You are on page 1of 10

Biochimica et Biophysica Acta 1816 (2011) 57–66

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / b b a c a n

Review

Role of p63 in cancer development


Vincenzo Graziano a,b, Vincenzo De Laurenzi a,b,c,⁎
a
Dipartimento di Scienze Biomediche, Universita“G. d'Annunzio” Chieti-Pescara, Via Colle dell'Ara 1, 66100 Chieti, Italy
b
Fondazione “G. D'Annunzio", Centro Studi sull'Invecchiamento, Ce.S.I., Via Colle dell'Ara 1, 66100 Chieti, Italy
c
BioUniverSA srl, DIFARMA, University of Salerno, Fisciano (SA) 84084, Italy

a r t i c l e i n f o a b s t r a c t

Article history: Since their initial identification p53 homologues p63 and p73 have been expected to play a role in cancer
Received 21 February 2011 development due to their close homology to p53, notoriously one of the most mutated genes in cancer.
Received in revised form 5 April 2011 However soon after their discovery the awareness that these genes were rarely mutated in cancer seemed to
Accepted 8 April 2011
indicate that they did not play a role in its development. However a large number of data collected in the
Available online 15 April 2011
following years indicated that altered expression rather than mutation could be found in different neoplasia
Keywords:
and play a role in its biology. In particular p63 due to its fundamental role in epithelial development seems to
p63 play a role in a number of tumors of epithelial origin. In this review we summarize some of the evidence
Oncogenes linking p63 to carcinogenesis.
Oncosuppressors © 2011 Elsevier B.V. All rights reserved.
Cancerogenesis
Molecular biology

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
2. Altered expression of p63 in cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
2.1. Breast cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
2.2. Squamous cell carcinomas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2.3. Lung cancer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2.4. Other cancers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3. Potential mechanism explaining the role of p63 in cancer progression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4. Role of p63 in metastasis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

1. Introduction three genes had similar and redundant functions [1–5]. However
over a decade of work has shown that while the three proteins
p63 belongs to the family of transcription factors that also includes share some common functions they also have completely unique
p53 and p73. All members of the family have three highly conserved physiological roles. Similarly to p53 both p63 and p73 can induce cell
domains: a transactivation domain (TA), a DNA binding domain cycle arrest and apoptosis activating similar pathways. This is a
(DBD) and an oligomerization domain (OD), p63 and p73 have an complex gene family and a number of different proteins are generated
additional protein interacting domain at their c-terminus known as from the three genes. In particular alternative splicing results in a
sterile alpha motif (SAM) (Fig. 1). Based on the high homology number of variants that differ in the carboxy terminal part of the
between the three proteins it has initially been postulated that the protein while usage of an alternative promoter generates amino-
terminally truncated proteins known as ΔN isoforms (Fig. 1). The ΔN
isoforms initially identified for p73 and p63 and more recently also for
p53 lack the TA domain and therefore can have opposite function to
⁎ Corresponding author at: Dipartimento di Scienze Biomediche, Universita“G.
the full length TA isoforms. Indeed it has been shown that these
d'Annunzio” Chieti-Pescara, Via Colle dell'Ara 1, 66100 Chieti, Italy. Tel.: + 39 0871
541580; fax: + 39 0871 541598. isoforms promote cell growth and can protect cells from apoptosis
E-mail address: delaurenzi@unich.it (V. De Laurenzi). [1,6].

0304-419X/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.bbcan.2011.04.002
58 V. Graziano, V. De Laurenzi / Biochimica et Biophysica Acta 1816 (2011) 57–66

Fig. 1. Schematic representation of p53 family genes and proteins. (A) p63, (B) p53, (C) p73. Gene structures (upper panels) show alternative splicing and different promoters (P1
and P2) used. Transcribed regions are in white while untranslated regions are in black. Protein structures (lower panels). Main structural domains are indicated: transactivation
domain (TA); DNA binding domain (DBD); oligomerization domain (OD); sterile alpha motif (SAM). All combinations of alternative N-terminus and C-terminus for all proteins are
supposed to be formed.

Results from genetic studies in mice reveal the different functions of ratio but die early as a consequence of inflammation and recurrent
these proteins. While p53 KO mice are viable and do not show any major infections, moreover they show hippocampal dysgenesis, hydrocepha-
developmental defect, but are tumor prone both p73 and p63 KO mice lus and heterozygous animals when aged develop an Alzheimer like
have developmental phenotypes. Total p73 KO are born at a mendelian phenotype [7,8]. Selective TA KO have a similar but milder neuronal
V. Graziano, V. De Laurenzi / Biochimica et Biophysica Acta 1816 (2011) 57–66 59

phenotype, but in addition show chromosomal instability and develop the different findings linking p63 to tumor progression and invasive as
spontaneous tumors [9,10]. p63 KO mice show profound alterations of well as metastasizing ability of different tumors.
skin and other stratified epithelia as well as epithelial appendages
including mammary and salivary glands, teeth and hair follicles. These 2. Altered expression of p63 in cancer
animals also show truncated limbs and craniofacial abnormalities as a
consequence of an alteration of the apical ectodermal ridge. Similarly 2.1. Breast cancer
mutations of p63 in humans are responsible for a number of syndromes
known as ectodermal dysplasias. The interpretation of the role of p63 in In normal mammary tissues p63 is normally expressed in the
epithelial development based on the KO phenotypes however remains nucleus of a single layer of basal myoepithelial cells surrounding
controversial and two interpretation models have been proposed by the glandular cells [19,20]. In addition most reports seem to suggest that the
two groups that have independently generated the mice. One model ΔNp63 isoform is the isoform mostly expressed in normal myoepithelial
proposes that p63 is essential for commitment and differentiation of cells [20] while TA isoform is expressed in the lining cells but absent in
epithelial precursors while the other suggests that it is essential for myoepithelial and basal cells [21]. Despite the fact that mutations of this
maintenance of the progenitor population [11,12]. A model that can in gene have not consistently been described in breast tumors a number of
part reconcile these two different views has also been proposed and is reports show that in some cases tumor cells have altered p63 expression
supported by some experimental data. In this last model ΔNp63 would (Table 1). In general p63 over-expression is found only in a small
be essential for the maintenance of the progenitor population in the percentage of breast carcinomas [19,22] however several studies show
basal layer while TAp63 would be required together with ΔN or that this usually occurs in less differentiated, more aggressive tumors or
subsequently to ΔN to allow complete differentiation [1,13–15]. some particular subtypes of breast cancer. Indeed p63 increased
Similarly studies from the same two groups on the ability of p63+/− expression has been reported in: triple negative basal-like poorly
mice to develop spontaneous tumors reach opposite conclusions differentiated ductal carcinomas not otherwise specified [22–26],
indicating p63 as an oncogene or a tumor suppressor [16–18]. sarcomatoid/metaplastic carcinomas [27], adenoid-cystic carcinomas
The described role of p63 isoforms in development of stratified [20,26,28], myoepithelial–epithelial carcinomas, malignant adeno-
epithelia implies that altered levels of these proteins can play a role myoepitheliomas [19,20], and invasive papillary carcinomas [29]. In
also in epithelial cancer development. In this review we summarize accordance with the hypothesis that p63 expression is a marker of

Table 1
p63 expression in cancer.

Tumors site Subtypes P63 % Isoforms Prognosis or progression References

Breast Basal-like carcinoma Increased 22% ΔN Increased expression in bigger tumors. [19,20,22,23,25,26,30,33,34]
Increased with nuclear pleomorphism.
Loss correlates with poorer prognosis
Sarcomatoid/metaplastic carcinoma Increased 50–69% Unknown Unknown [27]
Adenoid-cystic carcinoma Increased 100% ΔN Unknown [20,26,28]
Myoepithelial–epithelial carcinoma Increased Unknown ΔN Unknown [19,20]
Malignant adenomyoepithelioma Increased Unknown Unknown Unknown [19,20]
Invasive papillary carcinoma Increased 33.3% Unknown Unknown [29]
Head and neck Squamous cell carcinoma Increased 81–100% ΔN Correlates with both better or [38,40,42–45,49,55–58]
poorer prognosis
Small cell tumors of sino-nasal tract Increased 20–100% ΔN Unknown [39,50]
Merkel cell carcinoma Increased 53.2% Unknown Correlates with poorer prognosis [51]
Sarcomatoid carcinoma Increased 63% Unknown Unknown [41]
Lung Squamous cell carcinoma Increased 88–100% ΔN Correlates with better prognosis [46,49,64,67–69]
Large cell lung cancer Increased 42–66% Unknown Correlates with better prognosis [46,47,49,66–68]
Adenocarcinomas Increased 11–30% ΔN Correlates with better prognosis [46,47,49,66–68,70]
Cytoplasmic expression correlates with
poorer prognosis
Small cell lung cancer Increased 20% Unknown Correlates with poorer prognosis [68,70]
Correlates with higher grade tumors
Large cell neuroendocrine tumors Increased 50–70% Unknown Correlates with poorer prognosis [68,70]
Correlates with higher grade tumors
Sarcomatoid cancer Increased 50% Unknown Unknown [41]
Esophagus Squamous cell carcinoma Increased 52–98% ΔN No correlation [73–83]
Loss correlates with poorer prognosis.
Barret's esophagus Increased Unknown Unknown [77,78]
Bladder Transitional cell carcinoma Incresead 20–75% ΔN Increase in early stage tumors. [85–93]
Decreased Unknown TA Correlates with poorer prognosis.
Sarcomatoid cancer Unknown Unknown Unknown Unknown [41]
Prostate Adenoid-cystic (basal) carcinoma Increased 60–100% Unknown Unknown [105–107]
Adenocarcinoma Increased 3–10% ΔN Cytoplasmic expression correlates [102,108]
with poorer prognosis
Liver Cholangiocarcinoma Increased 23–100% Unknown Unknown [114,115]
Ovary Benign-borderline Brenner tumors Increased 100% Unknown Unknown [110]
Malignant Brenner tumors Increased 17% Unknown Unknown [110]
Serous tumors Increased 0–13% ΔN Correlates with poorer prognosis [37,112,113]
Endometrial tumors Increased 0–25% Unknown Unknown [37,112]
Metastatic transitional cancer cell Increased Unknown Unknown Unknown [111]
Non-Hodgkin Follicular lymphoma Increased 54–99% TA Correlates with poorer prognosis [121–124]
lymphoma Diffuse large B-cell lymphoma Increased 15–34% TA Correlates with poorer prognosis [121–124]
Correlates with better prognosis
Anaplastic large cell lymphoma Increased 44% Unknown Unknown [120]

% indicates the percentage of tumors showing altered p63 expression.


60 V. Graziano, V. De Laurenzi / Biochimica et Biophysica Acta 1816 (2011) 57–66

undifferentiated basal like breast tumors several studies have shown TA isoforms does not seem to have great importance [42,43,50,52].
that more differentiated breast carcinomas such as: lobular like Within tumors p63 seems to be expressed in non-keratinizing areas of
carcinoma, intra-ductal carcinoma, low grade ductal carcinoma, tubular squamous cancers (both well and poorly differentiated), while it lacks
and medullary carcinomas are always p63 negative [19,20,29]. in central keratinizing areas of well differentiated squamous cancers
Most of the clinical studies reported however do not distinguish [40].
between the different p63 isoforms expressed. When an attempt to More recent work supports this hypothesis demonstrating a role for
investigate which isoform is expressed in breast cancer cells has been ΔNp63 in the evolution of pre-cancerous lesions like leukoplakia [44],
done it seems that the isoform that is mostly if not exclusively lichen planus [52], actinic cheilitis [53], or erythroleukoplakia [54] in
expressed is the ΔN isoform [30]. It should be mentioned that malignancies. Interestingly p63 expression is reduced in oral lichen
occasionally TAp63 expression has been reported as an example planus (OLP), a pre-cancerous condition that often shows spontaneous
Nylander and colleagues using selective TA and ΔN antibodies show regression, suggesting a potential causative role in cancer progression.
that while ΔN expression is restricted to basal like tumors cells, TA is ΔNp63 is in fact under expressed in OLP lesions compared to the
expressed in most breast cancer cells regardless of the subtype [21]. adjacent normal mucosa while no changes have been reported for TA
A role for ΔNp63 in basal like breast cancers is in accordance with the isoform expression. This suggests that under-expression of ΔN isoform
fact that this isoform is normally expressed in cells of the basal layer and is essential for the apoptotic regression typically observed in these
it has been suggested to play a role in immortalization of basal epithelial lesions [52,54]. Nuclei of actinic cheilitis of the lips and lip squamous cell
cells in early steps of carcinogenesis [31]. In accordance with this Zucchi carcinomas show higher levels of p63 compared with normal lips
and co-workers show that p63 is necessary for the maintenance of epithelium [53]. Furthermore over-expression of ΔNp63 in leukoplakias
stemness and regeneration of cancer stem cell spheres derived from a is correlated to an increased risk of developing oral cancer, and patients
murine breast cancer cell line [32]. with oral cancer developed from a ΔNp63 negative pre-cancerous lesion
Several studies have attempted to correlate p63 expression with have a cancer-free-survival-rate longer than those derived from ΔNp63
prognosis (Table 1) but further work is required to obtain a clear picture positive ones [44].
of the value of this gene as a prognostic marker. A correlation of p63 The fact that p63 expression is often altered in these cancers has
expression with increased features of poor prognosis [33] has been suggested that it could be an important prognostic marker, however
reported, indeed expression increases from grade IIIa to grade IIIb–IV the number of studies published so far are controversial on this
breast carcinomas [19] and correlates with nuclear pleomorphism, a matter. Some studies show that ΔNp63 over-expression correlates
known feature of aggressive tumors [34]. Moreover in one study it was with a better prognosis [42,55,56], others show that it correlates with
reported that p63 is negative in cancers smaller than 2 cm, but its a worse prognosis [51,57,58]. Indeed, Merkel cell carcinomas strongly
expression increases with tumor size [19]. Conversely well differenti- staining for p63 are associated to a more aggressive, Ki67 positive
ated tumors over expressing p63 are associated to a good prognosis [23]. tumor resulting in a lower overall survival of the patients [51]. TAp63
It is possible that expression of p63 in cells of different origin has a down-regulation has also been associated with more advanced
different phenotype or more likely that the isoforms expressed by the clinical stage and a greater tumor diameter [59], again supporting
different tumors are different and result in a different outcome. Indeed the idea that an imbalance between the two isoforms in favor of ΔN is
one could imagine that while basal like tumors express the oncogenic responsible for tumor progression. In addition while some studies
ΔN isoforms, more differentiated tumors express the tumor suppressor seem to suggest that p63 over-expressing tumors show a better
TA form. Hopefully future studies investigating differential isoform response to cis-platin treatment [55,60–62] others show that p63
expression will clarify this point and allow the use of p63 as a prognostic over-expressing tumors are more resistant to radiation therapy
marker. treatment [48,63].
Response to therapeutic drugs in vitro also gives potentially
controversial results, as an example increased ΔNp63 levels have been 2.3. Lung cancer
reported to increase resistance to doxorubicin [35], but also to
increased response to cis-platin [30]. In vivo clinical studies show that p63 is strongly expressed in the nuclei of myoepithelial cells
p63 expression in tumor cells correlates with a higher response rate to surrounding bronchial and bronchiolar epithelium [37,41,46,47,64–68],
cis-platin based chemotherapy regimens [36]. Finally a retrospective but not in pulmonary epithelium nor in neuroendocrine cells normally
clinical study shows that well differentiated ER positive cancers found in the lung [41,66,68].
expressing m-RNA of p63 in nuclei are associated to a better response p63 is often over-expressed in different lung cancers, including:
to endocrine therapy than ER positive cancers that do not express it squamous cell lung cancers[41,46,49,64,67–69], large cell lung cancer,
[23]. adenocarcinomas (where it is also occasionally found in the cytoplasm)
[41,46,47,49,66–68,70], high grade neuroendocrine (NET) lung cancers
2.2. Squamous cell carcinomas [68,70], and sarcomatoid cancers [41]. This is particularly relevant in
squamous cell lung cancers where up to 90% show up-regulation of p63.
Altered p63 expression patterns have also been found in tumors On the contrary low-grade small cell lung cancers are constantly
derived from pluristratified epithelia (Table 1). While most of the negative [64,68,69] in line with the neuroendocrine origin of these cells.
work done so far reports a role of this gene in squamous cell The importance of p63 in lung cancer development is also outlined
carcinomas of the head and neck [37–45] involvement in squamous by the finding that p63 expression increases from hyperplasia to
tumors of the esophagous and the lung have also been reported and metaplasia, to dysplasia reaching its maximum in severe grade
will be discussed separately [37,46,47]. dysplasias [47,65,68]. While most of these clinical studies have not
p63 is over expressed in 81% to 100% [38,44,45] of head and neck distinguished between the different p63 isoforms more recent work
squamous cell carcinomas (HNSCC) [38–45,48,49] as well as in some suggests again a predominant role for the ΔN isoform [37,46,47,66].
particular types of head and neck cancers such as small cell tumors of Similarly to what described above for other types of tumors while p63
sino-nasal tract like nasopharyngeal carcinomas [39,50], Merkel cell is clearly over-expressed in some lung cancer its correlation with
carcinomas (also found in other localizations) [51] and head and neck prognosis in terms of overall survival and disease-free-survival is less
sarcomatoid carcinomas (spindle cell carcinomas) [41]. Again most of clear. While some reports show no correlation between expression and
the reports that have looked at the expression of different isoforms prognosis [68,71] in others increased expression of p63 in NSCLC
show that the ΔN isoform (particularly the α one) is usually the one (particularly the squamous form) correlates with good prognosis [46,72].
over-expressed in these tumors [42,43,50,52] while the expression of Conversely, over-expression in NET correlates with bad tumor features in
V. Graziano, V. De Laurenzi / Biochimica et Biophysica Acta 1816 (2011) 57–66 61

terms of patient performance status, tumor diameter, differentiation and 3. Potential mechanism explaining the role of p63 in cancer
Ki-67 expression [68] and prognosis [70]. Interestingly the cytoplasmic progression
expression described in adenocarcinomas associates with poor prognosis
[66] suggesting that in some cases ectopic expression rather than, or The number of evidences reported above clearly shows that
associated with, altered quantitative expression plays a role in cancer altered p63 expression is linked to tumor progression (Table 1),
progression. however the underlying molecular mechanisms are still in part
elusive. Recent reports suggest an important role of this transcription
factor in the maintenance of the cancer stem phenotype, those are the
2.4. Other cancers fraction of cells within a tumor that are capable of propagating it upon
transplantation into immuno-compromised mice [125]. In cells from
p63 is expressed in the basal layer of esophageal mucosa and plays a head and neck cancers ΔNp63 controls expression levels of CD44 and
role in its development, it is also strongly expressed in nuclei of basal layer keratins 6A and 14, known markers of more immature
squamous esophageal cancer cells (Table 1), mainly as the ΔN isoform precursors, while it strongly reduces supra-basal keratins 4 and 19
[73–83]. While p63 shows normal expression in esophageal mucosa expression suggesting its role in maintaining a less differentiated
collected from areas distant from the tumor its expression increases in phenotype [126]. In addition ΔNp63 induces expression of Inhibitor of
the basal layer of normal mucosa adjacent to the tumor [74] and is over differentiation 3 (Id-3) that is a transcriptional repressor of E-Caderin
expressed through the entire thickness of dysplastic lesions [74,78], which in turn results in inhibition of differentiation of cancer cells,
suggesting that this could be an early event in initial stages of again contributing to a stem cell like phenotype [127]. Moreover
carcinogenesis. The finding that p63 is also weakly up-regulated in Zucchi and co-worker show that p63 is essential for maintaining
Barret's esophagous supports a similar role in the development of cancer cell stemness in a population of breast cancer cells that are
adenocarcinomas [77,78]. p63 over-expression in esophageal tumors those capable of reforming tumors after transplantation in rats while
however does not correlate with prognosis while its absence has been the subpopulation that lacks p63 fails to do so [32]. Similarly it has
associated with poorer prognosis [75,76,81,82]. recently been shown that ΔNp63 is essential for the maintenance of
p63 is essential for the development of the ventral part of the urinary the basal phenotype of primary breast epithelial cells, Notch
bladder and is usually expressed in basal and supra-basal layers of the expression would down-regulate p63 allowing differentiation [128].
urothelium but not in the outer more differentiated umbrella layer As the other members of the family p63 affect the cell cycle, in
[41,84–88]. Bladder urothelial carcinoma cells (or sarcomatoid ones) particular ΔNp63 over-expression is known to result in increased cell
usually express p63 where it seems to play a role in early steps of proliferation and altered response to growth arrest signals [129], this
carcinogenesis while its expression decreases with tumor size and is probably due to inhibition of Cyclin Kinase Inhibitor (CKI) p21 and
advanced tumor stages [41,85–92]. While ΔN isoform seems to be the p57 [130]. Moreover ΔNp63 exerts a dominant negative effect on the
predominantly expressed isoform decreasing in higher grade tumors TA form of p63 and on the other family members: p53 and TAp73,
[86–88,90,93], more dramatic changes are observed for the TA isoform therefore protecting cells from cell death [42,50,131]. Indeed Leong
that shows a drastic reduction in larger less differentiated tumors, and co-workers have shown that in triple negative breast cancers
suggesting again that an imbalance between the TA and ΔN isoforms ΔNp63 blocks TAp73 dependent apoptosis, cis-platin would be able to
rather than the reduction of one particular isoform might play a role in kill these cells in part by increasing c-abl-dependent phosphorylation
this type of cancer. Indeed Park and co-workers show that while there is of TAp73 thus disrupting the interaction between p63 and p73 [30]
no correlation between ΔN levels and patient's prognosis TA reduction is and allowing p73-dependent apoptosis to occur. Moreover cis-platin
strongly correlated to poor prognosis in terms of cumulative survival induces proteosomal-mediated ΔNp63 degradation in head and neck
[88]. Interestingly loss of p63 (particularly of ΔNp63) correlates with cancer cells, this degradation is mediated by RACK1 functioning as an
tumor invasion and metastasis and consequently with poor prognosis E3 ligase [132]. UFD2a an ubiquitin E4 ligase, that normally prevents
[85–87,93,94]. ΔNp63 degradation is also down-regulated upon cis-platin treatment
Prostate requires p63 expression for its development and it is thus allowing ΔNp63 degradation [62]. In addition to a direct
expressed like in breast, in myoepithelial cells surrounding normal interaction with the other family members other mechanisms
acinar glands, therefore p63 is routinely used in diagnostics to through which p63 affects them have been proposed. As an example
evaluate the presence of normal basal cells thus distinguishing ΔNp63 increases Heat Shock Proteins 70 (HSP70) and this results in
between benign and malign glands [37,95–103]. Similarly to what reduced apoptosis partly due to p53 sequestering [133].
reported for other tissues the main isoform expressed in the basal ΔNp63 up-regulation can also contribute to tumorigenesis through
layer of the prostate is ΔNp63 [104]. However while the majority of the activation of the β-catenin signaling pathway, through reduction of
prostate cancer are p63 negative, p63 is consistently over-expressed its phosphorylation and consequent intra-nuclear accumulation [134].
in a subgroup of basal like or poorly differentiated tumors: adenoid- ΔNp63 regulates the AKT pathway [135] protecting cells from apoptosis,
cystic [105,106], [107] and poorly differentiated adenocarcinomas this is at least in part mediated by up-regulation of Fatty Acid Synthase
[102]. Moreover Dhillon and co-workers claim that some adenocar- (FASN) an enzyme involved in long chain fatty acids metabolism, critical
cinomas of the prostate show an aberrant cytoplasmic localization of for cell survival [98].
p63 that correlates with reduced apoptosis and a poor prognosis In addition to the reported roles for the ΔNp63 in the mechanisms of
[108]. Another study suggests that the ΔN isoform is the one mostly carcinogenesis recent data show that TAp63 acts as a tumor suppressor.
expressed in these tumors [109]. Indeed selective TAp63 mice KO models show increased number of
In addition to the tumors reported above scattered reports suggest a primary and metastatic tumors [136,137]. This would be due to the
role of p63 in other tumors (Table 1) such as: ovary tumors [37,110–113], ability of TAp63 to induce senescence. TAp63 induced senescence is p53
liver tumors [114,115], pancreas [116,117], poorly differentiated gastric independent but requires p21 and Rb [137]. Moreover TAp63 controls
cancers [118] grade II and III meningiomas with increased cytoplasmic expression of miRs through transcriptional control of Dicer, and altered
expression in less differentiated forms [119], squamous cervical cancers expression of some of these miRs would play a direct role in
and non-Hodgkin lymphomas (NHL) [120–124]. tumorigenesis [136]. The role of p63 in senescence however appears
In some of these tumors a correlation with prognosis has been to be tissue specific, as an example in keratinocytes loss of both TA and
attempted again with apparently conflicting results. As an example ΔNp63 or TAp63 alone leads to increased senescence [138,139].
expression of p63 in NHL can correlate with better prognosis [121] as Another open question is what leads to altered expression of p63
well as with a poorer prognosis [123]. in cancer cells. Many different mechanisms both at the transcriptional
62 V. Graziano, V. De Laurenzi / Biochimica et Biophysica Acta 1816 (2011) 57–66

and post-transcriptional level have been proposed and are all 156]. It has also been shown that IKKbeta Kinase promotes ubiquitin-
potentially valid and may contribute to p63 deregulation in different dependent proteosomal degradation of ΔNp63 sensitizing cells to
tumors. chemotherapy [157]. Finally P14-ARF, induces sumoylation with
p63 expression can be transcriptionally regulated and a number of SUMO-2 and degradation of ΔNp63 in keratinocytes [158].
transcription factors are known to regulate the p63 promoters, some
of these are known players in cancer formation. Notch decreases 4. Role of p63 in metastasis
ΔNp63 expression [128,140] and this seems to play a role in deter-
mining the luminal or basal phenotype in mammary cells [128]. While most of the work reported above seems to suggest that
TAp63 promoter contains two different binding sites for NF-kB and its altered levels of p63 are involved in cancer progression and can
down-regulation leads to inhibition of TA expression [141]. C/EBP influence tumor growth in situ and local invasion, the role of this gene
binds ΔNp63 and activates its transcription [142]. ΔNp63 has also on the ability of the tumor to form metastasis seems to be completely
been shown to be under the control of the PI3K pathway in epithelial different. Indeed a loss of all p63 isoforms seems to play an essential
cells. This pathway is often altered in human cancer, thus suggesting role in this phenomenon.
that in some cases altered expression of ΔNp63 could be a conse- Clinical studies show that in different types of cancer including:
quence of increased PI3K activity [143]. In addition TA and ΔNp63 head and neck, lung, esophageal and bladder, expression of p63
can regulate their own promoters affecting their own expression correlates to reduced invasiveness and metastasization while its loss
[144,145]. promotes invasion [42,46,55,72,76,82,85–88,93,94,159,160], more-
Recently it has been shown that miR-203 represses ΔNp63 playing an over it has recently been reported that TAp63 selective KO results in
essential role in controlling epidermal stemness and differentiation. This increased metastasis formation in mice [136]. However in breast
opens the possibility that altered miR-203 expression can result cancer this is more controversial, since some authors suggest that over
in altered p63 expression in cancer cells. Interestingly miR-203 is up- expression of p63 is associated to increased ability to form metastasis
regulated in squamous cell carcinoma cells in response to UVC [145–147]. [19,22,33] while others report that loss of p63 is associated with
Other miRs have been shown to regulate p63 such as miR-302 and miR- invasion [161].
92 in testicular cancer and in myeloid cells respectively [148,149]. The role of p63 in metastasis is probably linked to its role in
Multiple mechanisms leading to increased p63 stability through regulating cell adhesion [162] as well as to its involvement in
reduced ubiquitination have been suggested. WWP-1 an E3 ligase for epithelial–mesenchymal transition (EMT) that is considered the
ΔN and TA p63 is over-expressed in many forms of prostate and breast initial step required for cells to leave the original site and form
cancer explaining why in the majority of these cancers p63 is usually metastasis. In vitro studies suggest that disruption of ΔNp63 leads to
low. Down-modulation of this protein in cancer cell lines has opposite expression of mesenchymal and neuronal markers, several of which
effects, depending on the p63 isoform expressed by the cell. Where its are associated with the metastatic phenotype in vivo [160]. Indeed it
down regulation results in ΔN increase it sensitizes cells to drug induced has been shown that snail down-regulates ΔNp63 expression and this
death while where it results in TA increase it shows the opposite effect in turn results in a more invasive phenotype [163]. Down regulation of
[35,150]. ΔNp63 would lead to invasion through reduction of Vitamin D
An interesting observation is that MDM2, that is known to lead to receptor (VDR) and Id3 and increase of the Metallo Proteinase 2
ubiquitinian degradation of p53, seems to increase the steady state (MMP) [127,159]. Moreover, p63 would induce the expression of
levels and transactivation function of p63 as well as p73 [151]. In a tissue inhibitor of matrix metalloproteinase 1 (TIMP1) reducing the
similar way SCF (beta TrCP1) that binds and ubiquitylates p63 results invasive ability of the cells [22]. In addition Adorno and co-workers
in its stabilization [152]. Degradation of p63 is also mediated by Itch, demonstrate that p63 is not lost or mutated in cells that acquire a
another E3 ligase, this pathway is inhibited by Nedd4-binding partner metastatic phenotype, but more likely is turned off possibly through
1 (N4BP1) again an alteration of this degradative pathway, known the formation of a ternary complex with Smad and mutant p53
to regulate chemosensitivity, could lead to increased p63 levels [153– induced by TGFβ-signaling [161,164]. In this model p63 could be

Fig. 2. Model describing p63 expression in epithelial cancers. In normal epithelia p63 is mostly expressed in the nuclei of basal cells. In hyperplasia expression is also found in supra-basal
layers. Expression increases in dysplastic and cancers cells, however EMT is favored if p63 is turned off. Secondary cancers might re-express p63 allowing growth of metastatic cells.
V. Graziano, V. De Laurenzi / Biochimica et Biophysica Acta 1816 (2011) 57–66 63

reactivated after cells have left the primary tumor site and favor self- [22] A. Ribeiro-Silva, H. Becker de Moura, F. Ribeiro do Vale, S. Zucoloto, The differential
regulation of human telomerase reverse transcriptase and vascular endothelial
renewal and cancer stem cells regeneration and growth of metastatic growth factor may contribute to the clinically more aggressive behavior of p63-
tumors. In line with this idea p63 has been found over-expressed in positive breast carcinomas, Int. J. Biol. Markers 20 (4) (2005) 227–234.
secondary sites of primary p63 positive tumors [56,111,165–167]. [23] L. Hanker, T. Karn, E. Ruckhaeberle, R. Gaetje, C. Solbach, M. Schmidt, et al.,
Clinical relevance of the putative stem cell marker p63 in breast cancer, Breast
Cancer Res. Treat. 122 (3) (2009) 765–775 Epub 2009 Nov 7.
5. Concluding remarks [24] A. Ribeiro-Silva, L.N. Zamzelli Ramalho, S.B. Garcia, S. Zucoloto, Is p63 reliable in
detecting microinvasion in ductal carcinoma in situ of the breast? Pathol. Oncol.
Res. 9 (1) (2003) 20–23.
While far from having a clear picture of how p63 altered expression [25] E.A. Rakha, T.C. Putti, D.M. Abd El-Rehim, C. Paish, A.R. Green, D.G. Powe, et al.,
is involved in cancer progression a pattern starts to appear. Indeed the Morphological and immunophenotypic analysis of breast carcinomas with basal
and myoepithelial differentiation, J. Pathol. 208 (4) (2006) 495–506.
majority of data would support the idea that increase of ΔNp63 and or [26] H. Liu, Q. Fan, Z. Zhang, X. Li, H. Yu, F. Meng, Basal-HER2 phenotype shows poorer
TAp63 reduction results in a more aggressive less differentiated tumor survival than basal-like phenotype in hormone receptor-negative invasive
cell as well as in increased resistance to treatment and thus correlates breast cancers, Hum. Pathol. 39 (2) (2008) 167–174.
[27] J.S. Reis-Filho, F.C. Schmitt, p63 expression in sarcomatoid/metaplastic carcino-
with a worse prognosis. However metastasis formation would require
mas of the breast, Histopathology 42 (1) (2003) 94–95.
down-regulation of all p63 isoforms in at least a subpopulation of cells [28] M.G. Mastropasqua, E. Maiorano, G. Pruneri, E. Orvieto, G. Mazzarol, A.R. Vento,
(Fig. 2). In this respect cancers that are already p63 negative would form et al., Immunoreactivity for c-kit and p63 as an adjunct in the diagnosis of
metastasis more frequently and earlier on and therefore have a poorer adenoid cystic carcinoma of the breast, Mod. Pathol. 18 (10) (2005) 1277–1282.
[29] D. Stefanou, A. Batistatou, A. Nonni, E. Arkoumani, N.J. Agnantis, p63 expression in
outcome. This apparently opposite role can explain why often benign and malignant breast lesions, Histol. Histopathol. 19 (2) (2004) 465–471.
prognostic studies are controversial. [30] C.O. Leong, N. Vidnovic, M.P. DeYoung, D. Sgroi, L.W. Ellisen, The p63/p73
network mediates chemosensitivity to cisplatin in a biologically defined subset
of primary breast cancers, J. Clin. Invest. 117 (5) (2007) 1370–1380.
References [31] J. DiRenzo, S. Signoretti, N. Nakamura, R. Rivera-Gonzalez, W. Sellers, M. Loda,
et al., Growth factor requirements and basal phenotype of an immortalized
[1] G. Melino, X. Lu, M. Gasco, T. Crook, R.A. Knight, Functional regulation of p73 and mammary epithelial cell line, Cancer Res. 62 (1) (2002) 89–98.
p63: development and cancer, Trends Biochem. Sci. 28 (12) (2003) 663–670. [32] I. Zucchi, S. Astigiano, G. Bertalot, S. Sanzone, C. Cocola, P. Pelucchi, et al., Distinct
[2] V. De Laurenzi, G. Melino, Evolution of functions within the p53/p63/p73 family, populations of tumor-initiating cells derived from a tumor generated by rat
Ann. N.Y. Acad. Sci. 926 (2000) 90–100. mammary cancer stem cells, Proc. Natl. Acad. Sci. USA 105 (44) (2008) 16940–16945.
[3] V. De Laurenzi, A. Rossi, A. Terrinoni, D. Barcaroli, M. Levrero, A. Costanzo, et al., [33] S. Garcia, J.P. Dales, E. Charafe-Jauffret, S. Carpentier-Meunier, L. Andrac-Meyer, J.
P63 and p73 transactivate differentiation gene promoters in human keratino- Jacquemier, et al., Poor prognosis in breast carcinomas correlates with increased
cytes, Biochem. Biophys. Res. Commun. 273 (1) (2000) 342–346. expression of targetable CD146 and c-Met and with proteomic basal-like
[4] M. Levrero, V. De Laurenzi, A. Costanzo, J. Gong, G. Melino, J.Y. Wang, Structure, phenotype, Hum. Pathol. 38 (6) (2007) 830–841.
function and regulation of p63 and p73, Cell Death Differ. 6 (12) (1999) [34] Thike AA, Cheok PY, Jara-Lazaro AR, Tan B, Tan P, Tan PH. Triple-negative breast
1146–1153. cancer: clinicopathological characteristics and relationship with basal-like
[5] M. Levrero, V. De Laurenzi, A. Costanzo, J. Gong, J.Y. Wang, G. Melino, The p53/ breast cancer. Mod Pathol 23 (1) (2010) 123–133.
p63/p73 family of transcription factors: overlapping and distinct functions, J. Cell [35] Y. Li, Z. Zhou, C. Chen, WW domain-containing E3 ubiquitin protein ligase 1
Sci. 113 (Pt 10) (2000) 1661–1670. targets p63 transcription factor for ubiquitin-mediated proteasomal degradation
[6] G. Melino, V. De Laurenzi, K.H. Vousden, p73: friend or foe in tumorigenesis, Nat. and regulates apoptosis, Cell Death Differ. 15 (12) (2008) 1941–1951.
Rev. Cancer 2 (8) (2002) 605–615. [36] A. Rocca, G. Viale, R.D. Gelber, L. Bottiglieri, S. Gelber, G. Pruneri, et al., Pathologic
[7] M.K. Wetzel, S. Naska, C.L. Laliberte, V.V. Rymar, M. Fujitani, J.A. Biernaskie, et al., complete remission rate after cisplatin-based primary chemotherapy in breast
p73 regulates neurodegeneration and phospho-tau accumulation during aging cancer: correlation with p63 expression, Cancer Chemother. Pharmacol. 61 (6)
and Alzheimer's disease, Neuron 59 (5) (2008) 708–721. (2008) 965–971.
[8] A. Yang, N. Walker, R. Bronson, M. Kaghad, M. Oosterwegel, J. Bonnin, et al., p73- [37] J.S. Reis-Filho, P.T. Simpson, A. Martins, A. Preto, F. Gartner, F.C. Schmitt,
deficient mice have neurological, pheromonal and inflammatory defects but lack Distribution of p63, cytokeratins 5/6 and cytokeratin 14 in 51 normal and 400
spontaneous tumours, Nature 404 (6773) (2000) 99–103. neoplastic human tissue samples using TARP-4 multi-tumor tissue microarray,
[9] Wilhelm MT, Rufini A, Wetzel MK, Tsuchihara K, Inoue S, Tomasini R, et al. Virchows Arch. 443 (2) (2003) 122–132.
Isoform-specific p73 knockout mice reveal a novel role for delta Np73 in the DNA [38] M. Wu, A. Kafanas, L. Gan, D.S. Kohtz, L. Zhang, E. Genden, et al., Feasibility of
damage response pathway. Genes Dev. 24 (6) (2008) 549–560. immunocytochemical detection of tumor markers (XIAP, phosphohistone H1
[10] R. Tomasini, K. Tsuchihara, M. Wilhelm, M. Fujitani, A. Rufini, C.C. Cheung, et al., and p63) in FNA cellblock samples from head and neck squamous cell carcinoma,
TAp73 knockout shows genomic instability with infertility and tumor suppres- Diagn. Cytopathol. 36 (11) (2008) 797–800.
sor functions, Genes Dev. 22 (19) (2008) 2677–2691. [39] T.D. Bourne, A.M. Bellizzi, E.B. Stelow, A.H. Loy, P.A. Levine, M.R. Wick, et al., p63
[11] A. Yang, R. Schweitzer, D. Sun, M. Kaghad, N. Walker, R.T. Bronson, et al., p63 is expression in olfactory neuroblastoma and other small cell tumors of the
essential for regenerative proliferation in limb, craniofacial and epithelial sinonasal tract, Am. J. Clin. Pathol. 130 (2) (2008) 213–218.
development, Nature 398 (6729) (1999) 714–718. [40] D.E. Burstein, C. Nagi, D.S. Kohtz, L. Lee, B. Wang, Immunodetection of GLUT1,
[12] A.A. Mills, B. Zheng, X.J. Wang, H. Vogel, D.R. Roop, A. Bradley, p63 is a p53 p63 and phospho-histone H1 in invasive head and neck squamous carcinoma:
homologue required for limb and epidermal morphogenesis, Nature 398 (6729) correlation of immunohistochemical staining patterns with keratinization,
(1999) 708–713. Histopathology 48 (6) (2006) 717–722.
[13] E. Candi, R. Cipollone, P. Rivetti di Val Cervo, S. Gonfloni, G. Melino, R. Knight, p63 [41] J.S. Lewis, J.H. Ritter, S. El-Mofty, Alternative epithelial markers in sarcomatoid
in epithelial development, Cell. Mol. Life Sci. 65 (20) (2008) 3126–3133. carcinomas of the head and neck, lung, and bladder-p63, MOC-31, and TTF-1,
[14] E. Candi, D. Dinsdale, A. Rufini, P. Salomoni, R.A. Knight, M. Mueller, et al., TAp63 Mod. Pathol. 18 (11) (2005) 1471–1481.
and DeltaNp63 in cancer and epidermal development, Cell Cycle 6 (3) (2007) [42] J.W. Rocco, C.O. Leong, N. Kuperwasser, M.P. DeYoung, L.W. Ellisen, p63 mediates
274–285. survival in squamous cell carcinoma by suppression of p73-dependent
[15] E. Candi, A. Rufini, A. Terrinoni, D. Dinsdale, M. Ranalli, A. Paradisi, et al., apoptosis, Cancer Cell 9 (1) (2006) 45–56.
Differential roles of p63 isoforms in epidermal development: selective genetic [43] K. Nylander, P.J. Coates, P.A. Hall, Characterization of the expression pattern of
complementation in p63 null mice, Cell Death Differ. 13 (6) (2006) 1037–1047. p63 alpha and delta Np63 alpha in benign and malignant oral epithelial lesions,
[16] E.R. Flores, S. Sengupta, J.B. Miller, J.J. Newman, R. Bronson, D. Crowley, et al., Int. J. Cancer 87 (3) (2000) 368–372.
Tumor predisposition in mice mutant for p63 and p73: evidence for broader [44] P. Saintigny, A.K. El-Naggar, V. Papadimitrakopoulou, H. Ren, Y.H. Fan, L. Feng,
tumor suppressor functions for the p53 family, Cancer Cell 7 (4) (2005) 363–373. et al., DeltaNp63 overexpression, alone and in combination with other
[17] W.M. Keyes, H. Vogel, M.I. Koster, X. Guo, Y. Qi, K.M. Petherbridge, et al., P63 biomarkers, predicts the development of oral cancer in patients with
heterozygous mutant mice are not prone to spontaneous or chemically induced leukoplakia, Clin. Cancer Res. 15 (19) (2009) 6284–6291.
tumors, Proc. Natl. Acad. Sci. USA 103 (22) (2006) 8435–8440. [45] A. Weber, U. Bellmann, F. Bootz, C. Wittekind, A. Tannapfel, Expression of p53
[18] M.I. Koster, D. Dai, D.R. Roop, Conflicting roles for p63 in skin development and and its homologues in primary and recurrent squamous cell carcinomas of the
carcinogenesis, Cell Cycle 6 (3) (2007) 269–273. head and neck, Int. J. Cancer 99 (1) (2002) 22–28.
[19] A. Ribeiro-Silva, L.N. Zambelli Ramalho, S. Britto Garcia, S. Zucoloto, The [46] P.P. Massion, P.M. Taflan, S.M. Jamshedur Rahman, P. Yildiz, Y. Shyr, M.E.
relationship between p63 and p53 expression in normal and neoplastic breast Edgerton, et al., Significance of p63 amplification and overexpression in lung
tissue, Arch. Pathol. Lab. Med. 127 (3) (2003) 336–340. cancer development and prognosis, Cancer Res. 63 (21) (2003) 7113–7121.
[20] M. Barbareschi, L. Pecciarini, M.G. Cangi, E. Macri, A. Rizzo, G. Viale, et al., p63, a [47] P.P. Massion, P.M. Taflan, S.M. Rahman, P. Yildiz, Y. Shyr, D.P. Carbone, et al., Role
p53 homologue, is a selective nuclear marker of myoepithelial cells of the human of p63 amplification and overexpression in lung cancer development, Chest 125
breast, Am. J. Surg. Pathol. 25 (8) (2001) 1054–1060. (5 Suppl) (2004) 102S.
[21] K. Nylander, B. Vojtesek, R. Nenutil, B. Lindgren, G. Roos, W. Zhanxiang, et al., [48] N. Thurfjell, P.J. Coates, B. Vojtesek, P. Benham-Motlagh, M. Eisold, K. Nylander,
Differential expression of p63 isoforms in normal tissues and neoplastic cells, J. Endogenous p63 acts as a survival factor for tumour cells of SCCHN origin, Int. J.
Pathol. 198 (4) (2002) 417–427. Mol. Med. 16 (6) (2005) 1065–1070.
64 V. Graziano, V. De Laurenzi / Biochimica et Biophysica Acta 1816 (2011) 57–66

[49] K. Hibi, B. Trink, M. Patturajan, W.H. Westra, O.L. Caballero, D.E. Hill, et al., AIS is [78] P.A. Hall, A.C. Woodman, S.J. Campbell, N.A. Shepherd, Expression of the p53
an oncogene amplified in squamous cell carcinoma, Proc. Natl. Acad. Sci. USA 97 homologue p63alpha and DeltaNp63alpha in the neoplastic sequence of Barrett's
(10) (2000) 5462–5467. oesophagus: correlation with morphology and p53 protein, Gut 49 (5) (2001)
[50] T. Crook, J.M. Nicholls, L. Brooks, J. O'Nions, M.J. Allday, High level expression of 618–623.
deltaN-p63: a mechanism for the inactivation of p53 in undifferentiated [79] P. Taniere, G. Martel-Planche, J.C. Saurin, C. Lombard-Bohas, F. Berger, J.Y.
nasopharyngeal carcinoma (NPC)? Oncogene 19 (30) (2000) 3439–3444. Scoazec, et al., TP53 mutations, amplification of P63 and expression of cell cycle
[51] S. Asioli, A. Righi, M. Volante, V. Eusebi, G. Bussolati, p63 expression as a new proteins in squamous cell carcinoma of the oesophagus from a low incidence
prognostic marker in Merkel cell carcinoma, Cancer 110 (3) (2007) 640–647. area in Western Europe, Br. J. Cancer 85 (5) (2001) 721–726.
[52] J.C. Sniezek, K.E. Matheny, M.D. Westfall, J.A. Pietenpol, Dominant negative p63 [80] J.N. Glickman, A. Yang, A. Shahsafaei, F. McKeon, R.D. Odze, Expression of p53-
isoform expression in head and neck squamous cell carcinoma, Laryngoscope related protein p63 in the gastrointestinal tract and in esophageal metaplastic
114 (12) (2004) 2063–2072. and neoplastic disorders, Hum. Pathol. 32 (11) (2001) 1157–1165.
[53] F.C. Xavier, C.M. Takiya, S.R. Reis, L.M. Ramalho, p63 immunoexpression in lip [81] T. Hara, H. Kijima, S. Yamamoto, T. Kenmochi, Y. Kise, H. Tanaka, et al., Ubiquitous
carcinogenesis, J. Mol. Histol. 40 (2) (2009) 131–137. p63 expression in human esophageal squamous cell carcinoma, Int. J. Mol. Med.
[54] M. Ebrahimi, L. Boldrup, Y.B. Wahlin, P.J. Coates, K. Nylander, Decreased 14 (2) (2004) 169–173.
expression of the p63 related proteins beta-catenin, E-cadherin and EGFR in [82] M. Morita, H. Uramoto, S. Nakata, K. Ono, M. Sugaya, T. Yoshimatsu, et al.,
oral lichen planus, Oral Oncol. 44 (7) (2008) 634–638. Expression of deltaNp63 in squamous cell carcinoma of the esophagus,
[55] R. Zangen, E. Ratovitski, D. Sidransky, DeltaNp63alpha levels correlate with Anticancer Res. 25 (5) (2005) 3533–3539.
clinical tumor response to cisplatin, Cell Cycle 4 (10) (2005) 1313–1315. [83] A. Thépot, A. Hautefeuille, M.P. Cros, B. Abedi-Ardekani, A. Pétré, O. Damour, V.
[56] L.R. Oliveira, A. Ribeiro-Silva, S. Zucoloto, Prognostic significance of p53 and p63 Krutovskikh, P. Hainaut, Intraepithelial p63-dependent expression of distinct
immunolocalisation in primary and matched lymph node metastasis in oral components of cell adhesion complexes in normal esophageal mucosa and
squamous cell carcinoma, Acta Histochem. 109 (5) (2007) 388–396. squamous cell carcinoma. Int. J. Cancer 127 (9) (2010) 2051–2062.
[57] L. Lo Muzio, A. Santarelli, R. Caltabiano, C. Rubini, T. Pieramici, L. Trevisiol, p63 [84] W. Cheng, W.B. Jacobs, J.J. Zhang, A. Moro, J.H. Park, M. Kushida, et al., DeltaNp63
overexpression associates with poor prognosis in head and neck squamous cell plays an anti-apoptotic role in ventral bladder development, Development 133
carcinoma, Hum. Pathol. 36 (2) (2005) 187–194. (23) (2006) 4783–4792.
[58] M.P. Foschini, R. Cocchi, L. Morandi, G. Marucci, M.G. Pennesi, A. Righi, et al., E- [85] E. Comperat, P. Camparo, R. Haus, E. Chartier-Kastler, S. Bart, A. Delcourt, et al.,
cadherin loss and Delta Np73L expression in oral squamous cell carcinomas Immunohistochemical expression of p63, p53 and MIB-1 in urinary bladder
showing aggressive behavior, Head Neck 30 (11) (2008) 1475–1482. carcinoma. A tissue microarray study of 158 cases, Virchows Arch. 448 (3)
[59] G. Pruneri, L. Pignataro, M. Manzotti, N. Carboni, D. Ronchetti, A. Neri, et al., p63 (2006) 319–324.
in laryngeal squamous cell carcinoma: evidence for a role of TA-p63 down- [86] F. Koga, S. Kawakami, Y. Fujii, K. Saito, Y. Ohtsuka, A. Iwai, et al., Impaired p63
regulation in tumorigenesis and lack of prognostic implications of p63 expression associates with poor prognosis and uroplakin III expression in invasive
immunoreactivity, Lab. Invest 82 (10) (2002) 1327–1334. urothelial carcinoma of the bladder, Clin. Cancer Res. 9 (15) (2003) 5501–5507.
[60] J.W. Rocco, L.W. Ellisen, p63 and p73: life and death in squamous cell carcinoma, [87] F. Koga, S. Kawakami, J. Kumagai, T. Takizawa, N. Ando, G. Arai, et al., Impaired
Cell Cycle 5 (9) (2006) 936–940. Delta Np63 expression associates with reduced beta-catenin and aggressive
[61] N. Thurfjell, P.J. Coates, L. Boldrup, B. Lindgren, B. Backlund, T. Uusitalo, et al., phenotypes of urothelial neoplasms, Br. J. Cancer 88 (5) (2003) 740–747.
Function and importance of p63 in normal oral mucosa and squamous cell [88] B.J. Park, S.J. Lee, J.I. Kim, C.H. Lee, S.G. Chang, J.H. Park, et al., Frequent alteration
carcinoma of the head and neck, Adv. Otorhinolaryngol 62 (2005) 49–57. of p63 expression in human primary bladder carcinomas, Cancer Res. 60 (13)
[62] A. Chatterjee, S. Upadhyay, X. Chang, J.K. Nagpal, B. Trink, D. Sidransky, U-box- (2000) 3370–3374.
type ubiquitin E4 ligase, UFD2a attenuates cisplatin mediated degradation of [89] C. Langner, M. Ratschek, O. Tsybrovskyy, L. Schips, R. Zigeuner, P63 immuno-
DeltaNp63alpha, Cell Cycle 7 (9) (2008) 1231–1237. reactivity distinguishes upper urinary tract transitional-cell carcinoma and
[63] L. Boldrup, P.J. Coates, X. Gu, K. Nylander, DeltaNp63 isoforms differentially renal-cell carcinoma even in poorly differentiated tumors, J. Histochem.
regulate gene expression in squamous cell carcinoma: identification of Cox-2 as Cytochem. 51 (8) (2003) 1097–1099.
a novel p63 target, J. Pathol. 218 (4) (2009) 428–436. [90] Y. He, X. Wu, W. Tang, D. Tian, C. Luo, Z. Yin, et al., Impaired delta NP63
[64] M. Wu, B. Wang, J. Gil, E. Sabo, L. Miller, L. Gan, et al., p63 and TTF-1 expression is associated with poor tumor development in transitional cell
immunostaining. A useful marker panel for distinguishing small cell carcinoma carcinoma of the bladder, J. Korean Med. Sci. 23 (5) (2008) 825–832.
of lung from poorly differentiated squamous cell carcinoma of lung, Am. J. Clin. [91] C.J. Di Como, M.J. Urist, I. Babayan, M. Drobnjak, C.V. Hedvat, J. Teruya-Feldstein,
Pathol. 119 (5) (2003) 696–702. et al., p63 expression profiles in human normal and tumor tissues, Clin. Cancer
[65] M. Wu, L. Orta, J. Gil, G. Li, A. Hu, D.E. Burstein, Immunohistochemical detection Res. 8 (2) (2002) 494–501.
of XIAP and p63 in adenomatous hyperplasia, atypical adenomatous hyperplasia, [92] Buza N, Cohen PJ, Pei H, Parkash V. Inverse p16 and p63 expression in small cell
bronchioloalveolar carcinoma and well-differentiated adenocarcinoma, Mod. carcinoma and high-grade urothelial cell carcinoma of the urinary bladder. Int J
Pathol. 21 (5) (2008) 553–558. Surg Pathol 18 (2) (2010) 94–102.
[66] T. Narahashi, T. Niki, T. Wang, A. Goto, D. Matsubara, N. Funata, et al., Cytoplasmic [93] H. Fukushima, F. Koga, S. Kawakami, Y. Fujii, S. Yoshida, E. Ratovitski, et al., Loss
localization of p63 is associated with poor patient survival in lung adenocar- of DeltaNp63alpha promotes invasion of urothelial carcinomas via N-cadherin/
cinoma, Histopathology 49 (4) (2006) 349–357. Src homology and collagen/extracellular signal-regulated kinase pathway,
[67] B.Y. Wang, J. Gil, D. Kaufman, L. Gan, D.S. Kohtz, D.E. Burstein, P63 in pulmonary Cancer Res. 69 (24) (2009) 9263–9270.
epithelium, pulmonary squamous neoplasms, and other pulmonary tumors, [94] R. Zigeuner, O. Tsybrovskyy, M. Ratschek, P. Rehak, K. Lipsky, C. Langner,
Hum. Pathol. 33 (9) (2002) 921–926. Prognostic impact of p63 and p53 expression in upper urinary tract transitional
[68] G. Pelosi, F. Pasini, C. Olsen Stenholm, U. Pastorino, P. Maisonneuve, A. Sonzogni, cell carcinoma, Urology 63 (6) (2004) 1079–1083.
et al., p63 immunoreactivity in lung cancer: yet another player in the [95] S. Signoretti, D. Waltregny, J. Dilks, B. Isaac, D. Lin, L. Garraway, et al., p63 is a
development of squamous cell carcinomas? J. Pathol. 198 (1) (2002) 100–109. prostate basal cell marker and is required for prostate development, Am. J.
[69] H. Zhang, J. Liu, P.T. Cagle, T.C. Allen, A.C. Laga, D.S. Zander, Distinction of pulmonary Pathol. 157 (6) (2000) 1769–1775.
small cell carcinoma from poorly differentiated squamous cell carcinoma: an [96] M.H. Weinstein, S. Signoretti, M. Loda, Diagnostic utility of immunohistochem-
immunohistochemical approach, Mod. Pathol. 18 (1) (2005) 111–118. ical staining for p63, a sensitive marker of prostatic basal cells, Mod. Pathol. 15
[70] N.H. Au, A.M. Gown, M. Cheang, D. Huntsman, E. Yorida, W.M. Elliott, et al., P63 (12) (2002) 1302–1308.
expression in lung carcinoma: a tissue microarray study of 408 cases, Appl. [97] A. Kuzmanov, S. Hayrabedyan, M. Karaivanov, K. Todorova, Basal cell
Immunohistochem. Mol. Morphol. 12 (3) (2004) 240–247. subpopulation as putative human prostate carcinoma stem cells, Folia
[71] T. Iwata, H. Uramoto, K. Sugio, Y. Fujino, T. Oyama, S. Nakata, et al., A lack of Histochem. Cytobiol. 45 (2) (2007) 75–80.
prognostic significance regarding DeltaNp63 immunoreactivity in lung cancer, [98] V. Sabbisetti, A. Di Napoli, A. Seeley, A.M. Amato, E. O'Regan, M. Ghebremichael,
Lung Cancer 50 (1) (2005) 67–73. et al., p63 promotes cell survival through fatty acid synthase, PLoS One 4 (6)
[72] N.H. Au, M. Cheang, D.G. Huntsman, E. Yorida, A. Coldman, W.M. Elliott, et al., (2009) e5877.
Evaluation of immunohistochemical markers in non-small cell lung cancer by [99] Garraway IP, Sun W, Tran CP, Perner S, Zhang B, Goldstein AS, et al. Human
unsupervised hierarchical clustering analysis: a tissue microarray study of 284 prostate sphere-forming cells represent a subset of basal epithelial cells capable
cases and 18 markers, J. Pathol. 204 (1) (2004) 101–109. of glandular regeneration in vivo. Prostate 70 (5) 491–501.
[73] Y. Daniely, G. Liao, D. Dixon, R.I. Linnoila, A. Lori, S.H. Randell, et al., Critical role of [100] M. Zhou, R. Shah, R. Shen, M.A. Rubin, Basal cell cocktail (34betaE12 + p63)
p63 in the development of a normal esophageal and tracheobronchial improves the detection of prostate basal cells, Am. J. Surg. Pathol. 27 (3) (2003)
epithelium, Am. J. Physiol. Cell Physiol. 287 (1) (2004) C171–C181. 365–371.
[74] H. Hu, S.H. Xia, A.D. Li, X. Xu, Y. Cai, Y.L. Han, et al., Elevated expression of p63 protein in [101] L.D. Davis, W. Zhang, A. Merseburger, D. Young, L. Xu, J.S. Rhim, et al., p63
human esophageal squamous cell carcinomas, Int. J. Cancer 102 (6) (2002) 580–583. expression profile in normal and malignant prostate epithelial cells, Anticancer
[75] L.Y. Cao, Y. Yin, H. Li, Y. Jiang, H.F. Zhang, Expression and clinical significance of Res. 22 (6C) (2002) 3819–3825.
S100A2 and p63 in esophageal carcinoma, World J. Gastroenterol. 15 (33) [102] J.K. Parsons, W.R. Gage, W.G. Nelson, A.M. De Marzo, p63 protein expression is
(2009) 4183–4188. rare in prostate adenocarcinoma: implications for cancer diagnosis and
[76] Y. Takahashi, T. Noguchi, S. Takeno, Y. Kimura, M. Okubo, K. Kawahara, Reduced carcinogenesis, Urology 58 (4) (2001) 619–624.
expression of p63 has prognostic implications for patients with esophageal [103] O. Hameed, J. Sublett, P.A. Humphrey, Immunohistochemical stains for p63 and
squamous cell carcinoma, Oncol. Rep. 15 (2) (2006) 323–328. alpha-methylacyl-CoA racemase, versus a cocktail comprising both, in the
[77] H. Geddert, S. Kiel, H.J. Heep, H.E. Gabbert, M. Sarbia, The role of p63 and diagnosis of prostatic carcinoma: a comparison of the immunohistochemical
deltaNp63 (p40) protein expression and gene amplification in esophageal staining of 430 foci in radical prostatectomy and needle biopsy tissues, Am. J.
carcinogenesis, Hum. Pathol. 34 (9) (2003) 850–856. Surg. Pathol. 29 (5) (2005) 579–587.
V. Graziano, V. De Laurenzi / Biochimica et Biophysica Acta 1816 (2011) 57–66 65

[104] C. Grisanzio, S. Signoretti, p63 in prostate biology and pathology, J. Cell. Biochem. [133] G. Wu, M. Osada, Z. Guo, A. Fomenkov, S. Begum, M. Zhao, et al., DeltaNp63alpha
103 (5) (2008) 1354–1368. up-regulates the Hsp70 gene in human cancer, Cancer Res. 65 (3) (2005)
[105] K.A. Iczkowski, K.L. Ferguson, D.D. Grier, D. Hossain, S.S. Banerjee, J.E. McNeal, 758–766.
et al., Adenoid cystic/basal cell carcinoma of the prostate: clinicopathologic [134] M. Patturajan, S. Nomoto, M. Sommer, A. Fomenkov, K. Hibi, R. Zangen, et al.,
findings in 19 cases, Am. J. Surg. Pathol. 27 (12) (2003) 1523–1529. DeltaNp63 induces beta-catenin nuclear accumulation and signaling, Cancer Cell
[106] T.Z. Ali, J.I. Epstein, Basal cell carcinoma of the prostate: a clinicopathologic study 1 (4) (2002) 369–379.
of 29 cases, Am. J. Surg. Pathol. 31 (5) (2007) 697–705. [135] E. Ogawa, R. Okuyama, S. Ikawa, H. Nagoshi, T. Egawa, A. Kurihara, et al., p51/p63
[107] M.G. Mastropasqua, G. Pruneri, G. Renne, O. De Cobelli, G. Viale, Basaloid cell inhibits ultraviolet B-induced apoptosis via Akt activation, Oncogene 27 (6)
carcinoma of the prostate, Virchows Arch. 443 (6) (2003) 787–791. (2008) 848–856.
[108] P.K. Dhillon, M. Barry, M.J. Stampfer, S. Perner, M. Fiorentino, A. Fornari, et al., [136] Su X, Chakravarti D, Cho MS, Liu L, Gi YJ, Lin YL, et al. TAp63 suppresses
Aberrant cytoplasmic expression of p63 and prostate cancer mortality, Cancer metastasis through coordinate regulation of Dicer and miRNAs. Nature 467
Epidemiol. Biomarkers Prev. 18 (2) (2009) 595–600. (7318) (2010) 986–990.
[109] J.K. Parsons, E.A. Saria, M. Nakayama, R.L. Vessella, C.L. Sawyers, W.B. Isaacs, [137] X. Guo, W.M. Keyes, C. Papazoglu, J. Zuber, W. Li, S.W. Lowe, et al., TAp63 induces
et al., Comprehensive mutational analysis and mRNA isoform quantification of senescence and suppresses tumorigenesis in vivo, Nat. Cell Biol. 11 (12) (2009)
TP63 in normal and neoplastic human prostate cells, Prostate 69 (5) (2009) 1451–1457.
559–569. [138] W.M. Keyes, Y. Wu, H. Vogel, X. Guo, S.W. Lowe, A.A. Mills, p63 deficiency
[110] X.Y. Liao, W.C. Xue, D.H. Shen, H.Y. Ngan, M.K. Siu, A.N. Cheung, p63 expression in activates a program of cellular senescence and leads to accelerated aging, Genes
ovarian tumours: a marker for Brenner tumours but not transitional cell Dev. 19 (17) (2005) 1986–1999.
carcinomas, Histopathology 51 (4) (2007) 477–483. [139] X. Su, M. Paris, Y.J. Gi, K.Y. Tsai, M.S. Cho, Y.L. Lin, et al., TAp63 prevents
[111] A. Kalebi, M. Hale, p63 expression in ovarian tumours: immunopositivity in premature aging by promoting adult stem cell maintenance, Cell Stem Cell 5 (1)
metastatic transitional cell carcinoma of the ovary, Histopathology 53 (2) (2008) (2009) 64–75.
228. [140] B.C. Nguyen, K. Lefort, A. Mandinova, D. Antonini, V. Devgan, G. Della Gatta, et al.,
[112] S.L. Shao, Y. Cai, Q.H. Wang, L.J. Yan, X.Y. Zhao, L.X. Wang, Expression of GLUT-1, Cross-regulation between Notch and p63 in keratinocyte commitment to
p63 and DNA-Pkcs in serous ovarian tumors and their significance, Zhonghua differentiation, Genes Dev. 20 (8) (2006) 1028–1042.
Zhong Liu Za Zhi 29 (9) (2007) 697–700. [141] Wu J, Bergholz J, Lu J, Sonenshein GE, Xiao ZX. TAp63 is a transcriptional target of
[113] S. Marchini, M. Marabese, E. Marrazzo, P. Mariani, D. Cattaneo, R. Fossati, et al., NF-kappaB. J Cell Biochem 109 (4) (2010) 702–710.
DeltaNp63 expression is associated with poor survival in ovarian cancer, Ann. [142] S. Borrelli, B. Testoni, M. Callari, D. Alotto, C. Castagnoli, R.A. Romano, et al.,
Oncol. 19 (3) (2008) 501–507. Reciprocal regulation of p63 by C/EBP delta in human keratinocytes, BMC Mol.
[114] F.S. Ramalho, L.N. Ramalho, L. Della Porta, S. Zucoloto, Comparative immuno- Biol. 8 (2007) 85.
histochemical expression of p63 in human cholangiocarcinoma and hepatocel- [143] C.E. Barbieri, C.E. Barton, J.A. Pietenpol, Delta Np63 alpha expression is regulated
lular carcinoma, J. Gastroenterol. Hepatol. 21 (8) (2006) 1276–1280. by the phosphoinositide 3-kinase pathway, J. Biol. Chem. 278 (51) (2003)
[115] K. Nomoto, K. Tsuneyama, C. Cheng, H. Takahashi, R. Hori, Y. Murai, et al., 51408–51414.
Intrahepatic cholangiocarcinoma arising in cirrhotic liver frequently expressed [144] W.K. Chu, K.C. Lee, S.E. Chow, J.K. Chen, Dual regulation of the DeltaNp63
p63-positive basal/stem-cell phenotype, Pathol. Res. Pract. 202 (2) (2006) 71–76. transcriptional activity by DeltaNp63 in human nasopharyngeal carcinoma cell,
[116] O. Basturk, F. Khanani, F. Sarkar, E. Levi, J.D. Cheng, N.V. Adsay, DeltaNp63 Biochem. Biophys. Res. Commun. 342 (4) (2006) 1356–1360.
expression in pancreas and pancreatic neoplasia, Mod. Pathol. 18 (9) (2005) [145] N. Li, H. Li, P. Cherukuri, S. Farzan, D.C. Harmes, J. DiRenzo, TA-p63-gamma
1193–1198. regulates expression of DeltaN-p63 in a manner that is sensitive to p53,
[117] G. Marucci, C.M. Betts, L. Liguori, V. Eusebi, Basaloid carcinoma of the pancreas, Oncogene 25 (16) (2006) 2349–2359.
Virchows Arch. 446 (3) (2005) 322–324. [146] A.M. Lena, R. Shalom-Feuerstein, P. Rivetti di Val Cervo, D. Aberdam, R.A. Knight,
[118] A. Tannapfel, S. Schmelzer, M. Benicke, M. Klimpfinger, K. Kohlhaw, J. Mossner, G. Melino, et al., MiR-203 represses ‘stemness’ by repressing DeltaNp63, Cell
et al., Expression of the p53 homologues p63 and p73 in multiple simultaneous Death Differ. 15 (7) (2008) 1187–1195.
gastric cancer, J. Pathol. 195 (2) (2001) 163–170. [147] R. Yi, M.N. Poy, M. Stoffel, E. Fuchs, A skin microRNA promotes differentiation by
[119] E.J. Rushing, C. Olsen, Y.G. Man, Correlation of p63 immunoreactivity with tumor repressing ‘stemness’, Nature 452 (7184) (2008) 225–229.
grade in meningiomas, Int. J. Surg. Pathol. 16 (1) (2008) 38–42. [148] A.H. Scheel, U. Beyer, R. Agami, M. Dobbelstein, Immunofluorescence-based
[120] G. Gualco, L.M. Weiss, C.E. Bacchi, Expression of p63 in anaplastic large cell screening identifies germ cell associated microRNA 302 as an antagonist to p63
lymphoma but not in classical Hodgkin's lymphoma, Hum. Pathol. 39 (10) expression, Cell Cycle 8 (9) (2009) 1426–1432.
(2008) 1505–1510. [149] I. Manni, S. Artuso, S. Careccia, M.G. Rizzo, R. Baserga, G. Piaggio, et al., The
[121] A.E. Hallack Neto, S.A. Siqueira, F.L. Dulley, M.A. Ruiz, D.A. Chamone, J. Pereira, microRNA miR-92 increases proliferation of myeloid cells and by targeting
p63 protein expression in high risk diffuse large B-cell lymphoma, J. Clin. Pathol. p63 modulates the abundance of its isoforms, FASEB J. 23 (11) (2009)
62 (1) (2009) 77–79. 3957–3966.
[122] X.L. Zuo, Y. Zhou, X. Zhou, X.H. Liu, K.J. Zhang, H.Q. Yang, et al., Expression of [150] Peschiaroli A., Scialpi F., Bernassola F., El Sherbini el S., Melino G. The E3 ubiquitin
survivin and P63 protein in B cell non-Hodgkin's lymphoma and their effects on ligase WWP1 regulates DeltaNp63-dependent transcription through Lys63
cell apoptosis and proliferation, Zhongguo Shi Yan Xue Ye Xue Za Zhi 15 (1) linkages. Biochem Biophys Res Commun 402 (2) (2010) 425–430.
(2007) 99–102. [151] V. Calabro, G. Mansueto, T. Parisi, M. Vivo, R.A. Calogero, G. La Mantia, The human
[123] N. Fukushima, T. Satoh, N. Sueoka, A. Sato, M. Ide, T. Hisatomi, et al., Clinico- MDM2 oncoprotein increases the transcriptional activity and the protein level of
pathological characteristics of p63 expression in B-cell lymphoma, Cancer Sci. 97 the p53 homolog p63, J. Biol. Chem. 277 (4) (2002) 2674–2681.
(10) (2006) 1050–1055. [152] J.R. Gallegos, J. Litersky, H. Lee, Y. Sun, K. Nakayama, H. Lu, SCF TrCP1 activates
[124] G. Pruneri, S. Fabris, P. Dell'Orto, M.O. Biasi, S. Valentini, B. Del Curto, et al., The and ubiquitylates TAp63gamma, J. Biol. Chem. 283 (1) (2008) 66–75.
transactivating isoforms of p63 are overexpressed in high-grade follicular [153] M. Rossi, R.I. Aqeilan, M. Neale, E. Candi, P. Salomoni, R.A. Knight, et al., The E3
lymphomas independent of the occurrence of p63 gene amplification, J. Pathol. ubiquitin ligase Itch controls the protein stability of p63, Proc. Natl. Acad. Sci.
206 (3) (2005) 337–345. USA 103 (34) (2006) 12753–12758.
[125] L. Vermeulen, M.R. Sprick, K. Kemper, G. Stassi, J.P. Medema, Cancer stem cells— [154] T.M. Hansen, M. Rossi, J.P. Roperch, K. Ansell, K. Simpson, D. Taylor, et al., Itch
old concepts, new insights, Cell Death Differ. 15 (6) (2008) 947–958. inhibition regulates chemosensitivity in vitro, Biochem. Biophys. Res. Commun.
[126] L. Boldrup, P.J. Coates, X. Gu, K. Nylander, DeltaNp63 isoforms regulate CD44 and 361 (1) (2007) 33–36.
keratins 4, 6, 14 and 19 in squamous cell carcinoma of head and neck, J. Pathol. [155] A. Oberst, M. Malatesta, R.I. Aqeilan, M. Rossi, P. Salomoni, R. Murillas, et al., The
213 (4) (2007) 384–391. Nedd4-binding partner 1 (N4BP1) protein is an inhibitor of the E3 ligase Itch,
[127] K. Higashikawa, S. Yoneda, K. Tobiume, M. Saitoh, M. Taki, Y. Mitani, et al., Proc. Natl. Acad. Sci. USA 104 (27) (2007) 11280–11285.
DeltaNp63alpha-dependent expression of Id-3 distinctively suppresses the [156] G. Melino, R.A. Knight, G. Cesareni, Degradation of p63 by Itch, Cell Cycle 5 (16)
invasiveness of human squamous cell carcinoma, Int. J. Cancer 124 (12) (2006) 1735–1739.
(2009) 2837–2844. [157] Chatterjee A, Chang X, Sen T, Ravi R, Bedi A, Sidransky D. Regulation of p53 family
[128] Yalcin-Ozuysal O, Fiche M, Guitierrez M, Wagner KU, Raffoul W, Brisken C. member isoform DeltaNp63alpha by the nuclear factor-kappaB targeting kinase
Antagonistic roles of Notch and p63 in controlling mammary epithelial cell fates. IkappaB kinase beta. Cancer Res 70 (4) (2010) 1419–1429.
Cell Death Differ. 17 (10) (2010) 1600–1612. [158] M. Vivo, A. Di Costanzo, P. Fortugno, A. Pollice, V. Calabro, G. La Mantia,
[129] K.E. King, W.C. Weinberg, p63: defining roles in morphogenesis, homeostasis, Downregulation of DeltaNp63alpha in keratinocytes by p14ARF-mediated
and neoplasia of the epidermis, Mol. Carcinog. 46 (8) (2007) 716–724. SUMO-conjugation and degradation, Cell Cycle 8 (21) (2009) 3537–3543.
[130] C.T. Chiang, W.K. Chu, S.E. Chow, J.K. Chen, Overexpression of delta Np63 in a [159] R. Kommagani, M.K. Leonard, S. Lewis, R.A. Romano, S. Sinha, M.P. Kadakia,
human nasopharyngeal carcinoma cell line downregulates CKIs and enhances Regulation of VDR by deltaNp63alpha is associated with inhibition of cell
cell proliferation, J. Cell. Physiol. 219 (1) (2009) 117–122. invasion, J. Cell Sci. 122 (Pt 16) (2009) 2828–2835.
[131] Mundt HM, Stremmel W, Melino G, Krammer PH, Schilling T, Muller M. [160] C.E. Barbieri, L.J. Tang, K.A. Brown, J.A. Pietenpol, Loss of p63 leads to increased
Dominant negative (DeltaN) p63alpha induces drug resistance in hepatocellular cell migration and up-regulation of genes involved in invasion and metastasis,
carcinoma by interference with apoptosis signaling pathways. Biochem Biophys Cancer Res. 66 (15) (2006) 7589–7597.
Res Commun 396 (2) (2010) 335–341. [161] M. Adorno, M. Cordenonsi, M. Montagner, S. Dupont, C. Wong, B. Hann, et al., A
[132] A. Fomenkov, R. Zangen, Y.P. Huang, M. Osada, Z. Guo, T. Fomenkov, et al., RACK1 mutant-p53/Smad complex opposes p63 to empower TGFbeta-induced metas-
and stratifin target DeltaNp63alpha for a proteasome degradation in head and tasis, Cell 137 (1) (2009) 87–98.
neck squamous cell carcinoma cells upon DNA damage, Cell Cycle 3 (10) (2004) [162] D.K. Carroll, J.S. Brugge, L.D. Attardi, p63, cell adhesion and survival, Cell Cycle 6
1285–1295. (3) (2007) 255–261.
66 V. Graziano, V. De Laurenzi / Biochimica et Biophysica Acta 1816 (2011) 57–66

[163] K. Higashikawa, S. Yoneda, K. Tobiume, M. Taki, H. Shigeishi, N. Kamata, Snail- [166] J.L. Hornick, G.Y. Lauwers, R.D. Odze, Immunohistochemistry can help
induced down-regulation of DeltaNp63alpha acquires invasive phenotype of distinguish metastatic pancreatic adenocarcinomas from bile duct ade-
human squamous cell carcinoma, Cancer Res. 67 (19) (2007) 9207–9213. nomas and hamartomas of the liver, Am. J. Surg. Pathol. 29 (3) (2005)
[164] J.G. Clohessy, P.P. Pandolfi, Beta-tting on p63 as a metastatic suppressor, Cell 137 381–389.
(1) (2009) 28–30. [167] O. Kaufmann, E. Fietze, J. Mengs, M. Dietel, Value of p63 and cytokeratin 5/6 as
[165] H. Yanai, N. Takahashi, M. Omori, W. Oda, I. Yamadori, S. Takada, et al., immunohistochemical markers for the differential diagnosis of poorly differen-
Immunohistochemistry of p63 in primary and secondary vulvar Paget's disease, tiated and undifferentiated carcinomas, Am. J. Clin. Pathol. 116 (6) (2001)
Pathol. Int. 58 (10) (2008) 648–651. 823–830.

You might also like