription (18 –20). Third, E2 may potentiate tran-
scription by remodeling DNA structure at the chromatin level
(21). Indeed, studies have shown E2 interacts with proteins
(CREB-binding protein, p300, and p300/CBP-associated factor)
possessing histone acetyltransferase activity (22–24). It is
likely that different types of E2 proteins function differentially
in regulating preinitiation complex assembly and show varied
cooperativity with cellular enhancer-binding factors that rec-
ognize constitutive, inducible, or cell type-specific enhancer
elements located in the URR, as well as with general or gene-
specific transcription cofactors that modulate E2 activity on
HPV chromatin.
Although low levels of E2 occupy promoter-distal binding
sites, it is postulated that transcriptional repression occurs
with high levels of E2 leading to occupancy at promoter-prox-
imal E2-BSs thus displacing cellular factors critical for E6
promoter activity (9, 10, 25–27). We have shown previously
that E2 protein can also actively repress transcription by pre-
venting preinitiation complex formation (28). This active re-
pression was demonstrated to be independent of cellular en-
hancer-binding factors, Sp1, and general cofactors such as
TBP-associated factors, TFIIA, mediator and PC4 (28). Con-
versely, cellular proteins may bind to DNA and hinder access of
* This work was supported by Grants CA81017 and GM59643 from
the National Institutes of Health and Grant RPG-97-135-04-MBC from
the American Cancer Society. The costs of publication of this article
were defrayed in part by the payment of page charges. This article must
therefore be hereby marked “advertisement” in accordance with 18
U.S.C. Section 1734 solely to indicate this fact.
Mt. Sinai Health Care Foundation scholar. To whom correspond-
ence should be addressed: Dept. of Biochemistry, W-409, Case Western
Reserve University School of Medicine, 10900 Euclid Ave., Cleveland,
OH 44106-4935. Tel.: 216-368-8550; Fax: 216-368-3419; E-mail:
c-chiang@biochemistry.cwru.edu.
1
The abbreviations used are: HPV(s), human papillomavirus(es);
BE2, BPV-1 E2 protein; BPV-1, bovine papillomavirus type 1; CREB,
cAMP-response element-binding protein; CMV, cytomegalovirus; E2-
BS(s), E2-binding site(s); EMSA, electrophoretic mobility shift assay;
HPV-11, HPV type 11; HPV-16, HPV type 16; HPV-18, HPV type 18;
11E2, HPV-11 E2 protein; 16E2, HPV-16 E2 protein; 18E2, HPV-18 E2
protein; PC4, positive cofactor 4; pol, polymerase; TBP, TATA-binding
protein; TF, transcription factor; URR, upstream regulatory region.
THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 277, No. 47, Issue of November 22, pp. 45619–45629, 2002
© 2002 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.
This paper is available on line at http://www.jbc.org 45619
by guest on February 9, 2017http://www.jbc.org/Downloaded from
E2 protein, thus inhibiting E2 function (29 31). Papillomavi-
ruses may also use its own proteins to modulate E2 activation
function. These include various forms of E2 repressor proteins
created through alternative splicing to generate gene products
that contain the same DNA-binding domain linked to different
N-terminal regions (3234) or the use of viral E1 protein to
abrogate E2 transactivation function through direct protein-
protein interaction (35, 36). Because E2 only weakly activates
(2-fold) but strongly represses (up to 100-fold) the native HPV
promoter in both transfected cells (9) and in currently available
cell-free transcription systems (for review, see Ref. 28), our
initial effort was focused on dissecting the mechanisms of tran-
scriptional repression by various E2 proteins on their homolo-
gous E6 promoter (28).
Interaction between viral E1 and E2 proteins also regulates
papillomavirus DNA replication and viral episome mainte-
nance (3740). In addition to binding to viral DNA, E2 ensures
proper segregation of viral genomes during cellular replication
which is independent of the E2 DNA-binding domain (4143).
Analogous to abrogation of E2 transactivation function, it has
been suggested that E1 also regulates E2 binding to mitotic
chromosomes (44). Besides regulating viral transcription, rep-
lication, and viral episome maintenance, E2 has been impli-
cated in several cellular processes relevant to carcinogenesis.
E2 transfected into HPV DNA-containing cell lines can lead to
growth arrest, apoptosis, or abrogation of mitotic checkpoints
(4547). E2 repression of E6 and E7 oncogene expression lead-
ing to reactivation of p53 and pRb tumor suppressor pathways
has been the proposed mechanism for growth arrest in HPV-
positive cell lines (48, 49). Surprisingly, growth arrest and
repression of E6 and E7 expression require an intact E2 trans-
activation domain (50 52). Expression of E2 protein in a trans-
formed cell line devoid of HPV DNA also displayed growth
arrest through an undefined mechanism (45, 51). E2-triggered
apoptosis can occur through both p53-independent (53) and
p53-dependent pathways (54). Similar to E2-induced growth
arrest, E2 caused apoptosis in HPV-negative cell lines through
unknown pathways (54).
Although many functions have been attributed to papilloma-
viral E2 proteins, very few studies have been conducted to
compare directly the transcription activities among HPV and
BPV-1 E2 proteins (11, 55, 56). Experiments directly compar-
ing high risk HPV-16 and HPV-18 E2 proteins with low risk
HPV-6 E2 protein by transient transfection assays found that
high risk E2 proteins have higher activation activities than low
risk E2 protein, but this activity did not correlate with either
higher steady-state levels of protein in vivo or with DNA-
binding properties of the two classes of E2 proteins (56). There-
fore, it was hypothesized that high risk E2 proteins were in-
trinsically better transcriptional activators than low risk E2
proteins, which in turn regulate the levels of E6 and E7 onco-
proteins in transformed cells and thereby control the processes
leading to carcinogenesis. In a more recent study, high risk E2
was found to have higher affinity for transcriptional coactiva-
tor CREB-binding protein than that of low risk E2 (23). These
studies on E2 transcriptional activity were assayed primarily
in transiently transfected cells. Therefore, interpretation of the
results may be complicated by interference of other cellular
factors, potentially masking the true activity of E2 protein.
To address the differences in transcriptional activities be-
tween high and low risk HPV E2 proteins, we employed in vitro
transcription systems reconstituted either with individually
purified recombinant general transcription factors (TFIIB,
TFIIE, and TFIIF), epitope-tagged protein complexes (TFIID,
TFIIH, and pol II) and general cofactor PC4, or with a preas-
sembled RNA polymerase II (pol II) holoenzyme complex and
TBP (17, 28). Using this highly purified in vitro transcription
system along with an HPV responsive G-less cassette template
containing two E2-BSs (17), high risk HPV-16 E2 protein was
found to possess greater transactivation activity than those
exhibited by E2 proteins encoded by HPV-18, HPV-11, and
BPV-1. For repression assays, we employed an in vitro tran-
scription system reconstituted with pol II holoenzyme and TBP
with HPV template containing either homologous or heterolo-
gous URR linked to a G-less cassette. HPV-16 E2 (16E2) was
determined to be the strongest transcriptional repressor,
followed by BPV-1 E2 (BE2), HPV-18 E2 (18E2) and HPV-11
E2 (11E2). The transcriptional activities of these E2 proteins
correlated well with their corresponding binding affinities to
E2-BSs derived from the promoter-proximal regions of natu-
rally occurring HPV-11, HPV-16, and HPV 18 E6 promoters, as
well as BPV-1. Although previous in vivo studies revealed little
correlation between E2 transcriptional activities and DNA-
binding affinities, our in vitro studies indicate that higher
transcriptional activation and repression activities of various
E2 proteins correlate directly with their relative DNA-binding
affinities for E2-BSs.
EXPERIMENTAL PROCEDURES
Plasmid Constructions—Bacterial expression plasmids pF:E2-11d
(28), pF:16E2-11d, pF:18E2-11d, and p:BE2-11d were used to express
HPV-11, HPV-16, HPV-18, and BPV-1 E2 proteins, respectively. The
plasmid pF:16E2-11d was constructed by cloning the HPV-16 E2 open
reading frame, generated by PCR amplification using pCMV-16E2 (56)
as template with an NdeI site-containing sense primer (5-AAGTCGA-
CATATGGAGACTCTTTGCCAA-3)andaBamHI site-containing an-
tisense primer (5-AACTCGAGGATCCTCATATAGACATAAATCC-3),
into pF:E4-11d (28), after swapping the inserts between NdeI and
BamHI sites. The plasmid pF:18E2-11d was created similarly by using
pCMV-18E2 (56) as template with sense primer (5-AAGTCGACATA-
TGCAGACACCGAAGGAAA-3) and antisense primer (5-AACTCGAG-
GATCCTTACATTGTCATGTATCCC-3) for PCR amplification. Like-
wise, the plasmid pF:BE2-11d was cloned using pdBPV-1(1426) (57)
as template with sense primer (5-AAGTCGACATATGGAGACAGCA-
TGCGAA-3) and antisense primer (5-TTCTCGAGGATCCTATTGAT-
GCAAGC-3).
The G-less cassette template p18URR-GLess, containing the HPV-18
URR spanning nucleotides 6929 7857/181, used for in vitro transcrip-
tion was constructed by cloning the PCR product, generated from
genomic HPV-18 DNA (gift from L. T. Chow) using an EcoRI site-
containing sense primer (5-CTGGGTACCGAATTCGGATCCCTAT-
GATAAG-3) and an SacI site-containing antisense primer (5-CT-
GAGATCTGAGCTCTTTTATATACACCG-3) into the EcoRI-SacI-
linearized G-less cassette vector pGL (58). The p16URR-GLess G-less
cassette template, spanning HPV-16 URR nucleotides 70077904/172,
was constructed similarly by cloning the PCR product generated from
genomic HPV-16 DNA (gift from P. M. Howley) using sense primer
(5-CTGGGTACCGAATTCAGACCTAGATCAGTTTC-3) and antisense
primer (5-CTGAGATCTGAGCTCTTTTATACTAACCG-3) into pGL.
The other transcription templates, p11URR-GLess (originally named
p707270GLess/I
, see Ref. 28), p786270(23M)GLess/I
, pML53,
p2E2(IR)53, and pG
5
MLT, have already been described (17, 28).
Protein Purification—Bacterially expressed FLAG-tagged HPV-11
E2, HPV-16 E2, HPV-18 E2, and BPV-1 E2 were purified from
BL21(DE3)pLysS strain harboring pF:E2-11d, pF:16E2-11d, pF:18E2-
11d, and pF:BE2-11d, respectively, according to the protocol published
previously (17, 28). Isolation of recombinant FLAG-tagged human gen-
eral transcription factors (TFIIB, TBP, TFIIE
, and TFIIE
), FLAG-
tagged multiprotein complexes (TFIID, TFIIH, and pol II), six histidine-
tagged TFIIF subunits (RAP30 and RAP74), recombinant PC4, and
TFIID-deficient pol II holoenzyme was conducted as described previ-
ously (17, 28, 59 61).
In Vitro Transcription—The transcription repression assays were
performed in a two-component transcription system reconstituted with
TFIID-deficient pol II holoenzyme and TBP using 50 ng of p11URR-
GLess, p16URR-GLess, p18URR-GLess, or p786270(23M)GLess/I
G-less cassette template, 50 ng of pML53, 1 ng of TBP, and 3
l(90
ng) of pol II holoenzyme, in the absence or presence of 11E2, 16E2,
18E2, or BE2 protein as specified in the figures. Reactions were then
conducted and analyzed as described previously (28, 60).
Transcription and DNA Binding Activities of HPV and BPV E245620
by guest on February 9, 2017http://www.jbc.org/Downloaded from
The E2 transactivation assay was carried out in a highly purified in
vitro transcription system in a 30-
l reaction containing 50 ng of
pG
5
MLT, 50 ng of p2E2(IR)53, 10 ng of TFIIB, 20 ng of TFIID, 2.5 ng
each of TFIIE
and TFIIE
, 2 ng of TFIIF, 22.5 ng of TFIIH, and 5 ng
of pol II, in the presence or absence of 150 ng of PC4 and an increasing
amount of 11E2, 16E2, 18E2, or BE2 as specified, following the two-step
incubation protocol (17). Reactions were then performed and analyzed
as described previously (17, 60). Transcription signals were quantified
by PhosphorImager analysis (Molecular Dynamics).
Electrophoretic Mobility Shift Assay (EMSA)DNA probes used for
EMSA were generated by PCR amplification of respective HPV E6
promoter-proximal fragments and BPV-1 URR. The HPV-11 E6 pro-
moter-proximal fragments, HPV-11(790292), HPV-11(790292)3M,
HPV-11(790292)4M, and HPV-11(790292)34M, which span HPV-11
nucleotides 79027933/192 containing wild-type 3 and 4, mutated 3,
mutated 4, or mutated 3 and 4 E2-BSs, were created by PCR amplifi-
cation using sense primer (5-AACCCGGGTACCTACCCACACCCTA-
CATA-3) and antisense primer (5-GGACACAGATCTGAGCTCT-
GCTAATTTTTTGGG-3) with DNA templates pGL7072161, p24-N-
3M, p24-N-4M, and p24-N-34M (9, 28), respectively. The wild-type
HPV-16 E6 promoter-proximal DNA fragment HPV-16(7868 96), span-
ning HPV-16 nucleotides 7868 7904/196 and containing two wild-type
E2-BSs, was generated by PCR amplification of HPV-16 genomic DNA
(gift from P. M. Howley) using sense primer (5-CTGGGTACCGAAT-
TCTTACACATTTACAAGCA-3) and antisense primer (5-GGACACA-
GATCTGAGCTCTTTTGGTGCATAAA-3). The wild-type HPV-18 E6
promoter-proximal DNA fragment HPV-18(7834 101), spanning
HPV-18 nucleotides 7834 7857/1101 and containing two wild-type
E2-BSs, was generated by PCR amplification of HPV-18 genomic DNA
(gift from L. T. Chow) using sense primer (5-CTGGGTACCGAATTC-
TGGGCAGCACATACTAT-3) and antisense primer (5-GGACACAGA-
TCTGAGCTCATTGTGGTGTGTTTCTC-3). The BPV-1 DNA fragment
BPV-1(7874 85), spanning BPV-1 nucleotides 7874 7945/185 (num-
bering based on GenBank accession number NC_001522) and contain-
ing two wild-type E2-BSs, was generated by PCR amplification using
pdBPV-1(1426) as template (57) with sense primer (5-CTGGGTAC-
CGAATTCGCAGCATTATATTTTAAG-3) and antisense primer (5-G-
GACACAGATCTGAGCTCAACCGGGGTCTGTCAGC-3). These PCR-
amplified DNA fragments were then purified by gel electrophoresis
using a 2% agarose gel and used for end-labeling reactions with T4
polynucleotide kinase and [
-
32
P]ATP. DNA probes were purified fur-
ther by NICK column (Amersham Biosciences).
EMSA was conducted in a 10-
l reaction containing 0.5 fmol of
32
P-labeled DNA probe in the presence of 10 mMHEPES-Na (pH 7.9), 5
mMdithiothreitol, 0.2 mMEDTA, 4 mMMgCl
2
, 10% glycerol, 0.1 mg/ml
bovine serum albumin, 10 ng/
l poly(dI-dC), and increasing amounts of
respective purified E2 proteins (adjusted to 2
l in BC300). Reactions
were incubated at 30 °C for 1 h and then loaded onto a 5% polyacryl-
amide gel in 0.5TBE and run at 175 V for 2.5hat4°C. The gels
were then dried and exposed to x-ray films. The signals were quantified
by PhosphorImager analysis.
Calculation of Equilibrium Binding ConstantsThe intensity of the
remaining probe in each lane after EMSA was quantified using Image-
Quant (Molecular Dynamics) without (
) or with (
for each lane) an
increasing amount of 11E2, 16E2, 18E2, or BE2 protein. Fractional
occupancy (
) was then determined by Equation 1.
100% 1
/
兴兲 (Eq. 1)
Binding isotherms were obtained by plotting fractional occupancy as a
function of protein concentrations and analyzed by nonlinear least
squares analyses. The equilibrium binding constant, K
d
, was deter-
mined by analysis of the titration curves against Equation 2,
protein/共关proteinKd(Eq. 2)
where
fractional occupancy, [protein] protein concentration, and
K
d
the equilibrium binding constant (62).
RESULTS
Transcriptional Repression Mediated by High and Low Risk
HPV and BPV-1 E2 ProteinsTo compare the transcriptional
activities between high risk (HPV-16 and HPV-18) and low risk
(HPV-11) HPV and animal papillomavirus E2 proteins, we first
expressed full-length HPV-11, HPV-16, HPV-18, and BPV-1 E2
FIG.1. Comparison of HPV and
BPV-1 E2 proteins and the sequences
of different E2-BSs. A, structural fea-
tures of E2 proteins. The N-terminal
activation domain (AD) is linked to the
C-terminal DNA binding/dimerization do-
main (DBD) of each E2 protein via the
hinge (H) region. Numbers refer to bound-
aries of amino acids depicting the protein
domains found in 11E2, 16E2, 18E2, and
BE2 proteins, based upon reports pub-
lished previously (7275). The apparent
molecular mass (in kDa) of each FLAG-
tagged E2 protein, estimated by SDS-
PAGE, is listed on the right.B, Coomassie
Blue-stained gel of purified E2 proteins.
Recombinant FLAG-tagged BE2 (lane 1),
11E2 (lane 2), 16E2 (lane 3), and 18E2
(lane 4) were expressed in and purified
from bacteria, resolved by 10% SDS-
PAGE, and visualized after Coomassie
Blue staining. Positions of prestained pro-
tein size markers (in kDa) are indicated
on the left.C, transcriptional templates
containing natural HPV URR used for in
vitro transcription assays. Constructions
of various G-less cassettes driven by
HPV-11 URR (p11URR-GLess), HPV-16
URR (p16URR-GLess), or HPV-18 URR
(p18URR-GLess) were as described under
Experimental Procedures.Each tem-
plate contains the naturally occurring
HPV URR and the TATA box of the E6
promoter preceding a 377-nucleotide
G-less cassette. Nucleotide numbering for
boundaries of E2-BSs is based upon pre-
vious studies (9, 28) and HPV nucleotide
accession numbers AF125673 (for HPV-
16) and X05015 (for HPV-18).
Transcription and DNA Binding Activities of HPV and BPV E2 45621
by guest on February 9, 2017http://www.jbc.org/Downloaded from
proteins in bacteria. All E2 proteins share similar structural
features with an N-terminal activation domain linked to a
C-terminal DNA-binding/dimerization domain by a flexible
hinge region (Fig. 1A). These bacterially expressed E2 proteins
were purified to near homogeneity and clearly migrated at
positions corresponding to their predicted molecular sizes (Fig.
1B). The amounts of various purified E2 proteins were then
normalized by Western blotting with anti-FLAG monoclonal
antibody, which recognizes the same epitope sequence intro-
duced at the N terminus of individual E2 proteins (data not
shown). For functional studies, we also created three transcrip-
tion templates, containing enhancer elements and the E6 pro-
moter derived from each type of HPV, by cloning respective
URR into a G-less cassette of 377 nucleotides (Fig. 1C). We
demonstrated previously that correct initiation from the HPV
E6 promoter whose transcription is regulated by different
amounts of E2 proteins, similar to the in vivo observations, was
recapitulated in our cell-free transcription systems using the
G-less cassette templates (28).
When examined in a cell-free transcription system reconsti-
tuted with TBP and human pol II holoenzyme, which provides
pol II and general transcription factors TFIIB, TFIIE, TFIIF,
and TFIIH (60, 61), transcription from these URR-driven
G-less templates was inhibited by BE2, 11E2, 16E2, and 18E2,
in a dose-dependent manner (Fig. 2). Repression of the E6
promoter could be observed on both homologous and heterolo-
gous URR-driven templates, irrespective of the types of E2
proteins used (compare left panels of Fig. 2, AC). In contrast,
the internal control template pML53, which contains the ad-
enovirus major late core promoter linked to a G-less cassette of
280 nucleotides, did not respond to increasing amounts of E2
proteins, indicating a URR-specific repression mediated by E2
proteins. We had illustrated in our previous studies (28) that
repression of the E6 promoter from HPV-11 URR-containing
G-less cassette templates was mediated mainly through E2
binding to the promoter-proximal #3 and #4 E2-BSs, which are
immediately adjacent to the TATA box of the E6 promoter (see
Fig. 1C).
These comparative studies also revealed several interesting
findings. First, 16E2 is apparently a better repressor compared
with 18E2, 11E2, and BE2, because 10 ng of 16E2 almost
completely inhibited HPV transcription, whereas 10 ng of the
other E2 proteins only led to less than 50% of inhibition (com-
pare Fig. 2, AC,lanes 3,8,13, and 18, and the diagrams shown
on the right of each panel). Obviously, 50 ng (i.e. a 5-fold
higher amount) of 18E2, 11E2, and BE2 is required to reach a
level of repression similar to that achieved by 10 ng of 16E2
(see the diagrams in Fig. 2, AC). Again, stronger repression of
the E6 promoter by 16E2, relative to the other E2 proteins, was
observed on both homologous (Fig. 2B) and heterologous (Fig.
2, Aand C) HPV templates. Second, animal papillomavirus E2
protein (BE2) could also inhibit HPV E6 promoters as effi-
ciently as HPV E2 proteins (Fig. 2), consistent with many
previous studies using the BPV-1 E2 expression plasmid to
study HPV promoter regulation by transfection assays (11, 55,
56). Third, the repression activity of 18E2 is more similar to
that exhibited by 11E2 and BE2, but not to its closely related
family member (i.e. high risk 16E2). This observation could be
best explained by the available C-terminal DNA-binding/
dimerization domain structures of 16E2, 18E2, and BE2 (see
Discussion).
Transcriptional Activation Mediated by High and Low Risk
HPV and BPV-1 E2 ProteinsTo compare the transactivation
activity of E2 proteins encoded by HPV and BPV-1, we em-
ployed a highly purified in vitro transcription system reconsti-
tuted with only recombinant general transcription factors
(TFIIB, TFIIE, and TFIIF), general cofactor PC4, and epitope-
tagged multiprotein complexes (TFIID, TFIIH, and pol II). It
has been shown previously (17) that 11E2 can activate tran-
scription from p2E2(IR)53 DNA template, which contains two
copies of the HPV-11 #2 E2-BS linked to a major late core
promoter-driven G-less cassette of 280 nucleotides, but not
from an internal control template containing five copies of
Gal4-binding sites connected to a similar major late core pro-
moter construct with a longer G-less cassette (Fig. 3A,bottom
drawings). When we compared HPV and BPV-1 E2 proteins
directly in this reconstituted transcription system, we found
that 16E2 was the strongest activator achieving a maximum
level of activation at only 5 ng of protein compared with the
other E2 proteins (Fig. 3A, compare lanes 5,13,23, and 31, and
Fig. 3B). BE2 and 18E2 displayed similar levels of activation
(Fig. 3A,lane 7 versus lane 34, 11.6- and 13.7-fold activation,
respectively), but BE2 reached maximum activation at a lower
dose-dependent amount than 18E2 (Fig. 3A, compare lanes 7
and 34,50ngofBE2versus 150 ng of 18E2). 11E2 attained its
maximum level of activation at the same dosage as BE2 (Fig.
3A,lane 15, 50 ng of 11E2), but with only 7.4-fold activation.
After the maximum level of activation is achieved for each E2
protein, a general decline in fold activation, representing tran-
scriptional squelching, is observed (Fig. 3B). It is clear from
both experiments for activation and repression that 16E2 is the
dominant activator and repressor compared with 18E2, 11E2,
and BE2 in these comparative studies.
DNA-binding Activities of High and Low Risk HPV and
BPV-1 E2 ProteinsBecause E2 is a sequence-specific DNA-
binding protein, we speculated that the transcription activity of
E2 protein may correlate with its DNA-binding activity,
thereby accounting for the functional differences in transacti-
vation and repression between 16E2 and the other E2 proteins.
To explore this possibility, we performed EMSA using purified
E2 proteins and DNA probes derived from promoter-proximal
#3 and #4 E2-BSs of respective HPV E6 promoters and a URR
region containing #11 and #12 E2-BSs from BPV-1 (63). As
shown in Fig. 4A, two proteinDNA complexes (C1 and C2) were
detected with each E2 protein when a DNA probe containing
E2-BSs #3 and #4 of HPV-11 was used for EMSA. The C1
complex represents an E2 dimer binding to one E2-BS first
appearing at low concentrations of E2, whereas the C2 complex
is formed between two E2 dimers binding to both E2-BSs and
observed at increasing concentrations of E2 protein. At higher
concentrations of 16E2 and 18E2, we also observed additional
proteinDNA complexes migrating slower than the C2 complex
(Fig. 4A,lanes 30 and 40), presumably generated by oligomer-
ization of E2DNA complexes. An appearance of a band (indi-
cated by an asterisk) below the C1 complex of BE2 (Fig. 4A,
lanes 1120) is likely formed between minor degradation prod-
ucts of BE2 and the DNA probe. Consistent with BE2 being the
largest E2 protein in this assay, BE2DNA complexes migrated
slower than HPV E2DNA complexes. Similar patterns of
E2DNA complexes were also detected with DNA probes de-
rived from HPV-16 (Fig. 4B) and HPV-18 (Fig. 4C). When
BPV-1 probe was used, 16E2 and 18E2 showed migration pat-
terns similar to those observed with HPV probes (Fig. 4D,lanes
2140). The C1 and C2 complexes formed with BE2 apparently
dissociated during electrophoresis, causing two clusters of
large smears on the gel (Fig. 4D,lanes 1120). The affinity of
11E2 for the BPV probe is extremely low because no clear
complexes could be detected even with the use of 200 ng of
protein (Fig. 4D,lanes 110), which accidentally led to the
formation of large proteinDNA aggregates unable to enter the
gel. It is apparent from these EMSAs that the C1 complex was
formed at much lower concentrations of 16E2 and 18E2 than
Transcription and DNA Binding Activities of HPV and BPV E245622
by guest on February 9, 2017http://www.jbc.org/Downloaded from
11E2 and BE2 (compare lanes 2140 with lanes 120 of Fig. 4,
AC), indicating that high risk E2 has higher affinity for E2-
BSs than 11E2 and BE2.
To compare and quantify accurately the binding affinities
among various E2 proteins for each DNA probe, we calculated
the equilibrium binding constant (K
d
) estimated from three or
more gel shift assays by measuring the reduction of free probe
(Table I). Calculations of the equilibrium binding constant
clearly indicate that 16E2 has the highest affinity to both
homologous and heterologous HPV E6 promoter-proximal E2-
BSs, followed in order by 18E2, BE2, and 11E2 (see Table I, K
d
of 16E2:18E2:BE2:11E2 0.719:1.597:3.239:5.190 for HPV-11
FIG.2.Transcriptional repression mediated by high and low risk HPV and BPV-1 E2 proteins. A, transcriptional repression of the
HPV-11 E6 promoter mediated by different E2 proteins. In vitro transcription was performed in a two-component system comprising pol II
holoenzyme and TBP using p11URR-GLess template, which contains the HPV-11 URR spanning nucleotides 70727933/170, and the internal
control template pML53, in the absence () or presence of increasing amounts (2, 10, 50, and 200 ng) of BE2 (lanes 15, respectively), 11E2 (lanes
610), 16E2 (lanes 1115), or 18E2 (lanes 1620). Protein concentrations of different E2 proteins were normalized by Western blotting with
anti-FLAG M2 monoclonal antibody (Sigma). The line graph shows the relative transcription intensity of HPV-11 template with different amounts
of E2 protein. Relative intensity is defined as the signal intensity, quantified by PhosphorImager analysis, from p11URR-GLess relative to that
from the same DNA template performed in the absence of E2 (i.e. the first lane of each reaction set). B, transcriptional repression of the HPV-16
E6 promoter mediated by different E2 proteins. In vitro transcription was performed as described in A, except that a transcriptional template
p16URR-GLess containing the HPV-16 E6 promoter spanning HPV-16 nucleotides 70077904/172 was used. The line graph shows relative
transcriptional intensity for the HPV-16 E6 promoter at increasing concentrations for each E2 protein titration. C, transcriptional repression of
the HPV-18 E6 promoter mediated by different E2 proteins. In vitro transcription reactions were performed as described in A, except that a
transcriptional template p18URR-GLess containing the HPV-18 E6 promoter spanning HPV-18 nucleotides 6929 7857/181 was used. The line
graph shows relative transcriptional intensity for each HPV-18 E6 promoter signal derived from the titration of each E2 protein.
Transcription and DNA Binding Activities of HPV and BPV E2 45623
by guest on February 9, 2017http://www.jbc.org/Downloaded from
probe, and similar comparisons for HPV-16 and HPV-18
probes). In the case of BPV-1 DNA-derived probe containing
two E2-BSs, homologous BE2 bound with the highest affinity
(K
d
1.079) followed by 16E2 (K
d
1.567), 18E2 (K
d
8.422),
and 11E2 (K
d
25.360), consistent with results published
previously (11). This indicates that there may be some species
specificity between the DNA-binding activity of BPV and HPV
E2 proteins, which allows BE2 to bind better to BPV-1 DNA
and HPV E2 to bind better to HPV DNA. This phenomenon
likely reflects inherent variation between HPV and BPV E2-BS
sequences with HPV preferring A/T spacer nucleotides
(ACCGN
4
CGGT, see Fig. 1C), whereas BPV contains more G/C
spacer nucleotides within their respective E2-BSs. Differences
among E2 recognition of DNA (see Discussion) and decreased
stability of BE2 protein may also contribute to the variation in
results from EMSA performed with HPV DNA compared with
BPV-1 DNA (7, 64).
Binding Properties of Individual E2-BSs #3 and #4 in the
HPV-11 E6 Promoter-proximal RegionWe and others have
demonstrated previously that E2-mediated repression of the
E6 promoter via inhibition of preinitiation complex formation
functions mainly through E2 binding to the promoter-proximal
#4 E2-BS (9, 10, 25, 28). To explore the possibility that high
risk HPV E2 protein may bind with higher affinity to the
promoter-proximal #4 E2-BS than low risk 11E2 and BE2 to
mediate transcriptional repression, we performed EMSA using
an HPV-11 E6 promoter-proximal DNA probe with mutated #3
but wild-type #4 E2-BS (Fig. 5A). As expected, only one pre-
dominant complex (C1), formed via E2 binding to the #4 E2-BS,
was detected with HPV-11(790292)3M probe by different E2
proteins (Fig. 5A). Surprisingly, additional E2DNA complexes
(C2 and oligomers) were also observed with #3 E2-BS-mutated
DNA probe at higher concentrations of 16E2 and 18E2 (Fig. 5A,
lanes 10 and 1720), indicating that either high risk HPV E2
proteins do have higher DNA-binding activities able to over-
come half-site mutations, or the flanking sequences surround-
FIG.3. Transcriptional activation
mediated by different E2 proteins. A,
E2-dependent activation assays. Recon-
stituted in vitro transcription reactions
were performed with purified TFIIB,
TFIID, TFIIE, TFIIF, TFIIH, and pol II,
in the absence () or presence ()ofPC4
and increasing amounts of different E2
proteins using p2E2(IR)53 and pG
5
MLT
templates, as described under Experi-
mental Procedures.MLP, major core late
promoter. B, line graph displaying fold
activation for each E2 protein. Fold acti-
vation is defined as the signal intensity,
quantified by PhosphorImager, from
p2E2(IR)53 relative to that from the
same DNA template performed in the ab-
sence of E2 and PC4 (i.e. the first lane).
Transcription and DNA Binding Activities of HPV and BPV E245624
by guest on February 9, 2017http://www.jbc.org/Downloaded from
ing E2 half-site mutations may contribute to E2DNA recogni-
tion and binding. Comparison of the equilibrium binding
constants (K
d
, Table II) revealed that 16E2 indeed bound with
the highest affinity (K
d
0.871) to HPV-11(790292)3M probe,
followed by 18E2 (K
d
2.163), BE2 (K
d
5.026), and 11E2
(K
d
9.577). To test whether E2-BSs #3 and #4 in the HPV-11
E6 promoter-proximal region are equivalent for E2 binding
because they have the same consensus sequence, AC-
CGAAAACGGT (see Fig. 1C), we also performed EMSA with
DNA probe containing mutated #4 E2-BS (Fig. 5B). The results
were nearly identical to the mutated #3 probe except that
higher order HPV-16 E2DNA complexes (C2 and oligomer)
were clearly observed (Fig. 5B,lanes 810), again suggesting
that flanking sequences surrounding the E2-BS mutations in-
deed contribute to E2DNA binding, which is consistent with
previous studies (65). Although higher order E2DNA com-
plexes appeared with high risk E2 proteins when individual
E2-BSs were mutated, comparison of binding constants for
each E2 protein binding to either HPV-11(790292)3M or HPV-
11(790292)4M probe indicates that the binding properties of
each E2 protein to #3 and #4 E2-BSs are very similar (Table II,
compare 11E2 K
d
9.577 versus 10.261, 16E2 K
d
0.871 versus
0.674, 18E2 K
d
2.163 versus 1.807, and BE2 K
d
5.026 versus
7.569). As a control, additional EMSA was performed with a
DNA probe containing mutations at both #3 and #4 E2-BSs
(Fig. 5C). As expected, 11E2 was unable to bind to HPV-
11(790292)34M probe (Fig. 5C,lanes 110). Surprisingly,
proteinDNA complexes were detected with BE2, 16E2, and
18E2 (Fig. 5C,lanes 18 20,26 30, and 3540), albeit at sig-
nificantly lower affinities than binding to wild-type E2-BSs.
These studies further confirmed that high risk HPV E2 pro-
teins indeed have higher DNA-binding activity than low risk
E2 protein even to overcome mutations in the half-E2-BS,
consistent with previous studies reporting that the C-terminal
FIG.4. EMSA performed with E2
binding to DNA probes containing
two E2-BSs derived from different
types of HPVs and BPV-1. A, binding of
11E2, BE2, 16E2, and 18E2 to a wild-type
HPV-11 E6 promoter-proximal DNA frag-
ment containing two E2-BSs. EMSA was
performed in the absence () or presence
of increasing amounts of E2 protein (in
ng), as indicated above each lane, with a
32
P-labeled HPV-11 E6 promoter proxi-
mal-probe spanning HPV-11 nucleotides
79027933/192. The identities of the
shifted complexes, as indicated by C1,C2,
Oligomer, and *, are described under Re-
sults.B, binding of 11E2, BE2, 16E2, and
18E2 to a wild-type HPV-16 E6 promoter-
proximal DNA fragment containing two
E2-BSs. EMSA was performed as de-
scribed in A, except that an HPV-16 E6
promoter probe spanning HPV-16 nucleo-
tides 7868 7904/196 was used. C, bind-
ing of 11E2, BE2, 16E2, and 18E2 to a
wild-type HPV-18 E6 promoter-proximal
DNA fragment containing two E2-BSs.
EMSA was performed as described in A,
except that an HPV-18 E6 promoter probe
spanning HPV-18 nucleotides 7834
7857/1101 was used. D, binding of 11E2,
BE2, 16E2, and 18E2 to a wild-type
BPV-1 DNA fragment containing two E2-
BSs. EMSA was performed as described
in A, except that a BPV-1 DNA probe
spanning BPV-1 nucleotides 7874 7945/
185 was used.
TABLE I
Equilibrium binding constant (K
d
) for E2 binding to DNA probe with two E2-BSs
K
d
is in nMand was estimated from three or more gel shift assays.
Protein URR-derived DNA probe
HPV-11(790292) HPV-16(786896) HPV-18(7834101) BPV-1(787485)
11E2 5.190 0.680 4.740 0.617 4.969 0.690 25.360 4.279
16E2 0.719 0.061 0.721 0.050 0.701 0.062 1.567 0.118
18E2 1.597 0.231 1.610 0.102 1.706 0.175 8.422 0.638
BE2 3.239 0.414 3.306 0.301 4.518 0.437 1.079 0.187
Transcription and DNA Binding Activities of HPV and BPV E2 45625
by guest on February 9, 2017http://www.jbc.org/Downloaded from
domain of 16E2 has 180-fold higher affinity for nonspecific
DNA than the C-terminal domain of BE2 (66).
Correlation of Transcriptional Repression Activity with DNA-
binding Activity between 18E2 and 11E2There is a clear
distinction between repression activity of 16E2 and both 18E2
and 11E2 proteins (see Fig. 2). This activity correlated well
with the binding affinities among HPV E2 proteins to DNA
probes derived from the HPV E6 promoter-proximal region, in
that 16E2 had the highest affinity followed by 18E2 and then
11E2 (Tables I and II). If there is indeed a direct correlation
between binding affinities and E2 repressor activity, 18E2
should be a better transcriptional repressor than 11E2 because
it has a higher affinity for promoter-proximal E2-BSs than
11E2 (Table I and Fig. 4). However, this was not clearly ob-
served (Fig. 2). If the mechanism of transcriptional repression
at the E6 promoter is caused by disruption of preinitiation
complex formation through promoter-proximal DNA binding
(9, 10, 25, 28), then one possibility that may account for this
lack of difference is that there were not enough titration points
to measure the differences in activity between 18E2 and 11E2
adequately. The other possibility is that promoter-distal
E2-BSs may sequester E2 protein away from promoter-
proximal E2-BSs.
To address these issues, we performed an in vitro transcrip-
tion assay using the two-component transcription system with
an HPV-11 URR-containing G-less cassette template where
E2-BSs #2 and #3 were mutated (28) while carefully titrating
both 11E2 and 18E2 (Fig. 6A). At low concentrations of 11E2
and 18E2 (e.g. 1 ng), the activity from the HPV-11 E6 promoter
is almost reduced by half with 18E2, whereas 11E2 showed
negligible reduction in E6 promoter activity (Fig. 6A, compare
lane 15 with lane 5). Plotting relative intensity determined by
PhosphorImager analysis revealed that there is a significant
difference between the repression activities of 18E2 and 11E2
FIG.5. EMSA performed with E2
binding to DNA probes containing
mutated promoter-proximal E2-BSs
derived from HPV-11. EMSAs were
performed as described in Fig. 4, except
that
32
P-labeled HPV-11(790292)3M (A),
HPV-11(790292)4M (B), and HPV-
11(790292)34M (C) probes spanning
HPV-11 nucleotides 79027933/192 with
mutated #3, mutated #4, and mutated #3
and #4 E2-BSs were used.
TABLE II
Equilibrium binding constant (K
d
) for E2 binding to E2-BS-mutated HPV-11 probe
K
d
is in nMand was estimated from three or more gel shift assays.
Protein Mutated HPV-11 E6 promoter-proximal #3 and #4 E2-BSs
HPV-11(790292)3M HPV-11(790292)4M HPV-11(790292)34M
11E2 9.577 2.607 10.261 2.582 No binding
16E2 0.871 0.077 0.674 0.054 31.479 4.963
18E2 2.163 0.246 1.807 0.284 28.221 2.068
BE2 5.026 0.597 7.569 1.585 376.874 67.853
Transcription and DNA Binding Activities of HPV and BPV E245626
by guest on February 9, 2017http://www.jbc.org/Downloaded from
at low amounts of E2 proteins, which indeed correlates with
their relative DNA-binding affinities for E2-BSs (Fig. 6B).
DISCUSSION
In this report, we described the first comparative analysis of
transcription and DNA-binding activities of the full-length
high risk and low risk HPV and BPV-1 E2 proteins in well
defined cell-free environments devoid of other cellular proteins
whose presence in the cell may complicate the studies of intrin-
sic activities of E2 proteins. We uncovered a direct correlation
of transcriptional activity with the DNA-binding activity in
which high risk 16E2 is the strongest transcriptional activator/
repressor, compared with 18E2, 11E2, and BE2, and shows the
highest affinity for both homologous and heterologous HPV E6
promoter-proximal E2-BSs. Our studies thus provide unequiv-
ocal comparison of functional properties of E2 proteins encoded
by high risk and low risk HPVs as well as BPV-1, which play an
important role in virus-induced pathogenesis in both humans
and animals.
High Risk E2 Proteins Display More Transcriptional Activity
than Low Risk HPV and BPV-1 E2 ProteinsThe full-length
E2 protein can function as a transcriptional repressor to inhibit
the expression of virus-encoded gene products mainly through
promoter-proximal E2-BSs (25, 27, 28, 6769). We found that,
similar to high versus low risk HPV E6 and E7 oncoproteins,
there are also distinct differences in the repression activities
between high and low risk HPV E2 proteins. Using an in vitro
transcription system reconstituted with human pol II holoen-
zyme and TBP, we defined a strict order of transcriptional
repression activity with 16E2 acting as the strongest repressor,
followed by 18E2 and BE2, and lastly 11E2 (Figs. 2 and 6). The
potency of transcriptional repression activity exhibited by dif-
ferent E2 proteins strongly correlates with their DNA-binding
affinities for E2-BSs (Fig. 4 and Table I). 16E2 bound with the
highest affinity for the E6 promoter-proximal DNA fragments
derived from HPV-11, HPV-16, and HPV-18 and was followed
closely, again, by 18E2, BE2, and finally 11E2. The strong
DNA-binding activity observed with 16E2 and 18E2 even al-
lows these high risk E2 proteins to bind to DNA probes con-
taining half-site mutations in both promoter-proximal #3 and
#4 E2-BSs (Fig. 5), whose binding could not occur with low risk
11E2 protein.
Although comparison of transactivation activity among some
high and low risk HPV E2 proteins has been described in
previous studies using either transient transfection or in vitro
transcription performed with HeLa nuclear extract (11, 55, 56),
the presence of numerous unidentified cellular proteins in their
assays often complicates the interpretation. Therefore we have
developed a highly purified E2-dependent in vitro transactiva-
tion system (17) comprised only of essential recombinant gen-
eral transcription factors (TFIIB, TFIIE, and TFIIF), recombi-
nant general cofactor PC4, and epitope-tagged multiprotein
complexes (TFIID, TFIIH, and pol II). Using this reconstituted
cell-free transcription system, we found that 16E2 is the most
responsive and strongest transactivator compared with 18E2,
BE2, and 11E2 (Fig. 3), suggesting that 16E2 may be more
effective at recruiting components of the general transcription
machinery to the promoter region. This interesting possibility
remains to be investigated further.
Significance for the Order of Transcriptional Activity between
High and Low Risk HPV E2 ProteinsIn our studies, we found
a strict order of both transcriptional activation and repression
activity among various E2 proteins. High risk 16E2 displays
the highest activation and repression activity, followed by high
risk 18E2 and then low risk 11E2. In a recent epidemiological
FIG.6. Transcriptional repression
of the HPV-11 E6 promoter mediated
by 11E2 and 18E2. A, transcriptional re-
pression of the HPV-11 E6 promoter by
different amounts of E2 proteins. In vitro
transcription was performed in a two-
component system as described in Fig.
2A, except that p786270(23M)GLess/I
(28), a transcriptional template contain-
ing HPV-11 URR spanning nucleotides
78627933/170 with mutations at #2
and #3 E2-BSs, was used. B, line graph of
relative intensity for 11E2 and 18E2 pro-
teins. Relative intensity is defined as the
signal intensity, quantified by Phospho-
rImager, from p786270(23M)GLess/I
relative to that from the same DNA tem-
plate performed in the absence of E2 (i.e.
the first lane of each reaction set).
Transcription and DNA Binding Activities of HPV and BPV E2 45627
by guest on February 9, 2017http://www.jbc.org/Downloaded from
study, 99.7% of 1,000 cases of invasive cervical cancer were
HPV DNA-positive with HPV-16 DNA (53%) being the most
prevalent followed by HPV-18 DNA (15%) (2, 3). Previous stud-
ies have also shown that high risk HPV E6 and E7 are more
active in inducing cellular transformation than low risk E6 and
E7 (for review, see Refs. 4 and 5). A common hallmark of
cervical cancer is viral integration, which often disrupts the E2
open reading frame, leading to the loss of E6 promoter regula-
tion and thus increased the expression of HPV E6 and E7
oncoproteins. We speculate that HPV-16 developed a more
transcriptionally active E2 protein to regulate its highly active
E6 and E7 oncoproteins tightly and govern the viral life cycle
more stringently than low risk HPVs, which express less potent
E6 and E7 proteins. Once regulation of the E6 promoter by high
risk E2 proteins is lost because of viral integration, cancer may
then develop.
Differences in DNA Target Site Discrimination by Papilloma-
virus E2 ProteinsAlthough previous studies of binding site
affinities between HPV and BPV-1 E2 proteins revealed little
variation (11, 56), we found a distinct order of E2 binding to
both wild-type and mutated HPV E6 promoter-proximal E2BSs
among 16E2, 18E2, BE2, and 11E2 (Figs. 4 and 5 and Tables I
and II). One possible explanation for the variation between our
results and previously published studies might be attributed to
the DNA probes from which the equilibrium binding constants
were calculated. Although we used a much larger DNA probe
containing two E2-BSs derived from the natural HPV URRs,
others have used much smaller DNA probes containing only
one or two E2-BSs with minimal flanking regions (11, 56).
Smaller DNA probes would naturally have more inherent DNA
flexibility than larger DNA probes, which would favor BE2
binding to more flexible DNA, whereas HPV E2 has a predis-
position to bind to pre-bent DNA (7). Variations in Mg
2
salt
concentrations between DNA-binding experiments may also
play a role in the differences in calculating equilibrium binding
constants because there is evidence that Mg
2
enhances HPV
E2 recognition of specific E2-BSs (70). Furthermore, the four-
nucleotide spacers within our E2-BSs also vary from previous
DNA probes (11), consistent with previous studies identifying
that HPV E2 favors spacer nucleotides that are A/T-rich, spe-
cifically AATT, whereas BPV E2 does not display an ability to
discriminate between spacer nucleotides (7, 71). This may ex-
plain our results in which 16E2 favors binding to DNA probes
derived from HPV URR, whereas BE2 binds better to BPV-
derived E2-BSs that are less A/T-rich (Fig. 4 and Table I), a
conclusion in agreement with previous studies (11).
It should not be surprising that E2 proteins vary in binding
affinities, considering that there are distinct differences in the
quaternary structures of the DNA-binding domains between
16E2 and 18E2 (7). These differences are critical because they
dictate the spatial arrangement of side chains presented to the
major grooves for DNA sequence recognition (64). Further-
more, structural and biochemical studies of E2 proteins from
BPV-1 and HPV-16 have suggested they use different mecha-
nisms to discriminate their DNA targets (71) even though they
are highly homologous proteins. Recent E2DNA-binding stud-
ies have also reported that there are significant differences in
the binding affinities between the DNA-binding domains of
HPV-16 and BPV-1 (66). In fact, BE2 shares more structural
similarities to 18E2 than 16E2, whereas 16E2 is structurally
more similar to HPV-31 E2 protein (7). This may explain why
the transcriptional and DNA-binding activity of 18E2 is more
equivalent to BE2 than to its closely related human homolog
16E2. Although BE2 and 18E2 are structurally similar, and
they both induce the same global DNA deformation upon bind-
ing, their mechanisms of deformation are different. BE2 has a
positive charge localized near the C terminus of the recognition
helix, whereas 18E2 has a positive charge in the center of its
DNA interaction surface, and 16E2 is even less charged all
along the DNA interaction surface (7). This would lead to
differences in E2DNA contacts, with 16E2 making fewer con-
tacts with the DNA phosphate backbone than either BE2 or
18E2 (7), which would be consistent with increased DNA-bind-
ing specificity. Although the DNA-binding domain sequences of
high risk 16E2 and 18E2 share 54% identity and 77% homol-
ogy, they still differ in DNA binding. Thus it is likely that the
DNA-binding domain of low risk 11E2 may vary further in
structure from both 16E2 and 18E2, thus contributing to vari-
ations in DNA-binding affinity among high and low risk HPV
E2 proteins. This remains to be elucidated once the structure of
11E2 becomes available.
AcknowledgmentsWe thank L. T. Chow for genomic HPV-18 DNA,
R. Kovelman for pCMV-16E2 and pCMV-18E2 plasmids, P. M. Howley
for genomic HPV-16 DNA and pdBPV-1 (1426) plasmid, and K. Chiu
along with J. Watson for assistance in plasmid constructions. We are
also grateful to Mary C. Thomas for insightful discussion and critical
reading of the manuscript.
REFERENCES
1. zur Hausen, H. (2000) J. Natl. Cancer Inst. 92, 690 698
2. Dell, G., and Gaston, K. (2001) Cell. Mol. Life Sci. 58, 19231942
3. Mun˜oz, N. (2000) J. Clin. Virol. 19, 15
4. Mantovani, F., and Banks, L. (2001) Oncogene 20, 7874 7887
5. Mu¨nger, K., Basile, J. R., Duensing, S., Eichten, A., Gonzalez, S. L., Grace, M.,
and Zacny, V. L. (2001) Oncogene 20, 7888 7898
6. OConnor, M., Chan, S. Y., and Bernard, H.-U. (1995) in Human Papilloma-
viruses 1995 Compendium, Part III-A (Myers, G., Bernard, H.-U., Delius,
H., Baker, C., Icenogle, J., Halpern, A., and Wheeler, C. (eds) pp. 2140, Los
Alamos National Laboratory, Los Alamos, NM
7. Hegde, R. S. (2002) Annu. Rev. Biophys. Biomol. Struct. 31, 343360
8. Chong, T., Apt, D., Gloss, B., Isa, M., and Bernard, H.-U. (1991) J. Virol. 65,
59335943
9. Dong, G., Broker, T. R., and Chow, L. T. (1994) J. Virol. 68, 11151127
10. Tan, S.-H., Leong, L. E.-C., Walker, P. A., and Bernard, H.-U. (1994) J. Virol.
68, 64116420
11. Ushikai, M., Lace, M. J., Yamakawa, Y., Kono, M., Anson, J., Ishiji, T.,
Parkkinen, S., Wicker, N., Valentine, M.-E., Davidson, I., Turek, L. P., and
Haugen, T. H. (1994) J. Virol. 68, 66556666
12. Rank, N. M., and Lambert, P. F. (1995) J. Virol. 69, 63236334
13. Steger, G., Ham, J., Lefebvre, O., and Yaniv, M. (1995) EMBO J. 14, 329 340
14. Benson, J. D., Lawande, R., and Howley, P. M. (1997) J. Virol. 71, 80418047
15. Yao, J.-M., Breiding, D. E., and Androphy, E. J. (1998) J. Virol. 72, 10131019
16. Enzenauer, C., Mengus, G., Lavigne, A., Davidson, I., Pfister, H., and May, M.
(1998) Intervirology 41, 80 90
17. Wu, S.-Y., and Chiang, C.-M. (2001) J. Biol. Chem. 276, 3423534243
18. Li, R., Knight, J. D., Jackson, S. P., Tjian, R., and Botchan, M. R. (1991) Cell
65, 493505
19. Breiding, D. E., Sverdrup, F., Grossel, M., Moscufo, N., Boonchai, W., and
Androphy, E. J. (1997) Mol. Cell. Biol. 17, 7208 7219
20. Boner, W., Taylor, E. R., Tsirimonaki, E., Yamane, K., Campo, M. S., and
Morgan, I. M. (2002) J. Biol. Chem. 277, 2229722303
21. Lefebvre, O., Steger, G., and Yaniv, M. (1997) J. Mol. Biol. 266, 465478
22. Peng, Y.-C., Breiding, D. E., Sverdrup, F., Richard, J., and Androphy, E. J.
(2000) J. Virol. 74, 58725879
23. Lee, D., Lee, B., Kim, J., Kim, D. W., and Choe, J. (2000) J. Biol. Chem. 275,
70457051
24. Lee, D., Hwang, S. G., Kim, J., and Choe, J. (2002) J. Biol. Chem. 277,
64836489
25. Dostatni, N., Lambert, P. F., Sousa, R., Ham, J., Howley, P. M., and Yaniv, M.
(1991) Genes Dev. 5, 16571671
26. Demeret, C., Yaniv, M., and Thierry, F. (1994) J. Virol. 68, 70757082
27. Stubenrauch, F., Leigh, I. M., and Pfister, H. (1996) J. Virol. 70, 119 126
28. Hou, S. Y., Wu, S.-Y., Zhou, T., Thomas, M. C., and Chiang, C.-M. (2000) Mol.
Cell. Biol. 20, 113125
29. Jackson, M. E., and Campo, M. S. (1995) J. Virol. 69, 6038 6046
30. Boeckle, S., Pfister, H., and Steger, G. (2002) Virology 293, 103117
31. Hartley, K. A., and Alexander, K. A. (2002) J. Virol. 76, 5014 5023
32. Chiang, C.-M., Broker, T. R., and Chow, L. T. (1991) J. Virol. 65, 33173329
33. Stubenrauch, F., Hummel, M., Iftner, T., and Laimins, L. A. (2000) J. Virol. 74,
1178 1186
34. Stubenrauch, F., Zobel, T., and Iftner, T. (2001) J. Virol. 75, 4139 4149
35. Sandler, A. B., Vande Pol, S. B., and Spalholz, B. A. (1993) J. Virol. 67,
5079 5087
36. Ferran, M. C., and McBride, A. A. (1998) J. Virol. 72, 796 801
37. Mohr, I. J., Clark, R., Sun, S., Androphy, E. J., MacPherson, P., and Botchan,
M. R. (1990) Science 250, 1694 1699
38. Ustav, M., and Stenlund, A. (1991) EMBO J. 10, 449 457
39. Frattini, M. G., and Laimins, L. A. (1994) Proc. Natl. Acad. Sci. U. S. A. 91,
12398 12402
40. Piirsoo, M., Ustav, E., Mandel, T., Stenlund, A., and Ustav, M. (1996) EMBO
J. 15, 111
Transcription and DNA Binding Activities of HPV and BPV E245628
by guest on February 9, 2017http://www.jbc.org/Downloaded from
41. Lehman, C. W., and Botchan, M. R. (1998) Proc. Natl. Acad. Sci. U. S. A. 95,
4338 4343
42. Skiadopoulos, M. H., and McBride, A. A. (1998) J. Virol. 72, 2079 2088
43. Ilves, I., Kivi, S., and Ustav, M. (1999) J. Virol. 73, 4404 4412
44. Voitenleitner, C., and Botchan, M. (2002) J. Virol. 76, 3440 3451
45. Hwang, E. S., Riese, D. J., Jr., Settleman, J., Nilson, L. A., Honig, J., Flynn, S.,
and DiMaio, D. (1993) J. Virol. 67, 3720 3729
46. Desaintes, C., Demeret, C., Goyat, S., Yaniv, M., and Thierry, F. (1997) EMBO
J. 16, 504 514
47. Frattini, M. G., Hurst, S. D., Lim, H. B., Swaminathan, S., and Laimins, L. A.
(1997) EMBO J. 16, 318 331
48. Wells, S. I., Francis, D. A., Karpova, A. Y., Dowhanick, J. J., Benson, J. D., and
Howley, P. M. (2000) EMBO J. 19, 57625771
49. Goodwin, E. C., and DiMaio, D. (2000) Proc. Natl. Acad. Sci. U. S. A. 97,
1251312518
50. Dowhanick, J. J., McBride, A. A., and Howley, P. M. (1995) J. Virol. 69,
77917799
51. Goodwin, E. C., Naeger, L. K., Breiding, D. E., Androphy, E. J., and DiMaio, D.
(1998) J. Virol. 72, 39253934
52. Francis, D. A., Schmid, S. I., and Howley, P. M. (2000) J. Virol. 74, 2679 2686
53. Desaintes, C., Goyat, S., Garbay, S., Yaniv, M., and Thierry, F. (1999) Onco-
gene 18, 4538 4545
54. Webster, K., Parish, J., Pandya, M., Stern, P. L., Clarke, A. R., and Gaston, K.
(2000) J. Biol. Chem. 275, 8794
55. Bouvard, V., Story, A., Pim, D., and Banks, L. (1994) EMBO J. 13, 54515459
56. Kovelman, R., Bilter, G. K., Glezer, E., Tsou, A. Y., and Barbosa, M. S. (1996)
J. Virol. 70, 7549 7560
57. Sarver, N., Byrne, J. C., and Howley, P. M. (1982) Proc. Natl. Acad. Sci.
U. S. A. 79, 71477151
58. Wang, J. C., Sawadogo, M., and Van Dyke, M. W. (1998) Biochim. Biophys.
Acta 1397, 141145
59. Wu, S.-Y., Kershnar, E., and Chiang, C.-M. (1998) EMBO J. 17, 4478 4490
60. Wu, S.-Y., and Chiang, C.-M. (1998) J. Biol. Chem. 273, 1249212498
61. Wu, S.-Y., Thomas, M. C., Hou, S. Y., Likhite, V., and Chiang, C.-M. (1999)
J. Biol. Chem. 274, 23480 23490
62. Nelson, D. L., and Cox, M. M. (2000) Lehninger Principles of Biochemistry, 3rd
Ed., pp. 203242, Worth Publishers, New York
63. Li, R., Knight, J., Bream, G., Stenlund, A., and Botchan, M. (1989) Genes Dev.
3, 510 526
64. Kim, S.-S., Tam, J. K., Wang, A.-F., and Hegde, R. S. (2000) J. Biol. Chem. 275,
3124531254
65. Pepinsky, R. B., Prakash, S. S., Corina, K., Grossel, M. J., Barsoum, J., and
Androphy, E. J. (1997) J. Virol. 71, 828 831
66. Ferreiro, D. U., Lima, L. M. T. R., Nadra, A. D., Alonso, L. G., Goldbaum, F. A.,
and Prat-Gay, G. (2000) Biochemistry 39, 1469214701
67. Bernard, B. A., Bailly, C., Lenoir, M. C., Darmon, M., Thierry, F., and Yaniv,
M. (1989) J. Virol. 63, 43174324
68. Stenlund, A., and Botchan, M. R. (1990) Genes Dev. 4, 123136
69. Thierry, F., and Howley, P. M. (1991) New Biol. 3, 90 100
70. Lewis, H., and Gaston, K. (1999) J. Mol. Biol. 294, 885896
71. Hines, C. S., Meghoo, C., Shetty, S., Biburger, M., Brenowitz, M., and Hedge,
R. S. (1998) J. Mol. Biol. 276, 809 818
72. Abroi, A., Kurg, R., and Ustav, M. (1996) J. Virol. 70, 6169 6179
73. Zou, N., Lin, B. Y., Duan, F., Lee, K.-Y., Jin, G., Guan, R., Yao, G., Lefkowitz,
E. J., Broker, T. R., and Chow L. T. (2000) J. Virol. 74, 37613770
74. Bellanger, S., Demeret, C., Goyat, S., and Thierry, F. (2001) J. Virol. 75,
7244 7251
75. Antson, A. A., Burns, J. E., Moroz, O. V., Scott, D. J., Sanders, C. M.,
Bronstein, I. B., Dodson, G. G., Wilson, K. S., and Maitland, N. J. (2000)
Nature 403, 805809
Transcription and DNA Binding Activities of HPV and BPV E2 45629
by guest on February 9, 2017http://www.jbc.org/Downloaded from
Samuel Y. Hou, Shwu-Yuan Wu and Cheng-Ming Chiang
Proteins Correlates with E2 DNA Binding
Transcriptional Activity among High and Low Risk Human Papillomavirus E2
doi: 10.1074/jbc.M206829200 originally published online September 17, 2002
2002, 277:45619-45629.J. Biol. Chem.
10.1074/jbc.M206829200Access the most updated version of this article at doi:
Alerts:
When a correction for this article is posted When this article is cited
to choose from all of JBC's e-mail alertsClick here
http://www.jbc.org/content/277/47/45619.full.html#ref-list-1
This article cites 73 references, 53 of which can be accessed free at
by guest on February 9, 2017http://www.jbc.org/Downloaded from
... E 1 is essential for the viral DNA replication, while E 2 besides acting in cooperation with E 1 for the viral replication represent the main transcriptional regulator. In particular, binding to the LCR, E 2 tightly regulates the viral transcription mediated by the early promoter, showing both a transactivation and a strong repression function when present in low and high concentrations, respectively [5], [6]. Before the late stage, high levels of E 2 inhibit the early promoter, thus shutting down E 6 and E 7 oncogenes expression, paving the way for cellular differentiation. ...
... where the state variables are the concentrations [nM] of: the primary transcript x, mRNAs mE i , i = 1,2,4,5,6,7, proteins E 1 and E 2 . refers for the transcription of x enhanced by low values of E 2 concentration and repressed by elevated E 2 concentration [5] (Fig. 2A) according to (7). ...
... β 1 and β 2 are the rate constants for E 1 and E 2 protein translation assumed to be linearly related to the cognate mRNAs, see (4)(5). ...
Conference Paper
In high risk forms, human papillomaviruses (HPV) can either induce or promote cancerous lesions, especially cervical cancer which is considered the second most common cancer in the women worldwide. HPV life cycle is tightly linked to the infected cell differentiation program and its evolution is strictly joined to the switch between the early and the late viral polycistronic promoters.The aim of this study is to develop a novel mathematical model which collects and structures the available biologic knowledge on the early promoter regulation for HPV in episomal form. The model includes the main regulation by E2 viral protein as well as a novel discovered co-regulation function mediated by the viral E1 protein. Only by including both E2 and E1 regulatory effect the model is able to correctly predict the temporal behaviour of the early promoter switching off. A possible use of the model as in silico tool to evaluate new antiviral therapies is discussed.
... Os HPV de baixo risco são responsáveis por acarretar inúmeras lesões benignas, como verrugas, papiloma laríngeo e tumores ano-genitais, manifestando, por vezes, sintomas como: prurido e dores. Por sua vez, os HPV de alto risco também podem provocar lesões benignas na mucosa, porém, esses tipos carcinogênicos são principalmente os associados ao câncer de colo do útero (Hou et a., 2002;Ozbun, 2002. ...
... Salientamos que as proteínas E1 e E2 modulam, de fato, o ciclo infeccioso do HPV. A quantidade de E1 e E2 é um indicador do número de genomas virais presentes na célula e ao mesmo tempo, um modo de regular o número de genomas virais (Hubert & Laiminis, 2002;Hou et al., 2002;Lee et al., 2002). ...
... Em infecções com espécies e HPV de alto risco, as proteínas virais E6 e E7 são bastante ativas, interferindo profundamente no ciclo celular. Como resultado, a divisão celular se processa mais rapidamente do que em infecções com espécies de HPV de baixo risco, aumentando a probabilidade de ocorrer, acidentalmente, numa das células, uma integração do DNA no genoma celular (Hou et al., 2002;Wagner & Hewlett, 2003). ...
... By binding to these sites, E2 recruits at the LCR a number of cellular proteins necessary to support transcription, replication and mitotic segregation of the viral genome. The stability of E2/DNA complexes differs according to the E2BS flanking sequences [203][204][205], and this has been proposed to modulate the effect of E2 on early promoter transcription [206]. ...
... These results substantiate both the functional relevance and the specificity of 16E2/GTF2B interaction. Hou and colleagues [206] have demonstrated that 16E2 is an especially potent transcriptional activator in a reconstituted E2-dependent in vitro test and hypothesized that it might result from a more efficient recruitment of general transcription regulators to promoter sequences. The specific targeting of GTF2B by HPV16 E2 substantiates this hypothesis. ...
Article
Papillomaviruses are responsible for widespread infections in humans, causing pathegenesis ranging from inapparent infections to benign lesions, hyperplasia or cancers. Given the major public health concern due to HPV-associated cancers, most studies have focused on the early proteins expressed by the most clinically relevant HPVs most frequently found in cancers. Among the early proteins encoded by HPVs, the E2 protein regulates viral transcription, replication and mitotic segregation of the viral genome, mainly through the recruitment of host factors to the HPV regulatory region. E2 is therefore pivotal for both the viral productive cycle and for viral persistence, which is a major risk factor for cancer development. In addition, the E2 proteins have been shown to engage interactions important to directly modulate the host cell, thereby contributing to create suitable cell conditions for the successive stages of the HPV life cycle. Interestingly, some E2's roles have been demonstrated to be specific to the oncogenic HPVs, raising the idea that beyond its role in the general HPV regulation, E2 could also directly influence the fate of cancer development. This thesis aimed at providing an overview of E2's functions across multiple HPV genotypes and at identifying specific features that distinguish the different HPV pathological traits. We mapped the virus-host interaction networks of the E2 proteins from a panel of 12 HPVs selected to be representative of the HPV diversity. Clustering of E2's interaction profiles correlated with the HPV phylogeny, raising the notion that E2 could directly contribute to the HPV pathogenesis. This work also emphasizes that the E2 proteins, like many other viral proteins, tend to target highly connected cellular proteins (cellular hubs), which is presumed to be an evolutionary way to maximize viral impacts on the host. E2 predominantly targets a subset of key cellular processes, like transcriptional regulation, apoptosis, RNA metabolism, ubiquitination or intracellular transport, which both confirms already known E2's functions and points to potential new functions. In addition, this large-scale comparative approach offers a framework to pinpoint interactions that are specifically associated with the most represented HPVs in cancers and therefore can be used as targets for the development of new therapeutics. In particular, we identified a specific interaction between the E2 protein from HPV16 and a cellular protein, CCHCR1, involved in the regulation of keratinocyte proliferation. We determined that CCHCR1's interaction domain on E2 overlaps with that of BRD4, a major interactor of E2, inducing a physical competition between the two cellular proteins. This competitive binding affects BRD4-mediated enhancement of E2's transcriptional activity, suggesting that the interaction with CCHCR1 might have an impact on the role of E2 in the infected cell. In addition, we showed that CCHCR1 induces the docking of HPV16 E2 into the cytoplasm which could further affect E2's nuclear functions. We also demonstrated that CCHCR1 impairs HPV16 E2's induction of keratinocytes early differentiation, presumably resulting from the negative effect of CCHCR1 on the nuclear functions of E2. This effect could have drastic consequences on the oncogenic potential of HPV16 and could participate to high prevalence in cancers of HPV16. Taken together, these results enhance the general understanding of the impact of E2 during HPV infections and highlights its contribution in the HPV pathogenesis. E2 appears as a critical factor that participates in the global hijacking of the host cell to allow the virus to replicate despite the hostile environment. E2 also emerges as a viral component susceptible to directly influence the outcome of an HPV infection and to potentially impact on the preliminary steps of carcinogenic conversion.
... HPVs infect epithelial cells in the genital mucosa (alpha Papillomaviruses only), oral mucosa or skin (representatives of all five genera). HPVs cause a wide range of diseases from benign lesions to invasive tumors (Bzhalava et al., 2015), where twelve mucosal human papillomavirus (HPV) has been considered as carcinogenic to humans (types: (Bzhalava et al., 2015;Bzhalava et al., 2013;Faridi et al., 2011;Hou et al., 2002). ...
Article
Cervical cancer is the second most common cancer in women worldwide, and human papillomavirus infection is the main risk cause for developing this disease around the world. HPV E5 proteins are short proteins that comprise a hydrophobic domain that contains at least one amphipathic α-helix. HPV E5 is an oncoprotein and it is multifunctional. Among all functions described for HPV E5, perhaps, the modulation of the MHC class I constitute the most important, because of its implication during the immune response. In this work, a large-immunoinformatics analysis based on genomic data it was performed in order to provide more insight of this enigmatic oncoprotein during the immune response, through the analysis of the epitope densities and physical-chemical properties for HPV proteins using bioinformatics and immunoinformatics approaches combined with statistical analysis. It was found that HPV E5 have a higher epitope density in comparison with the rest of proteins in its proteome, which suggest that this protein is subjected to low immune pressure. Correlations analysis based on the epitope density and computed physical-chemical properties suggest that hydrophobicity (GRAVY: r = 0.6857, r² = 0.4702 and hydrophobic amino acid frequency: r = 0.6907, r² = 0.4771) is the main factor conducting the immune evasion in HPV E5.
... HPVs that produce skin lesions are low risk HPVs, but HPV-6 and HPV-11 are responsible for about 90 percent of all cases of genital warts in males and females but do not cause cervical cancer. The mechanism of infection of these viruses has been widely investigated; particularly the oncogene protein Papillomaviridae E2/E1 and E6/E7, since they are considered the essential part for the development of cancer cells (Hou et al., 2002). The U.S. Food and Drug Administration (FDA) have approved two vaccines, Gardasil® and Cervarix® that protect against infection by HPVtypes 16 and 18. ...
Chapter
Cancer has been one of the leading causes of death in the developed countries while in the developing countries; it comes second after deaths caused by acquired immunodeficiency syndrome (AIDS) and its complications. Approximately, 12.5 million new cases of cancer are being diagnosed worldwide each year, and considerable research efforts are in progress for finding a cure to this deadly disease. A number of approaches for optimization of cancer treatment have been investigated by different researchers all over the world. Among them nanostructured drug delivery systems such as liposomes and polymeric nanoparticles are the most prominent technologies. Conventional cancer therapies are associated with adverse effects and toxicity due to distribution of chemotherapeutic agents throughout the body affecting both normal and tumoral cells. Nanostructured drug delivery systems and novel therapeutic approaches to treat cancer form the basis of the chapter. Firstly, polymeric and lipidic nanoparticles are dealt with in detail. The novel therapeutic approaches exemplified by PEGylation, antibody drug conjugate (ADC) and anticancer vaccines are then discussed. Theranostics is a rapidly evolving field in nanotechnology which represents a multi-modal strategy in the battle against cancer, opening up a new field in which nanotechnology has set the stage for an evolutionary leap in the therapy and diagnosis of human cancer. The fundamentals of theranostic have been highlighted along with their applications.
Article
High risk forms of human papillomaviruses (HPVs) promote cancerous lesions and are implicated in almost all cervical cancer. Of particular relevance to cancer progression is regulation of the early promoter that controls gene expression in the initial phases of infection and can eventually lead to pre-cancer progression. Our goal was to develop a stochastic model to investigate the control mechanisms that regulate gene expression from the HPV early promoter. Our model integrates modules that account for transcriptional, post-transcriptional, translational and post-translational regulation of E1 and E2 early genes to form a functioning gene regulatory network. Each module consists of a set of biochemical steps whose stochastic evolution is governed by a chemical Master Equation and can be simulated using the Gillespie algorithm. To investigate the role of noise in gene expression, we compared our stochastic simulations with solutions to ordinary differential equations for the mean behavior of the system that are valid under the conditions of large molecular abundances and quasi-equilibrium for fast reactions. The model produced results consistent with known HPV biology. Our simulation results suggest that stochasticity plays a pivotal role in determining the dynamics of HPV gene expression. In particular, the combination of positive and negative feedback regulation generates stochastic bursts of gene expression. Analysis of the model reveals that regulation at the promoter affects burst amplitude and frequency, whereas splicing is more specialized to regulate burst frequency. Our results also suggest that splicing enhancers are a significant source of stochasticity in pre-mRNA abundance and that the number of viruses infecting the host cell represents a third important source of stochasticity in gene expression.
Conference Paper
High risk human papillomavirus (HPV) can induce cervical and oropharyngeal cancerous lesions. Different HPV strains, as well as still unknown mechanisms can be associated to a range of biochemical parameters that can importantly affect the HPV gene expression dynamics. For this reason, it is of pivotal importance to investigate how parameters variation can induce interesting behaviors such as viral latency in place of the normal gene replication activity. The aim of this study is to perform bifurcation analysis on a minimal model of the early promoter regulatory core controlled by E2 transcriptional regulation. The bifurcation analysis showed how E2 regulation can induce a bistability on the early promoter gene expression that could explain the interplay between viral latency and gene replication regimen.
Article
Full-text available
Gene regulatory network can help to analyze and understand the underlying regulatory mechanism and the interaction among genes, and it plays a central role in morphogenesis of complex diseases such as cancer. DNA sequencing technology has efficiently produced a large amount of data for constructing gene regulatory networks. However, measured gene expression data usually contain uncertain noise, and inference of gene regulatory network model under non-Gaussian noise is a challenging issue which needs to be addressed. In this study, a joint algorithm integrating genetic programming and particle filter is presented to infer the ordinary differential equations model of gene regulatory network. The strategy uses genetic programming to identify the terms of ordinary differential equations, and applies particle filtering to estimate the parameters corresponding to each term. We systematically discuss the convergence and complexity of the proposed algorithm, and verify the efficiency and effectiveness of the proposed method compared to the existing approaches. Furthermore, we show the utility of our inference algorithm using a real HeLa dataset. In summary, a novel algorithm is proposed to infer the gene regulatory networks under non-Gaussian noise and the results show that this method can achieve more accurate models compared to the existing inference algorithms based on biological datasets.
Article
Full-text available
NF-κB is a transcription factor and if activated, it induces apoptosis inhibition. Phalerin from Phaleria macrocarpa fruits expected to inhibit NF-κB activation. This research is to investigate anticancer mechanism of Phaleria macrocarpa fruits extracts (EBMD) in NF-κB pathway. Molecular docking assay was performed to determine phalerin affinity to IKK. Cytotoxic activity was observed by MTT assay. Double staining was performed to determine the apoptotic cells. Docking score of phalerin to IKK is -60. The IC50 value of EBMD is 629 μg/mL. Apoptosis profile shows (shown that) many cells undergoing apoptosis after treatment. Thus, EBMD potentially inhibits activation of NF-κB pathway and triggers apoptosis on HeLa cells. Keywords : NF-κB, Phaleria macrocarpa, sel HeLa, Bcl-2, IKK, molecular docking
Article
Full-text available
Human papillomaviruses (HPVs) have been linked to a variety of human diseases, most notably cancer of the cervix, a disease responsible for at least 200,000 deaths per year worldwide. Over 100 different types of HPV have been identified and these can be divided into two groups. Low-risk HPV types are the causative agent of benign warts. High-risk HPV types are associated with cancer. This review focuses on the role of high-risk HPV types in cervical tumorigenesis. Recent work has uncovered new cellular partners for many of the HPV early proteins and thrown light on many of the pathways and processes in which these viral proteins intervene. At the same time, structural and biochemical studies are revealing the molecular details of viral protein function. Several of these new avenues of research have the potential to lead to new approaches to the treatment and prevention of cervical cancer.
Article
Full-text available
The human papillomavirus (HPV) E2 protein regulates viral gene expression and is also required for viral replication. HPV-transformed cells often contain chromosomally integrated copies of the HPV genome in which the viral E2 gene is disrupted. We have shown previously that re-expression of the HPV 16 E2 protein in HPV 16-transformed cells results in cell death via apoptosis. Here we show that the HPV 16 E2 protein can induce apoptosis in both HPV-transformed and non-HPV-transformed cell lines. E2-induced apoptosis is abrogated by a trans-dominant negative mutant of p53 or by overexpression of the HPV 16 E6 protein, but is increased by overexpression of wild-type p53. We show that mutations that block the DNA binding activity of E2 do not impair the ability of this protein to induce apoptosis. In contrast, removal of both N-terminal domains from the E2 dimer completely blocks E2-induced cell death. Heterodimers formed between wild-type E2 and N-terminally deleted E2 proteins also fail to induce cell death. Our data suggest that neither the DNA binding activity of E2 nor other HPV proteins are required for the induction of apoptosis by E2 and that E2-induced cell death occurs via a p53-dependent pathway.
Article
Full-text available
A human RNA polymerase II (pol II) complex was isolated from a HeLa-derived cell line that conditionally expresses an epitope-tagged RPB9 subunit of human pol II. The isolated FLAG-tagged pol II complex (f:pol II) contains a subset of general transcription factors but is devoid of TFIID and TFIIA. In conjunction with TATA-binding protein (TBP) or TFIID, f:pol II is able to mediate both basal and activated transcription by Gal4-VP16 when a transcriptional coactivator PC4 is also provided. Interestingly, PC4, in the absence of a transcriptional activator, actually functions as a repressor to inhibit basal transcription. Remarkably, TBP is able to mediate activator function in this transcription system. The presence of TBP-associated factors, however, helps overcome PC4 repression and further enhance the level of activation mediated by TBP. Alleviation of PC4 repression can also be achieved by preincubation of the transcriptional components with the DNA template. Sarkosyl disruption of preinitiation complex formation further illustrates that PC4 can only inhibit transcription prior to the assembly of a functional preinitiation complex. These results suggest that PC4 represses basal transcription by preventing the assembly of a functional preinitiation complex, but it has no effect on the later steps of the transcriptional process.
Article
Full-text available
The papillomavirus E2 proteins are transcriptional regulators that bind to a consensus DNA sequence ACCG NNNN CGGT. Multiple copies of this binding site are found in the viral genomes. The affinities of the naturally occurring binding sites for the E2 proteins are predominantly dependent upon the sequence of the NNNN spacer. The hierarchies of binding site affinities among the sites present in the viral genomes result in differential occupancy during the viral life-cycle. In turn, this differential binding regulates transcription from viral promoters, including those for the oncogenes E6 and E7. Structural and biochemical studies have shown that E2 proteins bend the DNA to which they specifically bind. Atomic resolution structures of complexes of the bovine papillomavirus strain 1 (BPV-1) E2 protein and DNA show that the protein does not contact the spacer DNA. A direct comparison of the binding of the DNA-binding domains of the E2 proteins from BPV-1 and human papillomavirus strain 16 (HPV-16) to a series of binding sites as a function of the sequence of their central spacer and/or the presence of a nick or gap in one strand of the spacer DNA is presented in this paper. The BPV-1 E2 DNA-binding domain is only moderately sensitive to the nature of the central spacer; less than several fold differences in affinity were observed when the DNA sequence of the spacer was varied and/or a nick or gap was introduced. In contrast, the HPV-16 E2 DNA-binding domain binds to sites containing A:T-rich central spacers with significantly increased affinity. The introduction of a nick or gap into the spacer of these high affinity sequences is very detrimental to HPV-16 E2 binding while comparable nicks or gaps have only small effects in the low affinity sequences. These results suggest that the HPV-16 E2 protein recognizes the structure of the DNA spacer and that the mechanism of DNA-sequence specific binding of the homologous HPV-16 E2 and BPV-1 E2 proteins is significantly different.
Article
Full-text available
The bovine papillomavirus type 1 E2 transactivator protein is required for viral transcriptional regulation and DNA replication and may be important for long-term episomal maintenance of viral genomes within replicating cells (M. Piirsoo, E. Ustav, T. Mandel, A. Stenlund, and M. Ustav, EMBO J. 15:1-11, 1996). We have evidence that, in contrast to most other transcriptional transactivators, the E2 transactivator protein is associated with mitotic chromosomes in dividing cells. The shorter E2-TR and E8/E2 repressor proteins do not bind to mitotic chromatin, and the N-terminal transactivation domain of the E2 protein is necessary for the association. However, the DNA binding function of E2 is not required. We have found that bovine papillomavirus type 1 genomes are also associated with mitotic chromosomes, and we propose a model in which E2-bound viral genomes are transiently associated with cellular chromosomes during mitosis to ensure that viral genomes are segregated to daughter cells in approximately equal numbers.
Article
Full-text available
The late gene promoter P7535 of the epidermodysplasia verruciformis-associated human papillomavirus type 8 (HPV8) is regulated by the viral E2 protein. Transfection experiments performed with the human skin keratinocyte cell line RTS3b and P7535 reporter plasmids revealed transactivation at low amounts and a repression of basal promoter activity at high amounts of E2 expression vector. This repression was promoter specific and correlated with the amount of transiently expressed E2 protein. Mutational analyses revealed that the negative regulation of P7535 activity is mediated by the low-affinity E2 binding site P2, which is separated by one nucleotide from the P7535 TATA box. Biochemical and genetic analyses suggested that repression is due to a displacement of the TATA-box binding protein by E2 and an interference of E2 with promoter-activating cellular factors that specifically recognize the P2 sequence. The high conservation of the P2 sequence among several papillomaviruses (epidermodysplasia verruciformis-associated HPVs, HPV1, cottontail rabbit papillomavirus, and bovine papillomavirus type 1) in the vicinity of the late gene promoter cap site suggests that an interplay of E2 and cellular factors at this sequence element is important for the expression of structural proteins.
Article
Full-text available
The human papillomavirus (HPV) transcription/replication factor E2 is essential for the life cycle of HPVs. E2 protein binds to DNA target sequences in the viral long control regions to regulate transcription of the viral genome. It also enhances viral DNA replication by interacting with the viral replication factor E1 and recruiting it to the origin of replication and may also play a more direct role in replication. The cellular proteins with which E2 interacts to carry out these functions are largely unknown. To identify these proteins a yeast two-hybrid screen was carried out with the transcription/replication domain of HPV16 E2. This screen identified several candidate interacting partners for E2 including TopBP1 (topoisomerase II beta-binding protein 1). TopBP1 has eight BRCA1 carboxyl-terminal domains that are found in proteins regulating the DNA damage response, transcription, and replication. Here we demonstrate that HPV16 E2 and TopBP1 interact in vitro and in vivo and that TopBP1 can enhance the ability of E2 to activate transcription and replication. This is the first time that TopBP1 has been shown to function as a transcriptional coactivator and that E2 interacts with TopBP1. Removal of the amino-terminal domain of TopBP1 abolishes coactivation of transcription and replication. This interaction may have functional consequences upon the viral life cycle.
Article
Eukaryotic viruses can maintain latency in dividing cells as extrachromosomal nuclear plasmids. Segregation and nuclear retention of DNA is, therefore, a key issue in retaining copy number. The E2 enhancer protein of the papillomaviruses is required for viral DNA replication and transcription. Viral mutants that prevent phosphorylation of the bovine papillomavirus type 1 (BPV) E2 protein are transformation-defective, despite normal viral gene expression and replication function. Cell colonies harboring such mutants show sectoring of viral DNA and are unable to maintain the episome. We find that transforming viral DNA attaches to mitotic chromosomes, in contrast to the mutant genome encoding the E2 phosphorylation mutant. Second-site suppressor mutations were uncovered in both E1 and E2 genes that allow for transformation, maintenance, and chromosomal attachment. E2 protein was also found to colocalize to mitotic chromosomes, whereas the mutant did not, suggesting a direct role for E2 in viral attachment to chromosomes. Such viral hitch-hiking onto cellular chromosomes is likely to provide a general mechanism for maintaining nuclear plasmids.
Article
Transcription of the transforming genes E6 and E7 of the human papillomavirus type 18 (HPV18) can be repressed by the product of the E2 open reading frame. Mutations were introduced in the cis-responsive elements upstream from the E6 and E7 transcriptional promoter, P105. The effect of these mutations in the absence or presence of E2 was examined in the human cervical carcinoma cell line C33, transfected with expression plasmids. In the presence of HPV18 E2, repression was relieved only when both of the two E2 binding motifs, immediately proximal to the P105 TATA box, were mutated. Mutation of a third E2 binding site, 110 nucleotides upstream of the TATA box, led to a slight but consistent activation in the presence of the full-length E2. Therefore, three E2 binding sites appear to be involved in the negative regulation of the P105 promoter by HPV18 E2. In the presence of the BPV1 E2 protein, only the E2 binding site most proximal to the promoter TATA box was involved in the P105 repression. The comparative analysis of the in vitro and in vivo properties of the human and bovine E2 proteins on the HPV18 P105 promoter provides new insights into the mechanism of transcriptional repression, which appears to involve steric hindrance at the site of transcriptional initiation.