35
REVIEW ARTICLE Sp1 and the ‘hallmarks of cancer’ Kate Beishline and Jane Azizkhan-Clifford Department of Biochemistry and Molecular Biology, Drexel University College of Medicine, Philadelphia, PA, USA Keywords cancer; oncogene; Sp1; transcription; tumor suppressor Correspondence J. Azizkhan-Clifford, Department Biochemistry and Molecular Biology, Drexel University College of Medicine, 245 North 15th Street, MS 497, Philadelphia, PA 19102, USA Fax: 214 762 4452 Tel: 215 762 4446 Email: [email protected] (Received 23 April 2014, revised 26 September 2014, accepted 10 November 2014) doi:10.1111/febs.13148 For many years, transcription factor Sp1 was viewed as a basal transcrip- tion factor and relegated to a role in the regulation of so-called housekeep- ing genes. Identification of Sp1’s role in recruiting the general transcription machinery in the absence of a TATA box increased its importance in gene regulation, particularly in light of recent estimates that the majority of mammalian genes lack a TATA box. In this review, we briefly consider the history of Sp1, the founding member of the Sp family of transcription fac- tors. We review the evidence suggesting that Sp1 is highly regulated by post-translational modifications that positively and negatively affect the activity of Sp1 on a wide array of genes. Sp1 is over-expressed in many cancers and is associated with poor prognosis. Targeting Sp1 in cancer treatment has been suggested; however, our review of the literature on the role of Sp1 in the regulation of genes that contribute to the ‘hallmarks of cancer’ illustrates the extreme complexity of Sp1 functions. Sp1 both acti- vates and suppresses the expression of a number of essential oncogenes and tumor suppressors, as well as genes involved in essential cellular func- tions, including proliferation, differentiation, the DNA damage response, apoptosis, senescence and angiogenesis. Sp1 is also implicated in inflamma- tion and genomic instability, as well as epigenetic silencing. Given the apparently opposing effects of Sp1, a more complete understanding of the function of Sp1 in cancer is required to validate its potential as a therapeu- tic target. Introduction The transcription factor Specificity Protein 1 (Sp1) was first identified as a promoter-specific binding factor that is essential for transcription of the SV40 major Immediate Early (IE) gene [1,2]. Initially, Sp1 was considered a general transcription factor that is required for transcription of a large number of ‘house- keeping genes’, so-called because of their involvement in metabolism, cell proliferation/growth, and cell death [3]. It has become increasingly clear that many of the housekeeping genes play critical roles in cancer initia- tion and progression. Estimates suggest that there are at least 12 000 Sp binding sites within the human gen- ome, and most studies support the idea that Sp1 not only maintains basal transcription, but also contributes to the regulation, i.e. induction and inhibition, of tran- scription of a large number of cellular genes [46]. The Abbreviations Akt, protien kinase B; ATM, ataxia telangiectasia mutated; ATR, AT and Rad3 related; Brca1, breast cancer associated factor 1; b-TCRP, beta-transducin containing protein; EGFR, epidermal growth factor receptor; ER, estrogen receptor; ERK, extracellular signal-regulated kinase; GSK3b, glycogen synthase kinase-3 beta; IGF, insulin-like growth factor; IGF1R, insulin-like growth factor receptor 1; JNK1, Jun N- terminal kinase 1; MMPs, matrix metallproteinases; NSAIDS, non-steroidal anti-inflammatory drugs; PI3K, phosphoinositide 3-kinase; PKC, protein kinase C; PML, promyelocytic leukemia protein; ROS, reactive oxygen species; SCF, SkpCullinF box; VEGF, vascular endothelial growth factor; VHL, Von Hippel-Lindau Protein. 224 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the 'hallmarks of cancer

Embed Size (px)

Citation preview

REVIEW ARTICLE

Sp1 and the ‘hallmarks of cancer’Kate Beishline and Jane Azizkhan-Clifford

Department of Biochemistry and Molecular Biology, Drexel University College of Medicine, Philadelphia, PA, USA

Keywords

cancer; oncogene; Sp1; transcription; tumor

suppressor

Correspondence

J. Azizkhan-Clifford, Department

Biochemistry and Molecular Biology, Drexel

University College of Medicine, 245 North

15th Street, MS 497, Philadelphia, PA

19102, USA

Fax: 214 762 4452

Tel: 215 762 4446

E–mail: [email protected]

(Received 23 April 2014, revised 26

September 2014, accepted 10 November

2014)

doi:10.1111/febs.13148

For many years, transcription factor Sp1 was viewed as a basal transcrip-

tion factor and relegated to a role in the regulation of so-called housekeep-

ing genes. Identification of Sp1’s role in recruiting the general transcription

machinery in the absence of a TATA box increased its importance in gene

regulation, particularly in light of recent estimates that the majority of

mammalian genes lack a TATA box. In this review, we briefly consider the

history of Sp1, the founding member of the Sp family of transcription fac-

tors. We review the evidence suggesting that Sp1 is highly regulated by

post-translational modifications that positively and negatively affect the

activity of Sp1 on a wide array of genes. Sp1 is over-expressed in many

cancers and is associated with poor prognosis. Targeting Sp1 in cancer

treatment has been suggested; however, our review of the literature on the

role of Sp1 in the regulation of genes that contribute to the ‘hallmarks of

cancer’ illustrates the extreme complexity of Sp1 functions. Sp1 both acti-

vates and suppresses the expression of a number of essential oncogenes

and tumor suppressors, as well as genes involved in essential cellular func-

tions, including proliferation, differentiation, the DNA damage response,

apoptosis, senescence and angiogenesis. Sp1 is also implicated in inflamma-

tion and genomic instability, as well as epigenetic silencing. Given the

apparently opposing effects of Sp1, a more complete understanding of the

function of Sp1 in cancer is required to validate its potential as a therapeu-

tic target.

Introduction

The transcription factor Specificity Protein 1 (Sp1) was

first identified as a promoter-specific binding factor

that is essential for transcription of the SV40 major

Immediate Early (IE) gene [1,2]. Initially, Sp1 was

considered a general transcription factor that is

required for transcription of a large number of ‘house-

keeping genes’, so-called because of their involvement

in metabolism, cell proliferation/growth, and cell death

[3]. It has become increasingly clear that many of the

housekeeping genes play critical roles in cancer initia-

tion and progression. Estimates suggest that there are

at least 12 000 Sp binding sites within the human gen-

ome, and most studies support the idea that Sp1 not

only maintains basal transcription, but also contributes

to the regulation, i.e. induction and inhibition, of tran-

scription of a large number of cellular genes [4–6]. The

Abbreviations

Akt, protien kinase B; ATM, ataxia telangiectasia mutated; ATR, AT and Rad3 related; Brca1, breast cancer associated factor 1; b-TCRP,

beta-transducin containing protein; EGFR, epidermal growth factor receptor; ER, estrogen receptor; ERK, extracellular signal-regulated

kinase; GSK3b, glycogen synthase kinase-3 beta; IGF, insulin-like growth factor; IGF1R, insulin-like growth factor receptor 1; JNK1, Jun N-

terminal kinase 1; MMPs, matrix metallproteinases; NSAIDS, non-steroidal anti-inflammatory drugs; PI3K, phosphoinositide 3-kinase; PKC,

protein kinase C; PML, promyelocytic leukemia protein; ROS, reactive oxygen species; SCF, Skp–Cullin–F box; VEGF, vascular endothelial

growth factor; VHL, Von Hippel-Lindau Protein.

224 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

activities of Sp1 are regulated throughout the cell

cycle, and are modulated by post-translational modifi-

cation(s) in response to a large array of signals as dis-

cussed here. We also discuss the regulatory functions

of Sp1, as well as key protein–protein interactions,

with a focus on its role in cancer progression.

Sp1 was the founding member of the Sp transcrip-

tion factor family, the members of which contain

C2H2-type zinc fingers and resemble the larger family

of ‘Kr€uppel-like factors’ [3,7,8]. The Sp factors prefer-

entially bind GC boxes, while Kr€uppel-like factors

bind CACCC boxes [9–14]. Sp factors may be sepa-

rated into two groups: Sp1–4 and Sp5–9. The domain

organization of Sp1–4 is very similar, while Sp5–9 are

significantly different from Sp1–4, and are more simi-

lar to the other Kr€uppel-like factors. Sp1–4 have simi-

lar N–terminal transactivation domains (Fig. 1)

characterized by glutamine-rich regions, which, in

most cases, have adjacent serine/threonine-rich regions.

There is also a highly charged region referred to as the

C domain. Although the transactivation domains are

similar in overall amino acid content, they are not sig-

nificantly homologous to one another, which accounts,

at least in part, for differences in their functions. The

DNA-binding domains of Sp1–4 include three highly

homologous C2H2-type zinc fingers, which preferen-

tially bind to the same GC consensus site [50-(G/T)

GGGCGG(G/A)(G/A)(G/T)-30] [15,16]. Each of three

zinc fingers in Sp1 has its own specific sequence

preference, and all three are required for high-affinity

binding, with the first zinc finger having the highest

sequence specificity [17,18]. Interestingly, loss of all

three Sp1 zinc fingers abolishes not only DNA binding

but also nuclear localization. Nuclear localization

requires coordinated binding of zinc [19,20].

In contrast to other Sp family members, Sp1 con-

tains a C–terminal multimerization domain that medi-

ates super-activation of promoters containing multiple

adjacent Sp sites. Here, DNA-bound and unbound

Sp1 molecules interact, forming tetramers that develop

into larger multimers that promote association of

proximal promoter regions with distal enhancers by

forming looped regions of DNA [21,22]. This super-

activation mechanism is discussed further below. Mul-

timerization also exposes various regions of the Sp1

molecules for protein–protein interactions and post-

translational modifications.

Sp3 is the family member that is most similar to

Sp1. Unlike Sp1, Sp3 does not contain a multimeriza-

tion domain. Instead, it contains an inhibitory domain

located just N–terminal to its DNA-binding domain,

that mediates transcriptional repression in some

instances [23–25]. The D domain of Sp3, which bears

little similarity to the D domain of Sp1, may some-

times act as a transcriptional inhibitor [26]. Sp1 and

Sp3 have equal affinity for the GC box binding motifs,

and, as such, their relative abundance and binding

ratios influence gene regulation [27]. All these data are

consistent with the notion that the presence of a single

Sp site allows binding of either Sp1 or Sp3 to promote

Fig. 1. Transcription factor Sp1 and its closest related family members. Sp1 and its four closest family members are grouped together

based on their Cys2His2 zinc finger DNA binding domains (purple) and transactivation domains (light and dark blue), which are located N–

terminal to the DNA binding domain. The transactivation domains are composed of serine/threonine-rich regions, flanked by regions rich in

glutamine. The charged domain (green), although not essential for DNA binding, promotes binding of the zinc fingers to their DNA target.

Only Sp1 contains a multimerization domain (beige) which facilitates interaction between multiple Sp1 molecules and promotes super-

activation of genes. The position of the Buttonhead domain (Btd), which is conserved in Drosophila Sp1 and the similar Drosophila protein

Buttonhead, is indicated. Grey regions lack similarity between proteins.

225FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

basal transcription. Unlike Sp1, Sp3 does not have the

ability to super-activate promoters by bridging multi-

ple Sp binding sites. This finding supports the possibil-

ity of a potential inhibitory function for Sp3 in

promoters containing multiple Sp sites; in regulatory

regions with multiple sites, the presence of Sp3 at one

or more of these sites may interfere with Sp1 binding

and suppress the gene activation that requires multi-

merization between multiple Sp1 molecules on adjacent

or distal binding sites [23,24]. An example of this phe-

nomenon is regulation of the human topoisomer-

ase IIa promoter, for which activation occurs by

bridging between Sp1 bound at sites in the distal

enhancer and the proximal promoter. Here, proximal

and distal binding of Sp1 activates transcription, while

competition between Sp1 and Sp3 for binding at either

the distal enhancer or both binding regions results in

Sp3-dependent repression [28]. Displacement of a sin-

gle Sp1 molecule by Sp3 leads to a decrease in Sp1-

dependent transactivation [24,29]. The interplay

between Sp1 and Sp3 is discussed further below.

Both Sp1 and Sp3 are essential for embryonic devel-

opment, and knockout of either factor is embryo-lethal

in mice [30–32]. It has also been shown that Sp1 and

Sp3 are differentially regulated during development

[33], implying a role for these factors in cellular differ-

entiation and development.

Sp1 is a eukaryotic specific factor, and the full-

length Sp1 protein is highly conserved among mamma-

lian species. The family is defined by the DNA-binding

domain, which is also conserved in fish, reptiles and

birds. There are multiple Sp/Kr€uppel-like factor family

members that regulate various stages of development;

there is a high degree of amino acid conservation, par-

ticularly in the zinc-finger region, between Drosophila

Sp1 and those of higher species [34]. As indicated in

Fig. 1, Sp1 also contains a small region of homology

with the Drosophila protein buttonhead [35]. A yeast

homolog of any of the Sp1 factors has yet to be identi-

fied based on sequence homology or conservation of

its DNA binding ability. Moreover, no functional

homolog has been found either.

Sp1 contains two main transactivation domains, A

and B, which, in cooperation with the DNA binding

domain, support the majority of Sp1’s transcriptional

activity (Fig. 1). Domains A and B are highly gluta-

mine- and serine/threonine-rich, with domain B being

over twice as lard as domain A [36]. These transactiva-

tion domains have no sequence-specific transcriptional

activity alone, and require binding to promoters

through the DNA-binding domain to promote tran-

scription [37]. The C domain is highly charged, and

supports both DNA binding and transactivation,

although it is not essential for either process [37]. The

D domain, as discussed above, is important for multi-

merization, but other portions of the protein probably

also participate in this process.

Domains A, B, C and D are largely unstructured,

containing no known structural elements. In fact, the

only structured region of Sp1 is the zinc-finger domain,

for which an NMR structure has been elucidated

[38,39]. The highly disordered structure of the protein

suggests that Sp1 is probably capable of interacting with

a large number of protein partners, as discussed below.

Sp1 has three isoforms. Isoform A is the primary

form of the protein, containing 785 amino acids. Iso-

form B is a translational variant, in which an alterna-

tive start codon is used, resulting in a protein lacking

the first seven amino acids. Some studies, discussed

below, have suggested that these amino acids affect

the stability of the protein. The final variant, isoform

C, results from use of an alternative splice acceptor

site in the 3rd exon of the Sp1 pre-mRNA, which pro-

duces a protein that lacks amino acids 55–102 [40].

Few papers have discussed the functional significance

of the various isoforms, and more work is required to

identify the potentially different expression and func-

tions of each isoform.

Regulation of Sp1

As suggested by early publications [41,42], Sp1 is

highly modified by almost all the common forms of

post-translational modification, including phosphory-

lation, O-linked glycosylation, acetylation, SUMOyla-

tion, and ubiquitylation, and is targeted to

proteasome-mediated degradation pathways. Table 1

lists the post-translational modifications reported in

the literature, and their effect on Sp1 activity (when

documented).

Post-translation modifications

The 785 amino acid Sp1 protein contains 164 serine

and threonine residues, suggesting that it may be

highly phosphorylated and O-glycosylated. These

modifications affect not only Sp1 activity, but also

its stability, through modulation of proteolytic cleav-

age and proteasomal degradation. Several of the

modifications that have been identified in vitro or

through screens have yet to be associated with spe-

cific functions in cells. Signaling through specific kin-

ases probably has effects that have yet to be

discovered. Several of these modifications are dis-

cussed below within the context of modulating Sp1

stability and transcriptional activity.

226 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

A number of key signaling kinases have been shown

in vitro and in vivo to both directly and indirectly mod-

ify Sp1. These include extracellular signal-regulated

protein kinases (ERK) 1 and 2, c–Jun N–terminal

kinase 1 (JNK1), cyclin-dependent kinase 2 (CDK2)

and the phosphatidylinositol 3–kinase (PI3K)-like

kinases, ataxia telangiectasia mutated (ATM), ataxia tel-

angiectasia and Rad3-related kinase (ATR) and DNA-

dependent Protein Kinase (DNA-PK) (see Table 1 for

extensive list with references). These phosphorylation

events contribute to further post-translational modifi-

cation of Sp1, modulation of Sp1 stability, Sp1 DNA

binding affinity, and Sp1’s ability to transactivate tran-

scription. A number of serine/threonine residues have

been shown to be both phosphorylated and glycosylated,

and the dynamic between these two modifications is

probably important for regulation of function.

A single Sp1 molecule has the potential to be

extensively modified. The number of possible combina-

tions of modifications suggests that Sp1 may be

differentially modified even within the same cell, with

various levels of stability and differential functions

depending on the combination of modifications. As

mentioned above, the majority of the Sp1 protein is

unstructured, based on primary sequence information

[74], consistent with the idea that post-translational

modifications may modulate its structure. These modi-

fications and associated structural alterations support

the ability of Sp1 to interact with a large number of

binding partners, and to differentially modulate tran-

scription of a large number of genes.

Regulation of Sp1 stability

The levels of Sp1 protein are tightly regulated to main-

tain cellular homeostasis. Sp1 levels are modulated

throughout the cell cycle, and transcriptional regulation

is in part achieved by variable protein level. An early

study implicated reduced Sp1 O–glycosylation, associ-ated with glucose starvation, with its degradation [75].

Table 1. Sp1 post-translation modifications.

Residue Modification Enzyme Function References

Ser2 Phosphorylation; N–acetylserine Unknown Unknown [43,44]

Ser7 Phosphorylation Unknown Stability [43–45]

Lys16 SUMOylation; ubiquitinylation RNF4 Stability [45–48]

Ser56 Phosphorylation ATM/ATR Unknown [49]

Ser59 Phosphorylation CDK2; ERK1/2; PP2A Stability, DNA binding [45,50–53]

Ser101 Phosphorylation ATM/ATR – [49,54,55]

Ser111-114 Glycosylation OGT – Unpublishedb

Asp183 Cleavage Casp3 Unknown Unpublishedc

Ser220a Phosphorylation DNA-PK Transcription [56]

Thr278 Phosphorylation JNK1; ERK1/2 Stability [47]

Ser301 Glycosylation OGT – Unpublishedb

Thr355a Phosphorylation ERK1/2; JNK1 Transcription [57]

Thr453 Phosphorylation ERK1/2 Transcription [58–60]

Ser491 Glycosylation OGT Transcription [61,62]

Ser535/540 Glycosylation OGT – Unpublishedb

Asp584 Cleavage Casp3 – [63]

Ser612 Phosphorylation; glycosylation OGT Localization [64]

Thr640 Phosphorylation; glycosylation OGT Localization [64]

Ser641 Phosphorylation; glycosylation PKCf; OGT Transcription [64,65]

Thr651 Phosphorylation PKCf [66]

Thr668a Phosphorylation CKII; PPI; PKCf DNA binding [64,67,68]

Ser670 Phosphorylation PKCf – [68]

Thr681 Phosphorylation PP2A; PKCf – [50,68]

Ser698 Glycosylation Localization [64]

Ser702 Glycosylation OGT – [64]

Lys703 Acetylation P300; HDAC1 – [69–72]

Ser728 Phosphorylation GSK3b Degradation [73]

Ser732 Phosphorylation GSK3b Degradation [73]

Thr739 Phosphorylation Erk1/2, JNK1 Transcription, stability [47,48,58,73]

a Amino acid residue number is based on the full length 785 aa Sp1 protein. b Unpublished data from in vitro glycosylation assays Beishline

K, Azizkhan-Clifford J. c Unpulished data Torabi B, Azizkhan-Clifford J. Paranthetic Numbers indicate amino acid number in original Sp1

sequence which lacks amino acids 1–89.

227FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

The degradation occurred as a result of a cleavage event

in the N–terminus of the protein, which was dependent

on the presence of the first seven amino acids [75,76],

and was suggested to occur via interaction with the pro-

teasome [77]. Sp1 was also shown to interact directly

with b–transducin repeat-containing protein (b–TCRP),

a component of the Skp–Cullin–F box (SCF) ubiquitin

ligase complex [73], further supporting an interaction

with the proteasome. b–TCRP probably interacts with

Sp1 through a DSG (Asp-Ser-Gly) destruction box (b–TCRP binding motif) within the C–terminus of Sp1

[DSGAGS(727–732)], mediating proteasomal degrada-

tion of the protein. Evidence suggests that ubiquitin-

mediated degradation may be modulated by phosphor-

ylation of Sp1 by both ERK (on threonine 739) and

Glycogen synthase kinase-3 beta (GSK3b) (on threo-

nine residues 728 and 732) [73]. Despite strong evidence

for ubiquitin-mediated proteasomal degradation of

Sp1, residue(s) on Sp1 that are directly ubiquitylated

have yet to be identified.

The stability of Sp1 is also affected by SUMOyla-

tion, which occurs on lysine 16. This modification is

modulated by surrounding residues, and requires the

first seven amino acids of Sp1, which were previously

found to be associated with protease-mediated cleav-

age [45]. SUMOylation was shown to be associated

with destabilization of Sp1 [45,46,78], and later studies

showed that destabilization occurs through Ring finger

protein 4 (RNF4), a SUMO-dependent E3 ubiquitin

ligase. RNF4 interacts with both the SUMO-modified

region of Sp1 and its DNA-binding domain. This

interaction is associated with RNF4-dependent ubiqui-

tylation of Sp1, which is dependent on Sp1 SUMOyla-

tion [48]. The association of RNF4 with the DNA-

binding domain of Sp1 also suggested that RNF4 may

potentially affect Sp1 function independently of Sp1

ubiquitylation, and that Sp1 may modulate interaction

between RNF4 and other targets.

Phosphorylation of Sp1 modulates proteasome-

mediated degradation during mitosis. JNK1 phos-

phorylates Sp1 on threonine residues 278 and 739,

and these modifications are associated with stabiliza-

tion of Sp1 during mitosis [47]. Furthermore, JNK1-

dependent phosphorylation of threonine 739 inhibits

the interaction between Sp1 and RNF4, supporting

a role for JNK1 in promoting Sp1 stability [48]. The

effects of phosphorylation of threonine 278 on the

interactions between Sp1 and RNF4 or Sp1 and b–TCRP are unclear. This suggests the need for further

studies to address how Sp1 phosphorylation affects

SUMOylation and ubiquitylation, and the possibility

that Sp1 may be targeted by multiple E3 ubiquitin

ligase complexes.

Protein–protein interactions involving Sp1

Sp1 has been shown to interact with a number of cel-

lular factors to modulate expression of specific genes,

facilitate post-translational modification of Sp1, and

regulate Sp1 stability. Table 2 provides a comprehen-

sive list of factors that have been shown to interact

with Sp1. These include members of the basal

transcriptional machinery, other transcription factors,

cell-cycle regulators, chromatin modifiers and ATP-

dependent remodelers, as well as factors participating

in cellular processes such as DNA repair. A number of

these interactions are associated with specific cellular

functions, which are discussed in more detail through-

out later sections. A number of factors listed in table 2

have only been shown to interact in vitro [117] and

many have not been associated with specific cellular

functions. Sp1 may interact with multiple factors

simultaneously, particularly when present as a multi-

mer, resulting in a large number of complexes. This

makes it difficult to distinguish specific Sp1 functions

within a single complex, and also makes identifying

direct Sp1 interactions in vivo experimentally difficult.

Sp1 in cancer

While often described as a general transcription factor,

Sp1-dependent transcription is highly regulated

throughout development, cellular differentiation and

tumorigenesis. Sp1 differentially regulates a large num-

ber of key factors that are important in a number of

disease states, including cancer, which is the focus of

the remainder of this review.

Sp1 is over-expressed in a number of cancers,

including breast, gastric, pancreatic, lung, brain (gli-

oma) and thyroid cancers (Table 3) [156–160]. In

patient samples and cancer models, Sp1 levels correlate

with stage, invasive potential and metastasis. Sp1 levels

correlate with the survival of patients in almost all

cancers, with high levels of Sp1 being associated with

poor prognoses. On this basis, a number of studies

have addressed the possibility of targeting Sp1 to treat

cancer cells or sensitize tumors to other treatments

[161–166]. However, specific inhibitors that target Sp1

have not been developed, and the feasibility of target-

ing Sp1 specifically in cancer treatment has not been

demonstrated. This is at least in part due to the myr-

iad activities of Sp1 and incomplete mechanistic under-

standing of Sp1 function.

Hanahan and Weinberg have thoroughly reviewed

the cellular pathways essential for tumor formation

and cancer progression [170,171]. These include the

eight major hallmarks of cancer: sustained proliferative

228 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

Table 2. Sp1 interactions.

Protein Region References

General transcription factors

TBP A, B [79,80]

TAF4 A, B [61,83–89]

TAF7 Unknown [95]

TFIIAa Unknown [96]

dTAFII110 Unknown [61,62]

TAF–Ia/b Unknown [99]

Transcription factors

p53 Unknown [106,107]

E2F1 Amino acids 699–785 [102,110]

c–MYC Unknown [111]

c–Jun Amino acids 424–542 [113]

TReP–132 Unknown [114]

C/EBPb Unknown [116]

PML Unknown [118]

ELF1 Unknown [120]

Oct1 Unknown [124]

NF–YA Unknown [125]

Che–1 Unknown [126]

EGR1 Unknown [127]

DLX4 Unknown [128]

HIF1a Unknown [129]

YYI Unknown [131]

Fli–1 Unknown [132]

SREBP2 Unknown [133,134]

SMAD2 Unknown [135,136]

STAT6 Unknown [138]

Other factors

TG2 Unknown [140,141]

Huntingtin Unknown [142]

HIV–1 Tat Unknown [56,144]

RNF4 Amino acids 619–785 [48]

PKCf Unknown [121]

Chromatin remodeling factors

p300 Amino acids 138–232

Amino acids 262–487

[69,71,81,82]

SWI/SNF complex, BRG1, BAF155, BAF170 Unknown [90–94]

HMGA1 Unknown [97]

HMGA2 Unknown [98]

DNMT1 Unknown [100,101]

HDAC1 Amino acids 622–788 [69,102–105]

HDAC2 Unknown [103,108,109]

RbAp48 Unknown [103]

MCAF1/2 Unknown [112]

Tumor suppressors

Brca1 Unknown [115]

SKP2 Unknown [117]

MDM2 Unknown [119]

VHL Unknown [121–123]

b–TCRP Unknown [73]

DNA repair factors

Nbs1 Unknown [74]

Brca2 Unknown [117]

Rad51 Unknown [117]

ATM Unknown [130]

229FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

signaling, replicative immortality, resistance to cell

death and avoidance of immune destruction, induction

of angiogenesis, invasion and metastasis, and de-regu-

lation of cellular energetics [170,171]. Sp1 is critical for

the regulation of genes whose products are responsible

for each of these hallmarks. In addition, Sp1 may

influence two of the ‘enabling characteristics’ of can-

cer: inflammatory signaling, which alters the oxidative

state of the cell [170], and maintenance of genomic sta-

bility. Here we discuss key genes for which Sp1 has

been shown to play a regulatory role specifically in

tumors (Table 4).

Sp1 modulates cancer cell proliferative and

survival signals

Sustained proliferative signaling: IGF1R

The requirement for cancer cells to maintain their own

proliferative signals, independent of normal physiologi-

cal signaling, is well established. A large number of

essential signaling molecules in normal and cancer cells

are associated with Sp1-mediated transcriptional regu-

lation, including epidermal growth factor and its

receptor, fibroblast growth factor, and insulin-like

growth factor and its receptor (Table 4). In addition,

growth signaling directly regulates Sp1 and additional

Sp1-dependent proliferative signals.

One of the best examples of an Sp1-regulated

growth signaling molecule that is disrupted in cancer is

insulin-like growth factor 1 receptor (IGF1R). It is

well established that insulin-like growth factor (IGF)

signaling is utilized by cancer cells to maintain the pro-

liferative program, and, for this reason, IGF1R has

been proposed as a potential target in tumor treatment

[300,301]. When the IGF1R promoter was first charac-

terized, it was shown to contain as many as eight Sp

target sequences [195], and subsequent work showed

that a specific Sp1 site is essential for IGF1R transcrip-

tional regulation (Fig. 2) [213]. It is now understood

that growth signals directed through insulin-responsive

binding protein and Sp1 coordinate binding to GC-

rich and AT-rich regions of the IGF1R promoter to

increase transcription [302]. Further evidence suggests

that additional factors, including breast cancer-associ-

ated factor 1 (BRCA1), high-mobility group A1, estro-

gen receptor a (ERa) and ATM kinase, associate with

Sp1 (see Table 2) and modulate IGF1R transcription

[97,198,201].

BRCA1 has been shown to play a role in suppress-

ing Sp1-mediated regulation of IGF1R, which has sig-

nificant implications for the development and

Table 2. (Continued).

Protein Region References

Cell cycle

p21 Unknown [117]

CDK4 Unknown [117]

Cyclin A Unknown [137]

Nuclear receptors

SMRT N–terminus [139]

NCoR/BCoR N–terminus [139]

PPARc Unknown [143]

AR Unknown [145]

ERa Unknown [116,146–155]

Table 3. Sp1 expression in cancer.

Cancer

type References

Sp1

expression Survival Invasiveness, proliferation

Gastric [167] High Decrease –

Diffuse-type gastric [168] High Decrease Positive

Intestinal-type gastric [168] Low Decrease Positive/negative

Pancreatic [158] High Decrease Positive

Lung, early [169] High – –

Lung, stage IV [169] Low – Negative

Breast [160] High Decrease Positive

Glioma [159] High Decrease –

Thyroid [156] High – –

230 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

progression of breast cancer in carriers of the familial

BRCA mutation(s). Over several years, the Werner

group have established that BRCA1 negatively

regulates IGF1R transcription through direct interac-

tion with Sp1 that inhibits Sp1 binding at the IGF1R

promoter [199,200,262]. The interaction between Brca1

and Sp1 is stimulated by IGF1 [199,262]. BRCA1

directly interacts with Sp1 through amino acids 260–802, but the region of Sp1 that binds Brca1 has yet to

be identified [115]. It was later established that individ-

uals with the BRCA1 mutation have increased levels

of IGF1R compared to patients lacking the BRCA1

mutation [199]. This suggests that individuals with the

BRCA1 mutation are probably able to maintain

increased IGF1R transcription in the presence or

absence of IGF1 stimulation, due to a lack of Sp1-

dependent transcriptional inhibition.

EGFR

The epidermal growth factor receptor (EGFR) and its

family of tyrosine kinase receptors are associated with

sustained proliferation, as well as promoting invasion

and metastasis in cancer cells. Mutations in EGFR

have been identified in a number of cancers, and, in

addition, the levels of receptor are modulated to pro-

mote increased signaling in tumor tissues. For these

reasons, EGFR and its family of receptors have

Table 4. Cancer-related genes regulated by Sp1.

Gene References

Sustained proliferation/immortality

hTERT/hTERC [98,105,108,172–181]

p53/MDM2 [182–187]

p16 [188–190]

p21 [191–194]

IGF1R [97,106,115,195–201]

EGFR [202–207]

EGF [208]

FGF [209–212]

IGF [213–218]

Apoptosis

Survivin [162,164,172,173,219–224]

Trail–R2 [163,225,226]

Bcl–2 [70,227]

TRAIL [228]

MCL–1 [221,229]

XIAP [230]

Bak [70,168]

FasL [219,231–233]

Angiogenesis

VEGF [59,121,166,167,234–252]

TSP–1 [253]

PDGF [58,123,254–257]

uPA [258–260]

DNA damage/stress response

Brca1 [261,262]

ATM [130,263,264]

MDC1 [265]

Cdc25B [266,267]

RECQ4 [268]

PARP [269,270]

XRCC1 [271,272]

BLM [273]

CHEK2 [274,275]

XPC [276]

XPB [277,278]

XPD [277]

DDB1/2 [279]

DNA–PK/Ku70/80 [280]

XRCC5 [281]

ERCC6 (CSB) [282]

WRN [283]

Invasion and metastasis

MMP9 [60,284]

MT1–MMP [285,286]

RECK [60,287–289]

E–cadherin [169,290–294]

Integrin a5 [153,295,296]

MMP2 [90,137,159,284,297–299]

Fig. 2. Organization of GC boxes in Sp1-regulated promoters. Each

Sp1-regulated gene has a unique organization of transcription factor

binding sites. Multiple sites have been identified as being key for

Sp1-dependent increases or decreases in transcriptional activity of

the genes. It should be noted that the indicated enhancing and

inhibitory effects may be due to both binding or lack of binding at

these sites, depending on the gene context. Sp1 sites may be

found both upstream and downstream of transcription start sites,

and may coordinate with nearby binding sites for other factors. In

addition, other Sp factors, such as Sp3, may bind the same sites

and have different regulatory effects.

231FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

become targets for chemotherapeutic intervention in

recent years [303].

Sp1 was identified as a transcriptional regulator of

EGFR in the late 1980s, and Sp1 probably participates

in regulation of EGFR in cancer [202,204,205]. Upon

further assessment, a number of factors have been

shown to either inhibit or activate Sp1-dependent tran-

scription of the EGFR gene. Pro-myelocytic leukemia

protein (PML) has been shown to associate with Sp1,

and this interaction inhibits transcriptional activation

of the EGFR promoter. The interaction occurs through

the dimerization domain of PML and the DNA-bind-

ing domain of Sp1, thereby inhibiting Sp1’s ability to

bind to its target sequences within the EGFR promoter

and probably other promoter regions [118]. More

recent studies have shown that over-expression of

PML enhances localization of Sp1 to PML nuclear

bodies, and increases Sp1 degradation [304]. SUMOy-

lation of Sp1 was found to promote its localization to

PML nuclear bodies and subsequent Sp1 degradation.

Degradation of SUMOylated Sp1 has been associated

with RNF4, and is inhibited by phosphorylation

events during mitosis, when it has been suggested that

Sp1 is generally evicted from chromatin [305]. It may

be interesting to determine whether there is overlap

between the PML-mediated pathway and other aspects

of SUMO-mediated regulation of Sp1 throughout the

cell cycle [45,46,78].

Other factors that regulate EGFR in an Sp1-depen-

dent manner include the histone deacetylase HDAC1,

which has been shown to bind in a complex with

Kruppel-like factor 10 (KLF10/TIEG1) to Sp1 sites

within the promoter and to inhibit histone acetylation

to suppress transcription [306]. In addition, the reti-

noic acid receptor has been shown to bind Sp1 sites

within the EGFR promoter and inhibit Sp1 binding to

its target sequences and suppress transactivation [307].

This is in contrast to estrogen receptors a and b, bothof which have been shown to transactivate the EGFR

promoter through interaction with Sp1 rather than

through direct binding [206,308].

Enabling replicative immortality

It is well established that normal cells have finite repli-

cative potential. Cells gradually approach their replica-

tive limit until they enter a senescent state. A cell that

is capable of overcoming these proliferative barriers

becomes immortalized (in culture), and may become

tumorigenic (in vivo). Cellular senescence requires a

number of factors, but primarily involves the CDK

inhibitor p16(Ink4a), transcription factor p53, and sta-

bilization of telomeres at the ends of chromosomes.

Sp1 is an essential regulator of the genes encoding

p16, p53 and several components of telomere mainte-

nance, including telomerase (Table 4), making it an

essential factor to evade entrance into senescence and

achieve replicative immortality.

p53

Like Sp1, p53 is a transcription factor that is impli-

cated in the regulation of a large number of genes.

There is a wealth of data available regarding p53’s role

in tumorgenesis, suggesting that the dynamics between

Sp1 and p53 are key to understanding many tumor-

suppressing pathways. A number of these pathways

are highlighted here, demonstrating strong cooperation

between Sp1 and p53 in the regulation of many tumor

suppressors and oncoproteins, including p21, IGF1R,

BCL2 binding component 3 (PUMA), DNA (cytosine-

5) methyltransferase 1 (DMNT1) and others

[106,107,309,310]. It is important to note that regula-

tion of transcription by p53 and Sp1 is dynamic and

variable among genes. In some instances, the proteins

work together to increase gene transcription [100,107],

while other studies show opposing effects on transcrip-

tion of certain factors [283,311]. In-depth studies are

required to fully appreciate the significance of Sp1/p53

gene regulation.

Independent of the transcriptional activities of these

two factors together, Sp1 is implicated in regulation of

MDM2 proto-oncogene, E3 ubiquitin protein ligase

(MDM2), the key regulator of p53 stability in cancer

cells. A number of small nucleotide polymorphisms in

the MDM2 promoter have been shown to be associ-

ated with increased expression of the protein due to

enhanced transactivation by Sp1 [182–187]. The conse-

quence of this enhanced MDM2 expression is down-

regulation of p53 and increased cancer susceptibility.

Although there is some disagreement regarding the sig-

nificance of these small nucleotide polymorphisms

[312], other factors such as EGFR contain small nucle-

otide polymorphisms that are associated with increased

Sp1-dependent transcription, which affects the cellular

signaling environment to favor tumorigenesis [313].

This suggests a mechanism whereby Sp1 directly mod-

ulates cancer signaling, through genetic mutations that

directly enhance the ability of Sp1 to induce tumor-

promoting transcriptional changes.

p16

As mentioned above, the CDK inhibitor p16 is a key

regulator of cellular senescence. In normal dividing

cells, p16 functions to inhibit cyclin CDK4/6 activation

232 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

of the retinoblastoma protein, a key tumor suppressor

protein that controls cellular proliferation in the G1

phase [314]. p16 has been found to be either mutated or

silenced in a large number of malignancies in multiple

tissues [315]. However, evidence suggests that p16 may

be over-expressed in some cancers [316]. How p16 func-

tions in oncogenic signaling is still unclear [316], despite

its certain role in tumor formation and progression.

Sp1 is a key regulator of p16 expression [188–190].The proximal promoter of p16, a region essential for

p16 expression, contains multiple Sp sites (Fig. 2), and

expression of Sp1 is important for maintaining pro-

moter activity [189]. The p16 proximal promoter must

also be modified by p300, a histone acetyl transferase,

to increase H4 acetylation levels and allow for tran-

scriptional activation. p300 interacts directly with the

transactivation domains of Sp1, and is recruited to the

p16 promoter [188]. Sp1 was shown to be important

for the up-regulation of p16 in normal human fibro-

blasts, which undergo cellular senescence, suggesting

that Sp1 may play a role in the induction of senes-

cence. Although these studies show that p16 is modu-

lated by changes in Sp1 protein levels, they do not

provide convincing evidence demonstrating that up-

regulation of p16 during senescence is Sp1-dependent

[190]. Sp1 levels are also reportedly decreased during

DNA damage-induced cellular senescence. Decreased

Sp1 levels in response to DNA-damaging agents were

dependent on ATM and ATR kinases [317]. Phosphor-

ylation of Sp1 by these kinases is discussed below;

however, phosphorylation of Sp1 by ATM and/or

ATR has not been shown to influence its transcrip-

tional activity or stability [317].

Telomerase

Telomeres are a major determinant of replicative

potential. They protect chromosome ends by prevent-

ing chromosomal fusions, and thereby maintain geno-

mic integrity. In normal cells, telomere length

decreases during DNA replication in a phenomenon

known as the ‘end replication problem’. Once telomere

loss reaches a critical level, a cell enters senescence. In

immortalized cells, telomere length is maintained by a

special reverse transcriptase called telomerase. Most

tumor cells show increased expression of telomerase,

which is suppressed in most differentiated tissues. Sp1

is a key transcriptional regulator of telomerase subun-

its, and, for this reason, is essential in the establish-

ment of replicative immortality in cancer cells.

Initial characterization of the hTERT gene, which

encodes the catalytic subunit of telomerase, revealed

five binding sites for Sp factors and a binding site for

the Myc transcription factor, as well as lack of a

canonical TATA box (Fig. 2) [174,179,180]. Investiga-

tions showed that the five Sp binding sites in the

hTERT promoter supported transcription at a level

> 90% that of the of the full promoter, and E–boxregulation through Myc proteins required the Sp sites

[174].

The gene encoding the RNA component of telomer-

ase (hTERC) also contains multiple Sp binding sites

(Fig. 2) [181,318]. Although this promoter contains a

TATA box, it was found to be non-functional. Sp1

participates in hTERC expression; however, a CCAAT

box that binds the factor Nuclear Transcription Factor

Y Alpha (NF-YA) is sufficient for basal transcrip-

tional activity [318]. Of the four Sp binding sites in the

hTERC promoter, two were found to positively regu-

late hTERC expression, while the other two were

found to repress hTERC expression [317]. The Sp site

with the strongest effect on regulation was the site

closest to the CCAAT box, and this site negatively

regulated promoter activity [318]. Sp1 and NF-YA

have been shown to interact, and this interaction is

reduced by O–GlcNAcylation of Sp1 [125,319].

Increased O–GlcNAcylation is associated with cancer,

suggesting a potential mechanism for Sp1-dependent

increases in hTERC expression due to increased glyco-

lytic flux in tumors.

Sp1 has been shown to be associated with other fac-

tors that positively or negatively regulate telomerase

genes. One study addressed the involvement of MBD1-

containing chromatin-associated factor 1 (MCAF1,

ATF7IP), which binds Sp1 and promotes hTERT and

hTERC expression. It was also shown that decreases

in either MCAF1 or Sp1 decreased binding of RNA

polymerase II and Excision repair cross-complementa-

tion group 3 (ERCC3) to the hTERT and hTERC pro-

moters [320]. Multiple regions of the C–terminus of

Sp1 were found to interact with multiple portions of

MCAF1. The authors of this study suggest that

decreases in either MCAF1 or Sp1 affect the relative

methylation of the various CpG-rich regions of the

hTERT and hTERC promoters, resulting in long-term

suppression of telomerase expression in normal tissue

[320].

Sp1 also regulates hTERT expression by modulating

histone modifications, including acetylation. Even dur-

ing hTERT repression, Sp1 and Sp3 reportedly remain

bound to the promoter, and recruit histone deacetylas-

es [108]. Specifically, the N–terminal portion of Sp1

binds HDAC2 [108], decreasing histone acetylation at

the hTERT promoter and repressing hTERT expres-

sion. Further, studies using HDAC inhibitors also

implicated HDACs in hTERT repression by Sp1 [105].

233FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

Sp1 probably recruits additional HDACs such as

HDAC1, together with other factors associated with

HDAC complexes.

In addition to acetylation, histone methylation is

important for Sp1-dependent hTERT expression,

although methylation functions upstream of Sp1 bind-

ing. Liu et al. identified SET and MYND domain-con-

taining protein 3 (ZMYND1) as a specific histone

methyltransferase that is important for maintaining

histone H3 lysine 4 trimethylation of the hTERT pro-

moter [175]. In the absence of ZYMND1 and histone

H3 lysine 4 trimethylation, binding of both Sp1 and c-

Myc is attenuated, and hTERT expression is signifi-

cantly decreased [175].

Upstream signaling factors may promote Sp1-depen-

dent modulation of hTERT expression. One study

addressed the effect of survivin, another Sp1-regulated

factor, on hTERT expression [173]. Survivin was ini-

tially characterized as a caspase inhibitor that sup-

pressed apoptosis. Additional functions for survivin in

cellular signaling, including mitosis, have been charac-

terized [321]. During mitosis, survivin has been shown

to signal through Aurora B kinase to promote direct

or indirect phosphorylation of Sp1 and Myc to

increase hTERT expression [173]. CDK1 modulates

this signaling, suggesting Sp1 may be an important tar-

get for cell-cycle regulation by CDKs. Further work is

required to understand why survivin is important for

regulating hTERT expression.

Finally, the tumor-associated viruses Kaposi sar-

coma-associated herpes virus and human papilloma

virus (HPV) have been shown to modulate hTERT

expression in an Sp1-dependent manner [177,322],

supporting a role for Sp1 in hTERT-mediated

immortalization/tumorigenesis. These viruses express

the viral proteins Kaposi sarcoma latency-associated

nuclear antigen and HPV E6 protein, respectively,

that either directly or indirectly modulate the hTERT

promoter activity through Sp1 sites [177,322]. This

supports a role for Sp1 in tumor formation in

latently infected cells in both pathologies associated

with HPV and Kaposi sarcoma.

The hTERT promoter provides an excellent example

of the relationship between Sp1 binding sites and

regions of G–quadruplex DNA. G–quadruplex DNA,

or G4 DNA, refers to DNA that has adopted a type

of non-B form DNA structure, involving unique stack-

ing of nucleic acids in G–rich sequences of DNA and

RNA [323]. These sequences are present throughout

the human genome, and are more concentrated in telo-

meric regions, fragile sites for DNA double- and sin-

gle-strand damage, and transcriptional regulatory

regions of genes. A distinct correlation between the

occurrence of G4 DNA and Sp1 binding has been

established [324,325]. Sequence analysis of the hTERT

promoter suggests that G–quadruplex structures may

form within the Sp1 binding motifs, and it was sug-

gested they may negatively affect Sp1 binding [178].

Interestingly, other in vitro studies and biophysical

data suggest that Sp1 binds G–quadruplex structures

that contain the minimal consensus motif (50-GGGCGG-30), as well as G–quadruplex structures

that do not contain this motif [326]. These unique

DNA structures have been shown to both activate and

inhibit transcription, depending on their location rela-

tive to the transcription start site [323]. In addition to

hTERT, these complexes have been identified in the

regulatory regions of a number of genes encoding key

oncogenic factors, underscoring the importance of

understanding how G4 DNA contributes to transcrip-

tional regulation, as well as to the maintenance of

genomic stability.

Evading growth suppression

In the absence of nutrient and growth factor stimula-

tion, most cell populations are growth-suppressed or

quiescent. In addition, cellular stresses and DNA dam-

age may suppress growth to prevent cells from divid-

ing in the presence of a stressful environment. In

cancer cells, resistance to these stress signals may be

generated, and pathways that normally suppress cell

growth may be re-wired. Sp1 and its family members

regulate a number of cellular stress response factors,

such as cell-cycle regulators and DNA damage

response proteins (Table 4). In many of cases, Sp1 acts

as a basal transcriptional regulator, and has not been

shown to transcriptionally regulate these factors in

response to any specific stimulus. However, there is

evidence that Sp1 may modulate the cellular response

to specific stresses, such as that caused by ionizing

radiation exposure.

p21

The cyclin-dependent kinase inhibitor p21 (Waf1/

Cip1) is essential for cell-cycle arrest in response to

cellular stress signals [327]. Based on this and addi-

tional findings, p21 has been implicated in cancer

formation and progression [328]. The regulation of

p21 by Sp1 was first characterized in leukemic cells

[192]. This study identified two main regions in the

p21 promoter that may be bound by Sp1: nucleo-

tides �129 to �99 and �86 to �57 relative to the

transcriptional start site (Fig. 2). These binding sites

are responsible for activation of p21 transcription in

234 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

response to okadaic acid and phorbol ester, both of

which induce differentiation in these cells [192].

These sites were more proximal to the promoter

than the originally identified p53 binding sites that

control activation of p21 in response to cell stress

and DNA damage [329].

Multiple signaling pathways are associated with

Sp1 regulation of the p21 promoter. Early studies

showed that signaling through the progesterone

receptor stimulated Sp1-dependent up-regulation of

the p21 promoter. This hormone signaling probably

occurs in a manner similar to androgen and estrogen

receptor signaling through Sp1. In this case, proges-

terone receptor signaling also required the acetyl

transferase p300, which is probably recruited in an

Sp1-dependent manner [330]. The Sp1-dependent reg-

ulation of p21 has also been associated with the

activities of signal transducer and activator of tran-

scription protein 3 and 6 (STAT3 and STAT6); the

direct interaction of Sp1 with STAT6 may account

for the stimulation of p21 expression in breast can-

cer cells [138,331].

Several years after their initial characterization, Pa-

gliuca et al. studied the regulation of a number of

CDK inhibitors, including p21, by Sp factors [332].

They showed that Sp1 was a strong activator of tran-

scription, while Sp3 appeared to only weakly activate

the minimal proximal promoters of p21 and other

family members. This suggested that there may be

some interplay between Sp1 and Sp3 in regulation of

the CDK inhibitors. A number of reports have dem-

onstrated clear roles for Sp1 and Sp3 in activation

and repression, respectively, of the p21 promoter. Ka-

vurma and Khachigian identified an interesting phe-

nomenon regarding the three most proximal (A–C)Sp binding sites [194]; they showed that, although

mutation of all three of these sites abolishes promoter

activity, these three sites are essential for repression,

presumably by Sp3. Interestingly, mutation of the C

and E sites resulted in super-activation of the pro-

moter through Sp1. These studies are in agreement

with earlier work by Koutsodontis et al., who found

that Sp1 and Sp3 bound the proximal Sp binding

sites (A–E), with the highest affinity being observed

for site C. They showed that activation through these

sites was severely repressed by expression of an Sp1

mutant lacking the D domain, which is required for

multimerization and super-activation [193]. A decrease

in the Sp3 level resulted in an Sp1-dependent increase

in expression of p21. Although the decrease in Sp3

protein levels probably has a number of indirect

effects, Sp3 may inhibit Sp1-dependent activation of

p21, and its loss may facilitate Sp1 binding.

Another potential mechanism for Sp3-dependent

suppression of the p21 promoter was recently

proposed. When the three most proximal Sp sites

are bound by Sp1, they support activation; however,

when these sites are bound by Sp3, there is evidence

of polymerase stalling and inhibition of p21 tran-

script elongation. Given additional evidence implicat-

ing the more distal Sp binding sites as repressor

elements [191], it is likely that Sp3 inhibits the p21

promoter by preventing Sp1-dependent super-activa-

tion of transcription, which may occur through loop-

ing, to allow interaction of Sp1 bound at the

proximal sites with that bound at the distal sites.

Binding of Sp3 to either of these regulatory regions

may inhibit looping and super-activation, as shown

at other promoters [28].

The p21 promoter is reportedly methylated on

CpG islands in cancer cells [333]. This methylation

has been shown to inhibit binding of Sp1 and Sp3,

providing a potential mechanism for suppression of

p21 transcription mediated by inhibition of Sp1

binding at the promoter in these cells. The relation-

ship between DNA methylation and Sp1 is discussed

further below. However, p21 provides an interesting

example of how epigenetic modifications silence

genes that are normally positively regulated by Sp1,

even within the context of Sp1 over-expression in

cancer cells.

ATM

Sp1 has been shown to be associated with transcrip-

tional regulation of ATM, a key regulator in the cellu-

lar response to genotoxic stress, specifically DNA

double-strand breaks [130]. Serum starvation resulted

in decreased Sp1 binding to the ATM promoter, prob-

ably as a result of decreased Sp1 protein levels. Treat-

ment of cells with ionizing radiation, which activates

ATM activity, stimulated the ATM promoter through

Sp1. The same study showed a direct association of

Sp1 with the kinase domain of the ATM protein, as

well as the N–terminal, substrate and chromatin-asso-

ciated domains [130]. This evidence suggests that Sp1

is modulated in response to decreased growth signaling

and genotoxic stress, and that ATM may regulate its

own expression by phosphorylating Sp1. ATM has

been shown to modulate transcription of other genes

through Sp1, in addition to other transcription factors,

indicating that this may be a conserved mechanism of

transcriptional regulation in response to genotoxic

stresses [201]. Finally, cells isolated from a patient with

ataxia telangiectasia show disrupted Sp1 regulation

and abnormal responses to EGF signaling, again

235FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

suggesting that ATM is involved in several Sp1-regu-

lated processes related to evading growth suppression

[263].

Resisting cell death and immune escape

Genetic mutation and changes in cellular signaling

within a cancer cell produce a high level of intracellu-

lar stress, which, under normal circumstances, would

promote cell death. In addition, many tumor cells

withstand external signals from the immune system as

it attempts to eliminate cancer cells from normal tissue

populations. In order for tumors to develop, cancer

cells must be able to overcome the intrinsic and extrin-

sic signals that normally induce apoptosis. Regulation

of a number of pro- and anti-apoptotic factors by Sp1

has been investigated, including B-cell CLL/lymphoma

2 (Bcl-2) [227], tumor necrosis factor (ligand) super-

family, member 10 (TRAIL) and its receptors tumor

necrosis factor receptor superfamily, member 10a/b

(DR4/5) [163,225,226,228], myeloid cell leukemia 1

(MCL1) [314], X-linked inhibitor of apoptosis, E3

ubiqutin protein ligase (XIAP) [230], cellular FLICE-

like inhibitory protein (FLIP) [334], Fas ligand (FasL)

[219,231,233] and BCL-2 antagonist/killer 1 (BAK)

[335]. Tumor cells become resistant to extrinsic apop-

totic signals from the immune system by decreasing

the expression of the cell-surface TRAIL receptors

DR4/DR5 and the TRAIL-specific inhibitor FLIP

[336,337]. Sp1-dependent regulation of these factors

not only promotes tumor cell resistance to apoptosis

and immune escape, but may also help promote cancer

cell sensitivity to induction of apoptosis by modulation

of these factors [163]. Most notably, the Sp1-depen-

dent regulation of the anti-apoptotic factor survivin

(baculoviral IAP repeat-containing 5a, BIRC5 gene),

which is closely associated with neoplastic resistance to

cell death, has been thoroughly studied in a number of

tumor cell types [338]. Survivin promotes cell survival

by inhibiting both extrinsic and intrinsic apoptotic

pathways, making it essential for promoting cell sur-

vival in tumors and an important factor in avoiding

immune surveillance.

The BIRC5 promoter contains six putative GC/GT

box targets for Sp factors that are positively regulated

by binding of both Sp1 and Sp3 (Fig. 2) [339]; muta-

tion of key Sp1 sites abolishes basal levels of expres-

sion of the BIRC5 gene [223]. Furthermore, over-

expression of survivin in tumors is directly associated

with an increase in Sp1 protein [219,340].

In response to apoptotic stimuli, survivin levels

decrease in an Sp1-dependent manner. Although this

decrease potentially occurs through loss of p300

recruitment, the underlying mechanism has not been

thoroughly addressed [222]. As mentioned above, a

number of studies have suggested that the interaction

between p300 and Sp1 is important for Sp1-mediated

transcriptional changes [69,71,222,228,330,341–343].Doxorubicin treatment has been shown to suppress

expression of the BIRC5 gene. This suppression may

occur through Sp1/p53-dependent recruitment of

DNMT1 to the BIRC5 promoter [100]. Recruitment

of DNMT1 allows further recruitment of the histone

methyltransferase G9a and the histone deacetylase

HDAC1, and modification of the promoter by his-

tone H3 lysine 9 dimethylation, resulting in epige-

netic silencing of BIRC5. Although this study shows

a direct interaction between the N–terminal portion

of Sp1 and DNMT1, Sp1 remains constitutively

bound to the BIRC5 promoter even in the absence

of doxorubicin treatment, and recruitment of

DNMT1 to the BIRC5 promoter is dependent on

p53 recruitment. DNMT1 has also been shown to

interact with p53, consistent with p53-dependent

recruitment of DNMT1; this dependence is further

supported by the finding that decreases in BIRC5

expression in response to doxorubicin did not occur

in a p53�/� cell line [100]. In addition, we have

found that induction of apoptosis by DNA-

damaging agents, including doxorubicin, induces cas-

pase-mediated cleavage of Sp1 at Asp183. Caspase-

mediated cleavage creates two peptides, a 30 kDa

portion containing the N–terminal inhibitory domain

and the majority of the A domain, and a 70 kDa

fragment containing the remainder of the protein,

including the B domain and the DNA-binding

domain (B. Torabi & J. Azizkhan-Clifford, Drexel

University College of Medicine, unpublished results).

This cleavage was shown to be involved in induction

of apoptosis, and cells expressing a mutant of Sp1

that cannot be cleaved (Sp1 D183A) displayed

increased resistance to apoptosis. Further studies are

underway to determine how Sp1 cleavage affects the

transcription of key apoptotic factors.

Signaling through the ERK1/2 pathway may modu-

late BIRC5 expression. ERK activity negatively regu-

lates BIRC5 expression, and inhibition of ERK

stabilizes survivin in cells treated with the drug querce-

tin, which down-regulates BIRC5 expression [220]. It

appears that quercetin may negatively regulate Sp1

binding to the BIRC5 promoter, but the mechanism

underlying this effect is poorly understood. There is

evidence that ERK-dependent phosphorylation of Sp1

may both positively and negatively regulate transcrip-

tion of specific genes [59,60,73,287]. Signaling through

ERK may modulate Sp1-dependent regulation of

236 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

BIRC5 expression, but more direct evidence is required

to verify whether this is in fact the case.

YM155, which is currently in clinical trials as a

suppressor of BIRC5, functions by decreasing pro-

moter activity of the BIRC5 gene [344]. The effects

of the drug, which binds to the BIRC5 promoter,

appear to be specific to BIRC5. Studies of this drug

suggested that the down-regulation of BIRC5 expres-

sion by YM155 occurs through its binding to

regions that bind Sp1. YM155 does not affect the

Sp1 protein level, but may function by directly or

indirectly, affecting Sp1 localization and binding to

the BIRC5 promoter [344].

Promoting tumor-associated angiogenesis

To fulfill the nutrient and oxygen requirements of

growing tumors, angiogenic pathways are activated. A

large number of cellular factors are associated with

angiogenic pathways, both in tumors and normal tis-

sue. Sp1 is associated with the regulation of a number

of pro- and anti-angiogenic genes, including, but not

limited to, those encoding vascular endothelial growth

factor (VEGF), thrombospondin 1 (TSP–1), platelet-

derived growth factor (PDGF) and urokinase-type

plasminogen activator (uPA) (Table 4) [167,234–236,249,252–254,257,258,345]. VEGF is one of the

most highly studied factors associated with tumor

angiogenesis [346,347], and it is well established that

Sp1 plays an essential role in VEGF regulation.

Early characterization of the VEGF promoter identi-

fied a number of Sp1 binding sites near the transcrip-

tional start site of the VEGF gene (Fig. 2) [243,250].

Initial reports indicated that these Sp1 sites were essen-

tial for VEGF expression in response to PDGF stimu-

lation [237]. Further, regulation of VEGF by Sp1 was

shown to be modulated by interactions with other fac-

tors [236,239,242,249], as well as through kinase sig-

naling [240,244,247].

Under normal cellular conditions, hormone signal-

ing participates in the regulation of VEGF expression.

Estrogen promotes endometrial growth and implanta-

tion through modulation of VEGF [238,239]. Studies

have now shown that association of ERa with the

VEGF promoter is dependent on the Sp1 binding sites

within that promoter [238,239]. This estrogen-depen-

dent signaling is important within the context of breast

cancer. Estrogen stimulates ERa-dependent up-regula-

tion of VEGF in breast cancer cells, and this up-regu-

lation is dependent on the interaction of ERa with

Sp1/Sp3 [249]. In addition, androgen receptor also

associates with Sp1 at the VEGF promoter, and modu-

lates VEGF expression [236]. These findings implicate

Sp1 in hormone-dependent regulation of VEGF expres-

sion.

Several intracellular signaling cascades are associ-

ated with Sp1-dependent activation of VEGF. Reactive

oxygen species (ROS) are associated with activation of

a number of signaling pathways, including Ras/Raf/

MEK/ERK signaling (MEK, MAP kinase-ERK

kinase), that mediate Sp1-dependent up-regulation of

VEGF [247]. ROS signaling appears to modulate Sp1-

but not Sp3-dependent changes [247]. ERK1/2 directly

phosphorylates Sp1 on Thr453 and Thr739, and these

modifications are important for increasing VEGF pro-

moter activity [59].

Hepatocyte growth factor (HGF) signaling, which

modulates PI3K, ERK1/2 and protein kinase Cf(PKCf) pathways to modulate phosphorylation of

Sp1, also promotes VEGF expression [245]. PI3K func-

tions upstream of protein kinase B (Akt), an essential

signaling molecule in growth and proliferative path-

ways. Early studies showed that PI3K function was

associated with VEGF expression. Expression of active

Akt increases VEGF expression through increased

binding of Sp1 to sites in the VEGF promoter [244].

Increased active Akt is also associated with increased

Sp1 phosphorylation, although it is not known

whether this is direct or occurs through other down-

stream signaling kinases [244]. These studies also

showed that Sp1 depletion had no effect on the induc-

tion of VEGF in response to hypoxia, which stimulates

hypoxia inducible factor 1a to modulate VEGF. This

suggests that there are multiple mechanisms for modu-

lation of VEGF transcription that are both Sp1-depen-

dent and -independent.

Sp1 has been shown to interact with PKCf, whichmodifies its DNA-binding domain [121]. This modifi-

cation directly affects Sp1-dependent regulation of

VEGF, as well as the ability of Sp1 to interact with

Von Hippel–Lindau protein (VHL), a tumor suppres-

sor protein that negatively modulates VEGF expression

[121]. VHL mutations are found in tumors associated

with von Hippel–Lindau disease, as well as in other

tumors. Loss of VHL is associated with Sp1-dependent

up-regulation of VEGF mRNA [242]. It is now under-

stood that VHL and Sp1 interact with one another,

and it has been proposed that VHL may suppress Sp1

activity at the VEGF promoter. However, whether

VHL affects Sp1 binding or Sp1 transactivation of

VEGF has not been established.

Sp1 O–GlcNAcylation has also been suggested to

play a role in the regulation of VEGF [246], and this

supported by our work with retinal cells demonstrating

that VEGF is induced by glucose via a mechanism

requiring Sp1 and its modification by O-linked

237FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

glycosylation [244,348]. Increased O–GlcNAcylation of

proteins has been linked to the altered metabolic state

in cancer cells [349,350], as well as diabetic pathologies

[351,352]. Multiple reports have shown that O–linkedglycosylation of Sp1 by O–GlcNac transferase results

in modulation of Sp1 stability and transcriptional

activity. These data support a role for O–GlcNAcyla-

tion of Sp1 in the transcriptional modulation of

VEGF, as well as that of other genes encoding factors

involved in tumorigenesis and diabetes-associated

pathologies.

Activation of tumor cell invasion and metastasis

High levels of Sp1 protein have been show to correlate

with cancer cell migration and metastasis in a number

of tumor models and patient samples, including gas-

tric, pancreatic and breast cancers [158,160,167]. In

contrast, one group showed a negative correlation

between invasiveness and levels of Sp1 protein in late-

stage lung cancer [169]. These studies underscore the

importance of Sp1 in the regulation of a number of

pro- and anti-invasive factors, including matrix metal-

loproteinases (MMPs), membrane type 1 matrix metal-

loproteinase (MT1–MMP), the MMP inhibitor

reversion-inducing cysteine-rich protein with Kazal

motifs (RECK), E–cadherin and integrin a5 (Table 3)

[60,90,137,285,286,296,299,353].

Matrix metalloproteinases

MMPs participate in remodeling of the extracellular

matrix to facilitate cell migration and ultimately

metastasis. Sp1 has been implicated in the regulation

of MMPs and MMP-like proteins. An early study sug-

gested that Sp1 was a key regulator in the transcrip-

tion of MMP2 [299]. Brg1, a component of the

mammalian SWI/SNF (SWitch/Sucrose NonFerment-

able) complex that is important for activation of the

MMP2 promoter, is probably recruited to the MMP2

promoter through interaction with Sp1, increasing

expression of MMP2 mRNA and subsequently MMP2

protein [90]. This recruitment was associated with

decreased Sp3 binding to the promoter of the gene

encoding MMP2, which, in this context, may have

been acting as an inhibitor of transactivation. This

study suggested a mechanism by which Sp1 positively

regulates MMP2 expression.

The CDK inhibitor p16 was also identified as a neg-

ative regulator of Sp1-mediated MMP2 expression.

p16 expression prevented cyclin A from interacting

with Sp1, probably resulting in decreased CDK-medi-

ated Sp1 phosphorylation [137]. Studies of cyclin A/

CDK effects on Sp1 activity support a pathway in

which CDK2 phosphorylates Sp1 on Ser59, thereby

enhancing the transcriptional activity of Sp1 [52].

These data suggest that p16 inhibits CDK-dependent

phosphorylation of Sp1, resulting in decreased tran-

scription of Sp1 target genes, including MMP2. In

cancer cells, loss of p16 may mediate increased MMP2

gene expression via increased Sp1 transactivation. In

addition, Sp1 regulates p16 expression, as discussed

below.

The cell-cycle regulator S-phase kinase-associated

protein 2 (SKP2) has also been shown to influence

Sp1-mediated MMP expression. SKP2 is an E3 ubiqu-

itin ligase that is over-expressed in cancers. It has been

shown to modulate the stability of the CDK inhibitor

p27, as well as to promote increased invasion and

migration of cancer cells [284]. Over-expression of

SKP2 is associated with increases in both MMP2 and

MMP9, and expression of both proteases requires the

transcriptional activity of Sp1 [284]. This suggests two

possible mechanisms by which SKP2 promotes Sp1-

dependent transcription. First, SKP2 may directly

affect CDK-dependent phosphorylation of Sp1

through modulation of the p27 CDK inhibitor. In this

scenario, over-expression of SKP2 decreases p27

expression, thereby enhancing CDK-dependent Sp1

phosphorylation and thus MMP expression. Alterna-

tively, SKP2 may interact with Sp1 and directly modu-

late its DNA binding, as previously demonstrated

[117]. Further work is required to assess the relation-

ship between Sp1 and SKP2.

In addition to regulating the expression of secreted

MMP molecules, Sp1 also regulates the expression of

MT1–MMP, which encodes an MMP that contains a

transmembrane domain for membrane localization.

Functionally, MT1–MMP proteins also stimulate

migration and metastasis. Conflicting reports suggest

that Sp1 both positively and negatively regulates

expression of MT1–MMP [286,354]. Gene expression

in the endothelial cells of the microvasculature is mod-

ulated by mechanical/shear stress. Shear stress has

been shown to increase Sp1 phosphorylation in these

cells and to enhance Sp1 binding to the MT1–MMP

promoter [286]. This result conflicted with previous

studies showing that increased MT1–MMP expression

requires displacement of Sp1 from the promoter via

binding of the transcription factor Early growth

response 1 (EGR1) [354]. More recently, studies have

shown that MT1–MMP is highly expressed in invasive

prostate tumor cell lines and is positively regulated by

Sp1 [285], further complicating our understanding of

how Sp1 regulates MT1–MMP expression. Addition-

ally, Sp1-mediated expression of MT1–MMP is posi-

238 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

tively regulated by signaling through b-akt murine thy-

moma viral oncogene homolog 1 (AKT), JNK and

ERK pathways [285]. The initial studies on Sp1-medi-

ated MT1–MMP regulation were performed in rat

microvascular cells, while more recent studies have uti-

lized human cancer cells lines; it is possible that the

differences in results are due to differences between

normal and cancer cells, cell type or species. It is cer-

tainly not without precedence that Sp1 may both acti-

vate and inhibit, depending on the cellular context.

RECK

In addition to regulating MMPs directly, Sp1 also

transcriptionally regulates inhibitors of MMPs, such as

RECK. Initial observations suggested that Sp1 nega-

tively regulates RECK [287–289] through ERK-depen-

dent phosphorylation of Sp1 (on Thr453 and Thr739),

which increases HDAC1 recruitment to the RECK

promoter, and inhibits RECK transcription [60]. ERK

activity was enhanced by the activity of the EGFR

receptor family member HER-2/neu [60], which is fre-

quently over-expressed in breast cancer.

As with MT1–MMP, Sp1 appears to both positively

and negatively regulate the RECK gene: a common

theme in Sp1-dependent transcriptional regulation. Sp1

is known to recruit both activators and repressors of

transcription to gene promoters, depending on the cel-

lular context and post-translational modifications pres-

ent on the Sp1 protein. This allows Sp1 to

differentially modulate gene expression depending

upon the cellular environment.

E–cadherin

Sp1 may also regulate anti-migratory factors such as

E–cadherin. E–cadherin is a cell–cell adhesion glyco-

protein that is dysregulated in a number of cancers

[355–357]. Decreases in E–cadherin are associated with

increased cellular proliferation and metastasis. The

potential contribution of Sp1 to regulation of the E–cadherin gene CDH1 has been described previously

[292,294]. More recent studies in a mouse lung cancer

model and in human lung cancer cells suggest that Sp1

over-expression increases expression of the CDH1

gene. This likely explains the down-regulation of Sp1

in late-stage lung cancer cells, which would support

metastasis by decreasing cell–cell adhesion [169]. The

role of Sp1 as a positive regulator of E–cadherin is

supported by the finding that expression of the E6

protein in HPV16 decreases Sp1 binding to and trans-

activation of the CDH1 promoter [291]. Finally, CpG

island methylation may modulate expression of the

CDH1 gene in neoplastic tissue [358]. Sp1 participates

in protecting CpG islands from methylation in normal

tissue [293]. The involvement of Sp1 in CpG island

regulation is discussed below.

Clearly, Sp1 is necessary for regulation of transcrip-

tion of a number of factors that participate in cell

migration and invasion. The conflicting reports proba-

bly represent differences between normal and cancer

cells, as well as differences among cancer models. Sp1

positively and negatively regulates both pro- and anti-

invasive factors. Additionally, the expression and func-

tion of all of these factors are further influenced by a

number of other cellular pathways, ultimately resulting

in a balanced cellular phenotype.

Inflammation

Inflammation is recognized as a critical component of

tumor initiation and progression. Infection, chronic

irritation and inflammation contribute to tumorigene-

sis. The tumor microenvironment is infiltrated with

inflammatory cells that produce a variety of cytotoxic

mediators, such as ROS, serine and cysteine proteases,

MMPs, tumor necrosis factor a, interleukins, interfer-ons and enzymes, all of which contribute to inflamma-

tory proliferation, survival and migration. As

discussed above, immune surveillance and secretion of

inflammatory cytokines increases the inflammatory

environment of the tumor. Increased proliferation and

metabolism also adds to the inflammatory environ-

ment of a tumor. Thus inflammation acts as a tumor

promoter in association with many carcinogens.

The relationship between Sp1 and inflammatory sig-

naling is well established. Use of non-steriodal anti-

inflammatory drugs (NSAIDs) in cancer prevention

and in the treatment of various tumor types has been

proposed. Decreased inflammatory signaling in

response to NSAID treatment has been associated with

decreases in Sp1 protein levels, as well as in Sp1-depen-

dent transcriptional activity [234,251,298,359–361]. One

study addressing the direct effect of NSAIDs on Sp1

transcriptional activity demonstrated inhibition of

ERK activity, which decreased phosphorylation of Sp1

and activation of its downstream effector MMP2 [298].

Additionally, treatment with the NSAIDs tolfenamic

acid and aspirin has been shown to decrease Sp1-

dependent transcription. These associations provide a

direct link between the regulation of Sp1 by inflamma-

tory signaling and the regulation of essential tumor-

promoting factors. An interesting epidemiological

study supported the notion that Sp1 is intimately asso-

ciated with inflammation [362]. Colorectal cancer

patients with a wild-type COX2 promoter benefited

239FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

from NSAIDs and the associated decreased expression

of cytochrome c oxidase subunit II (COX2), whereas a

G?C mutation in the COX2 promoter, which pre-

vented Sp1 binding, allowed no beneficial effect of

NSAIDs in these patients [362].

Sp1 has also been implicated in inflammatory signal-

ing associated with a number of pro-inflammatory

genes. Activation of osteopontin in hepatitis C infec-

tion, which has been linked to hepatocellular carci-

noma, has been shown to be Sp1-dependent;

osteopontin production is at least in part responsible

for the epithelial/mesenchymal transition in progres-

sion of hepatocellular carcinoma. Mina, an epigenetic

gene regulatory protein known to function in multiple

physiological and pathological contexts, including pul-

monary inflammation, cell proliferation, cancer and

immunity, requires Sp1 for its induction [363]. T-cell-

specific T-box transcription factor (TBET), a regulator

of interferon c in natural killer cells, T cells, B cells

and dendritic cells, that responds to pro-inflammatory

cytokines, is regulated by Sp1 binding to its promoter.

Mithramycin, which blocks Sp1 binding, was shown to

down-regulate TBET expression, thereby implicating

Sp1 in the inflammatory response [364].

Acute inflammation leads to an increase in zinc

transport into the cell, probably to help protect

against ROS produced by surrounding inflammatory

cells. This influx in zinc also promotes cytokine pro-

duction and further cytokine signaling. In contrast to

acute inflammation, chronic inflammation leads to

decreases in intracellular zinc, which may contribute

to apoptotic and senescent signaling, as well as geno-

mic instability associated with chronic inflammation.

Decreased Sp1 protein levels and apoptotic signaling

have been associated with cytoplasmic sequestration of

zinc [360]. Zinc is a co-factor for Sp1 DNA binding,

and affects Sp1 protein levels and transcriptional

activity. Zinc is also essential for maintaining the

proper structure of the Sp1 zinc fingers [365]. The

DNA binding ability of Sp1 and its nuclear localiza-

tion are dependent on the presence of zinc [20,365].

Therefore, modulation of cellular zinc levels directly

affects Sp1 activity. The increase in zinc associated

with acute inflammation may activate Sp1, and may

help promote tumor-promoting events such as tran-

scriptional up-regulation of factors required for sur-

vival. Zinc levels are elevated in tumors, as well as in

the surrounding tumor micro-environment, but remain

low in the circulating blood serum of cancer patients

[366]. Studies addressing how changes in zinc levels in

response to inflammatory signals affect Sp1 activity

may provide interesting insight into how Sp1 promotes

tumor formation.

Sp1 functions in maintaining genomic stability

There is a strong body of evidence supporting the

idea that genomic instability and DNA damage pro-

mote genetic changes leading to oncogenic transfor-

mation. Cells must maintain the ability to repair

DNA lesions as simple as base modifications and

replication mismatches, to more threatening DNA

lesions, such as single- and double-strand breaks in

order to protect the integrity of the genome. Proper

recognition and processing of DNA lesions is essen-

tial for maintaining DNA sequence fidelity, and for

preventing large-scale rearrangements of the genome

itself. In addition, proper chromosome segregation

must be maintained during mitosis, through mecha-

nisms at the centromere and kinetochore, the mitotic

spindle and the centrosomes. Errors in chromosome

segregation result in chromosomal instability or the

loss or gain of whole chromosomes (aneuploidy) dur-

ing mitosis.

Sp1 may participate in whole chromosome stability

in mitotically dividing cells. Sp1 localizes to the cen-

trosome in interphase cells [367]. Sp1 depletion leads

to aneuploidy and the formation of micronuclei,

probably due to an increase in centrosome number

as seen in cells lacking full-length Sp1 [367]. This

phenotype appears to be dependent on the first 183

amino acids of the protein, because a mutant Sp1

lacking these residues is unable to rescue the multi-

centrosome phenotype [367].

Of the eight Sp family members, Sp1 is unique in

that it contains SQ/TQ cluster domains within its two

transactivation domains; SQ/TQ cluster domains are

targets for phosphorylation by PI3K-like kinases,

including ATM. Sp1 is phosphorylated by ATM in

response to a variety of DNA-damaging agents and

viral infection [49,54,55]. This phosphorylation, which

may be visualized as a 10 kDa electrophoretic mobility

shift from 95 to 105 kDa, or detection by an antibody

that recognizes Sp1 phosphorylated on Ser101, occurs

largely on serines and is rapid and sustained for sev-

eral hours after DNA damage. The disappearance of

this phosphorylation is coincident with the disappear-

ance of the phosphorylated form of Histone H2A vari-

ant X (cH2AX), and presumably the completion of

DNA repair. The mobility shift is probably due to

phosphorylation of several serine residues, and phos-

phorylation of Ser101 appears to be a priming site,

based on the complete absence of phosphorylation of

a mutant of Sp1, S101A, in response to ionizing radia-

tion [54]. It is important to note that hyper-phosphory-

lation of Sp1 is also associated with proliferation

signaling [368,369], although the residues that are

240 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

phosphorylated during proliferation and damage are

not all known. Importantly, ATM promotes Sp1-

dependent increases in gene transcription even in

serum-starved cells [370].

Cells depleted of Sp1 show increased sensitivity to

DNA damage at low levels of ionizing radiation, and

this sensitivity can be rescued by expression of wild-

type Sp1, but not by expression of the phosphorylation

mutant S101A [54]. This suggests that Sp1 phosphory-

lation is required for its function in response to dam-

age. Sp1 also regulates transcription of a number of

factors that respond to DNA damage. Interestingly,

microarray analysis of cells lacking Sp1 compared to

normal cells, and of cells exposed to ionizing radiation,

both in the presence and absence of Sp1, suggests that,

in response to ionizing radiation, a relatively small

number of genes are up regulated in an Sp1-dependent

manner (K. Beishline & J. Azizkhan-Clifford, Drexel

University College of Medicine, Paules et al., National

Institute of Environmental Health Sciences, unpub-

lished results). Although Sp1 depletion alone is associ-

ated with a large number of gene expression changes,

evidence suggests that the response of Sp1 to genotoxic

stress may not be primarily transcriptional. Most nota-

bly, nuclear lysates isolated from cells after DNA dam-

age showed no increase in Sp1 binding to its cognate

binding site [49]. In addition, unlike a number of other

transcriptional regulators modulated by ATM in

response to DNA damage, Sp1 is recruited to sites of

DNA breaks (ionizing radiation-induced foci), where it

co-localizes with phosphorylated ATM and other

repair factors [74]. The presence of Sp1 at these sites

of damage, an observed decrease in double-strand

break repair in cells depleted of Sp1, and the rescue of

this repair defect by a transcriptionally inert portion of

the protein [74] suggest that Sp1 plays a role in the

repair process, rather than a transcriptional role, in

response to DNA damage-associated stress signals

Sp1, DNA methylation and cancer

Methylation of CpG sequences in DNA was first iden-

tified in microorganisms in the 1960s, and has been an

area of intensive investigation in eukaryotic cells since

the 1980s, when its importance in vertebrate develop-

ment and genetics was recognized [371]. Clusters of

CpG sequences, known as CpG islands, were mapped

to multiple regions throughout the genome, including

the 50 ends of housekeeping genes and a number of tis-

sue-specific genes. After years of research, it has

become apparent that DNA methylation and CpG

islands are essential for the epigenetic regulation of

genes throughout embryonic development, cellular

differentiation, and a number of disease states, includ-

ing cancer [372].

Over 20 years ago, researchers showed that CpG

islands often contained Sp1 binding sites. These

sequences, present in regulatory regions of housekeep-

ing genes and other essential genomic loci, were pro-

tected from hyper-methylation during stages of

embryogenesis, when methylation patterns are gener-

ated [373]. Further analysis of DNA methylation in

mammalian systems suggested that changes in methyl-

ation of DNA act to activate or repress transcription,

supporting a role for Sp1 in protecting the 50 regionsof housekeeping and other genes from being silenced

epigenetically [374]. Furthermore, there is evidence to

suggest that when methylation spreads into Sp1 bind-

ing elements, it inhibits binding of Sp1 and further

contributes to the silencing of transcription [375]. All

of these data suggest a role for Sp1 not only in the

dynamic regulation of gene transcription, but also in

epigenetic regulation through controlling long-term

gene silencing by DNA methylation.

The understanding of how Sp1 and its family mem-

bers contribute to the methylation status of CpG

islands is derived from a large number of studies

addressing how changes in methylation of oncogenes

and tumor suppressors in neoplastic tissues may be

attributed to Sp1. An example is methylation of the

gene RASSF1A, encoding an Sp1-regulated tumor sup-

pressor that is frequently methylated in cancers [376].

In RASSF1A, changes in histone modifications lead to

eviction of Sp1 from the promoter region, and the loss

of Sp1 allows CpG methylation of the promoter. In

proliferating epithelial cells in vitro, de-acetylation of

histones at the RASSF1A promoter and replacement

of the acetylation marks by histone H3 lysine 9 trime-

thylation accompanied decreased transactivation by

Sp1. Cells within the population that were able to

overcome senescent barriers were found to maintain

suppression of RASSF1A gene expression through

CpG methylation of the Sp1 binding sites. The pro-

moters of other factors important in cancer signaling

reportedly have altered methylation states during

tumorigenesis [332,333,377], suggesting that the

increased levels of Sp1 observed in cancer cells may

not overcome the suppression of selective genes

through promoter methylation. It is important to note

that current studies addressing methylation patterns in

senescent cells compared to cancer cells suggest that

only very specific patches of methylation are important

in cancer phenotypes, while general hypomethylation

of the majority of the genome is a common trait in

both senescent and cancer cells [378]. In senescent

cells, these patterns suppress the expression of prolifer-

241FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

ative genes but protect the continued expression of

necessary housekeeping genes. These patterns suggest

protection of CpG islands within promoter regions but

addition of methylation marks on upstream and down-

stream elements [378]. These events correlate well with

findings that Sp1 protects from methylation CpGs to

which it is bound within promoters, while also sup-

porting studies in which Sp1 was found to promote

CpG methylation of surrounding regions through

interaction with factors such as DNMT1 [100,378].

The literature supports a strong role for Sp1 in deter-

mining the epigenetic status of both normal and cancer

cells, and further studies are required to shed addi-

tional light on this complex regulation.

Concluding remarks

The studies discussed in this review, which focuses on

the role of Sp1 in cancer, demonstrate that Sp1 is a

highly regulated transcription factor that is involved in

regulating expression of a large number of genes that

contribute to the ‘hallmarks of cancer’. Sp1 both acti-

vates and suppresses the expression of a number of

essential oncogenes and tumor suppressors, as well as

genes involved in a number of essential cellular func-

tions, including proliferation, differentiation, the DNA

damage response, apoptosis, senescence and angiogen-

esis. Given the apparently opposing effects of Sp1, for

example in regulating oncogenes and suppressors, pro-

survival genes and pro-apoptotic genes, a more com-

plete mechanistic understanding of its activities on dif-

ferent promoters and under different cellular

conditions is essential before it may be exploited as a

target for cancer treatment. Additionally, efforts are

required to develop inhibitors that inactivate specific

functions of Sp1. Further study of its non-transcrip-

tional functions is also essential for understanding its

full role in tumorigenesis.

Author contributions

Dr. Beishline and Dr. Clifford worked together to

develop the outline and gather the literature.

References

1 Dynan WS & Tjian R (1983) The promoter-specific

transcription factor Sp1 binds to upstream sequences

in the SV40 early promoter. Cell 35, 79–87.2 Dynan WS & Tjian R (1983) Isolation of transcription

factors that discriminate between different promoters

recognized by RNA polymerase II. Cell 32, 669–680.

3 Black AR, Black JD & Azizkhan-Clifford J (2001) Sp1

and kruppel-like factor family of transcription factors

in cell growth regulation and cancer. J Cell Physiol

188, 143–160.4 Cawley S, Bekiranov S, Ng HH, Kapranov P, Sekinger

EA, Kampa D, Piccolboni A, Sementchenko V, Cheng

J, Williams AJ et al. (2004) Unbiased mapping of

transcription factor binding sites along human

chromosomes 21 and 22 points to widespread

regulation of noncoding RNAs. Cell 116, 499–509.5 Oleaga C, Welten S, Belloc A, Sole A, Rodriguez L,

Mencia N, Selga E, Tapias A, Noe V & Ciudad CJ

(2012) Identification of novel Sp1 targets involved in

proliferation and cancer by functional genomics.

Biochem Pharmacol 84, 1581–1591.6 Gilmour J, Assi SA, Jaegle U, Kulu D, van de Werken

H, Clarke D, Westhead DR, Philipsen S & Bonifer C

(2014) A crucial role for the ubiquitously expressed

transcription factor Sp1 at early stages of

hematopoietic specification. Development 141, 2391–2401.

7 Bouwman P & Philipsen S (2002) Regulation of the

activity of Sp1-related transcription factors. Mol Cell

Endocrinol 195, 27–38.8 Kaczynski J, Cook T & Urrutia R (2003) Sp1- and

Kruppel-like transcription factors. Genome Biol 4, 206.

9 Thiesen HJ & Bach C (1990) Target Detection Assay

(TDA): a versatile procedure to determine DNA

binding sites as demonstrated on SP1 protein. Nucleic

Acids Res 18, 3203–3209.10 Kingsley C & Winoto A (1992) Cloning of GT box-

binding proteins: a novel Sp1 multigene family

regulating T-cell receptor gene expression. Mol Cell

Biol 12, 4251–4261.11 Hagen G, Dennig J, Preiss A, Beato M & Suske G

(1995) Functional analyses of the transcription factor

Sp4 reveal properties distinct from Sp1 and Sp3. J Biol

Chem 270, 24989–24994.12 Crossley M, Whitelaw E, Perkins A, Williams G,

Fujiwara Y & Orkin SH (1996) Isolation and

characterization of the cDNA encoding BKLF/TEF-2,

a major CACCC-box-binding protein in erythroid cells

and selected other cells. Mol Cell Biol 16, 1695–1705.13 Matsumoto N, Laub F, Aldabe R, Zhang W, Ramirez

F, Yoshida T & Terada M (1998) Cloning the cDNA

for a new human zinc finger protein defines a group of

closely related Kruppel-like transcription factors. J

Biol Chem 273, 28229–28237.14 Shields JM & Yang VW (1998) Identification of the

DNA sequence that interacts with the gut-enriched

Kruppel-like factor. Nucleic Acids Res 26, 796–802.15 Nagaoka M, Shiraishi Y & Sugiura Y (2001) Selected

base sequence outside the target binding site of zinc

finger protein Sp1. Nucleic Acids Res 29, 4920–4929.

242 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

16 Kadonaga JT, Jones KA & Tjian R (1986) Promoter-

specific activation of RNA polymerase II transcription

by Sp1. Trends Biochem Sci 11, 20–23.17 Kriwacki RW, Schultz SC, Steitz TA & Caradonna JP

(1992) Sequence-specific recognition of DNA by zinc-

finger peptides derived from the transcription factor

Sp1. Proc Natl Acad Sci USA 89, 9759–9763.18 Thiesen HJ & Schroder B (1991) Amino acid

substitutions in the SP1 zinc finger domain alter the

DNA binding affinity to cognate SP1 target site.

Biochem Biophys Res Commun 175, 333–338.19 Ito T, Azumano M, Uwatoko C, Itoh K & Kuwahara

J (2009) Role of zinc finger structure in nuclear

localization of transcription factor Sp1. Biochem

Biophys Res Commun 380, 28–32.20 Ito T, Kitamura H, Uwatoko C, Azumano M, Itoh K

& Kuwahara J (2010) Interaction of Sp1 zinc finger

with transport factor in the nuclear localization of

transcription factor Sp1. Biochem Biophys Res

Commun 403, 161–166.21 Su W, Jackson S, Tjian R & Echols H (1991) DNA

looping between sites for transcriptional activation:

self-association of DNA-bound Sp1. Genes Dev 5, 820–826.

22 Mastrangelo IA, Courey AJ, Wall JS, Jackson SP &

Hough PV (1991) DNA looping and Sp1 multimer

links: a mechanism for transcriptional synergism and

enhancement. Proc Natl Acad Sci USA 88, 5670–5674.23 Dennig J, Beato M & Suske G (1996) An inhibitor

domain in Sp3 regulates its glutamine-rich activation

domains. EMBO J 15, 5659–5667.24 Hagen G, Muller S, Beato M & Suske G (1994) Sp1-

mediated transcriptional activation is repressed by Sp3.

EMBO J 13, 3843–3851.25 Majello B, De Luca P, Hagen G, Suske G & Lania L

(1994) Different members of the Sp1 multigene family

exert opposite transcriptional regulation of the long

terminal repeat of HIV-1. Nucleic Acids Res 22, 4914–4921.

26 Hammill D, Jain N, Armstrong S & Mueller CR

(2005) The D-domain of Sp3 modulates its protein

levels and activation of the p21(CIP1/WAF1)

promoter. Biochem Biophys Res Commun 335, 377–384.

27 Wierstra I (2008) Sp1: emerging roles–beyondconstitutive activation of TATA-less housekeeping

genes. Biochem Biophys Res Commun 372, 1–13.28 Williams AO, Isaacs RJ & Stowell KM (2007) Down-

regulation of human topoisomerase IIalpha expression

correlates with relative amounts of specificity factors

Sp1 and Sp3 bound at proximal and distal promoter

regions. BMC Mol Biol 8, 36.

29 Li L & Davie JR (2010) The role of Sp1 and Sp3 in

normal and cancer cell biology. Ann Anat 192, 275–283.

30 Marin M, Karis A, Visser P, Grosveld F & Philipsen S

(1997) Transcription factor Sp1 is essential for early

embryonic development but dispensable for cell

growth and differentiation. Cell 89, 619–628.31 Kruger I, Vollmer M, Simmons DG, Elsasser HP,

Philipsen S & Suske G (2007) Sp1/Sp3 compound

heterozygous mice are not viable: impaired

erythropoiesis and severe placental defects. Dev Dyn

236, 2235–2244.32 Gollner H, Dani C, Phillips B, Philipsen S & Suske G

(2001) Impaired ossification in mice lacking the

transcription factor Sp3. Mech Dev 106, 77–83.33 Ma W, Horvath GC, Kistler MK & Kistler WS (2008)

Expression patterns of SP1 and SP3 during mouse

spermatogenesis: SP1 down-regulation correlates with

two successive promoter changes and translationally

compromised transcripts. Biol Reprod 79, 289–300.34 Wimmer EA, Frommer G, Purnell BA & Jackle H

(1996) buttonhead and D-Sp1: a novel Drosophila gene

pair. Mech Dev 59, 53–62.35 Athanikar JN, Sanchez HB & Osborne TF (1997)

Promoter selective transcriptional synergy mediated by

sterol regulatory element binding protein and Sp1: a

critical role for the Btd domain of Sp1. Mol Cell Biol

17, 5193–5200.36 Kadonaga JT, Courey AJ, Ladika J & Tjian R (1988)

Distinct regions of Sp1 modulate DNA binding and

transcriptional activation. Science 242, 1566–1570.37 Courey AJ & Tjian R (1988) Analysis of Sp1 in vivo

reveals multiple transcriptional domains, including a

novel glutamine-rich activation motif. Cell 55, 887–898.

38 Oka S, Shiraishi Y, Yoshida T, Ohkubo T, Sugiura Y

& Kobayashi Y (2004) NMR structure of transcription

factor Sp1 DNA binding domain. Biochemistry 43,

16027–16035.39 Narayan VA, Kriwacki RW & Caradonna JP (1997)

Structures of zinc finger domains from transcription

factor Sp1. Insights into sequence-specific protein-

DNA recognition. J Biol Chem 272, 7801–7809.40 Infantino V, Convertini P, Iacobazzi F, Pisano I,

Scarcia P & Iacobazzi V (2011) Identification of a

novel Sp1 splice variant as a strong transcriptional

activator. Biochem Biophys Res Commun 412, 86–91.41 Jackson SP, MacDonald JJ, Lees-Miller S & Tjian R

(1990) GC box binding induces phosphorylation of

Sp1 by a DNA-dependent protein kinase. Cell 63, 155–165.

42 Jackson SP & Tjian R (1988) O-glycosylation of

eukaryotic transcription factors: implications for

mechanisms of transcriptional regulation. Cell 55, 125–133.

43 Rigbolt KT, Prokhorova TA, Akimov V, Henningsen

J, Johansen PT, Kratchmarova I, Kassem M, Mann

M, Olsen JV & Blagoev B (2011) System-wide

243FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

temporal characterization of the proteome and

phosphoproteome of human embryonic stem cell

differentiation. Sci Signal 4, rs3.

44 Olsen JV, Vermeulen M, Santamaria A, Kumar C,

Miller ML, Jensen LJ, Gnad F, Cox J, Jensen TS,

Nigg EA et al. (2010) Quantitative phosphoproteomics

reveals widespread full phosphorylation site occupancy

during mitosis. Sci Signal 3, ra3.

45 Spengler ML, Guo LW & Brattain MG (2008)

Phosphorylation mediates Sp1 coupled activities of

proteolytic processing, desumoylation and degradation.

Cell Cycle 7, 623–630.46 Spengler ML & Brattain MG (2006) Sumoylation

inhibits cleavage of Sp1 N-terminal negative regulatory

domain and inhibits Sp1-dependent transcription. J

Biol Chem 281, 5567–5574.47 Chuang JY, Wang YT, Yeh SH, Liu YW, Chang WC

& Hung JJ (2008) Phosphorylation by c-Jun NH2-

terminal kinase 1 regulates the stability of

transcription factor Sp1 during mitosis. Mol Biol Cell

19, 1139–1151.48 Wang YT, Yang WB, Chang WC & Hung JJ (2011)

Interplay of posttranslational modifications in Sp1

mediates Sp1 stability during cell cycle progression. J

Mol Biol 414, 1–14.49 Iwahori S, Yasui Y, Kudoh A, Sato Y, Nakayama S,

Murata T, Isomura H & Tsurumi T (2008)

Identification of phosphorylation sites on transcription

factor Sp1 in response to DNA damage and its

accumulation at damaged sites. Cell Signal 20, 1795–1803.

50 Vicart A, Lefebvre T, Imbert J, Fernandez A & Kahn-

Perles B (2006) Increased chromatin association of Sp1

in interphase cells by PP2A-mediated

dephosphorylations. J Mol Biol 364, 897–908.51 Kim HS & Lim IK (2009) Phosphorylated

extracellular signal-regulated protein kinases 1 and 2

phosphorylate Sp1 on serine 59 and regulate cellular

senescence via transcription of p21Sdi1/Cip1/Waf1. J

Biol Chem 284, 15475–15486.52 Fojas de Borja P, Collins NK, Du P, Azizkhan-

Clifford J & Mudryj M (2001) Cyclin A-CDK

phosphorylates Sp1 and enhances Sp1-mediated

transcription. EMBO J 20, 5737–5747.53 Beausoleil SA, Villen J, Gerber SA, Rush J & Gygi SP

(2006) A probability-based approach for high-

throughput protein phosphorylation analysis and site

localization. Nat Biotechnol 24, 1285–1292.54 Olofsson BA, Kelly CM, Kim J, Hornsby SM &

Azizkhan-Clifford J (2007) Phosphorylation of Sp1 in

response to DNA damage by ataxia telangiectasia-

mutated kinase. Mol Cancer Res 5, 1319–1330.55 Iwahori S, Shirata N, Kawaguchi Y, Weller SK, Sato

Y, Kudoh A, Nakayama S, Isomura H & Tsurumi T

(2007) Enhanced phosphorylation of transcription

factor sp1 in response to herpes simplex virus type 1

infection is dependent on the ataxia telangiectasia-

mutated protein. J Virol 81, 9653–9664.56 Chun RF, Semmes OJ, Neuveut C & Jeang KT (1998)

Modulation of Sp1 phosphorylation by human

immunodeficiency virus type 1 Tat. J Virol 72, 2615–2629.

57 Zheng XL, Matsubara S, Diao C, Hollenberg MD &

Wong NC (2001) Epidermal growth factor induction

of apolipoprotein A-I is mediated by the Ras-MAP

kinase cascade and Sp1. J Biol Chem 276, 13822–13829.

58 Bonello MR & Khachigian LM (2004) Fibroblast

growth factor-2 represses platelet-derived growth

factor receptor-alpha (PDGFR-alpha) transcription via

ERK1/2-dependent Sp1 phosphorylation and an

atypical cis-acting element in the proximal PDGFR-

alpha promoter. J Biol Chem 279, 2377–2382.59 Milanini-Mongiat J, Pouyssegur J & Pages G (2002)

Identification of two Sp1 phosphorylation sites for

p42/p44 mitogen-activated protein kinases: their

implication in vascular endothelial growth factor gene

transcription. J Biol Chem 277, 20631–20639.60 Hsu MC, Chang HC & Hung WC (2006) HER-2/neu

represses the metastasis suppressor RECK via ERK

and Sp transcription factors to promote cell invasion.

J Biol Chem 281, 4718–4725.61 Roos MD, Su K, Baker JR & Kudlow JE (1997) O

glycosylation of an Sp1-derived peptide blocks known

Sp1 protein interactions. Mol Cell Biol 17, 6472–6480.62 Yang X, Su K, Roos MD, Chang Q, Paterson AJ &

Kudlow JE (2001) O-linkage of N-acetylglucosamine

to Sp1 activation domain inhibits its transcriptional

capability. Proc Natl Acad Sci USA 98, 6611–6616.63 Rickers A, Peters N, Badock V, Beyaert R,

Vandenabeele P, Dorken B & Bommert K (1999)

Cleavage of transcription factor SP1 by caspases

during anti-IgM-induced B-cell apoptosis. Eur J

Biochem 261, 269–274.64 Majumdar G, Harrington A, Hungerford J, Martinez-

Hernandez A, Gerling IC, Raghow R & Solomon S

(2006) Insulin dynamically regulates calmodulin gene

expression by sequential o-glycosylation and

phosphorylation of sp1 and its subcellular

compartmentalization in liver cells. J Biol Chem 281,

3642–3650.65 Zhang Y, Liao M & Dufau ML (2006)

Phosphatidylinositol 3-kinase/protein kinase Czeta-

induced phosphorylation of Sp1 and p107 repressor

release have a critical role in histone deacetylase

inhibitor-mediated derepression [corrected] of

transcription of the luteinizing hormone receptor gene.

Mol Cell Biol 26, 6748–6761.66 Dephoure N, Zhou C, Villen J, Beausoleil SA,

Bakalarski CE, Elledge SJ & Gygi SP (2008) A

244 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

quantitative atlas of mitotic phosphorylation. Proc

Natl Acad Sci USA 105, 10762–10767.67 Armstrong SA, Barry DA, Leggett RW & Mueller CR

(1997) Casein kinase II-mediated phosphorylation of

the C terminus of Sp1 decreases its DNA binding

activity. J Biol Chem 272, 13489–13495.68 Tan NY, Midgley VC, Kavurma MM, Santiago FS,

Luo X, Peden R, Fahmy RG, Berndt MC, Molloy MP

& Khachigian LM (2008) Angiotensin II-inducible

platelet-derived growth factor-D transcription requires

specific Ser/Thr residues in the second zinc finger

region of Sp1. Circ Res 102, e38–e51.69 Hung JJ, Wang YT & Chang WC (2006) Sp1

deacetylation induced by phorbol ester recruits p300 to

activate 12(S)-lipoxygenase gene transcription. Mol

Cell Biol 26, 1770–1785.70 Waby JS, Chirakkal H, Yu C, Griffiths GJ, Benson

RS, Bingle CD & Corfe BM (2010) Sp1 acetylation is

associated with loss of DNA binding at promoters

associated with cell cycle arrest and cell death in a

colon cell line. Mol Cancer 9, 275.

71 Suzuki T, Kimura A, Nagai R & Horikoshi M (2000)

Regulation of interaction of the acetyltransferase

region of p300 and the DNA-binding domain of Sp1

on and through DNA binding. Genes Cells 5, 29–41.72 Kang JE, Kim MH, Lee JA, Park H, Min-Nyung L,

Auh CK & Hur MW (2005) Histone deacetylase-1

represses transcription by interacting with zinc-fingers

and interfering with the DNA binding activity of Sp1.

Cell Physiol Biochem 16, 23–30.73 Wei S, Chuang HC, Tsai WC, Yang HC, Ho SR,

Paterson AJ, Kulp SK & Chen CS (2009)

Thiazolidinediones mimic glucose starvation in

facilitating Sp1 degradation through the up-regulation

of beta-transducin repeat-containing protein. Mol

Pharmacol 76, 47–57.74 Beishline K, Kelly CM, Olofsson BA, Koduri S,

Emrich J, Greenberg RA & Azizkhan-Clifford J (2012)

Sp1 facilitates DNA double-strand break repair

through a nontranscriptional mechanism. Mol Cell

Biol 32, 3790–3799.75 Han I & Kudlow JE (1997) Reduced O-glycosylation

of Sp1 is associated with increased proteasome

susceptibility. Mol Cell Biol 17, 2550–2558.76 Su K, Roos MD, Yang X, Han I, Paterson AJ &

Kudlow JE (1999) An N-terminal region of Sp1

targets its proteasome-dependent degradation in vitro.

J Biol Chem 274, 15194–15202.77 Su K, Yang X, Roos MD, Paterson AJ & Kudlow JE

(2000) Human Sug1/p45 is involved in the proteasome-

dependent degradation of Sp1. Biochem J 348 (Pt 2),

281–289.78 Wang YT, Chuang JY, Shen MR, Yang WB, Chang

WC & Hung JJ (2008) Sumoylation of specificity

protein 1 augments its degradation by changing the

localization and increasing the specificity protein 1

proteolytic process. J Mol Biol 380, 869–885.79 Emili A, Greenblatt J & Ingles CJ (1994) Species-

specific interaction of the glutamine-rich activation

domains of Sp1 with the TATA box-binding protein.

Mol Cell Biol 14, 1582–1593.80 Torigoe T, Izumi H, Yoshida Y, Ishiguchi H,

Okamoto T, Itoh H & Kohno K (2003) Low pH

enhances Sp1 DNA binding activity and interaction

with TBP. Nucleic Acids Res 31, 4523–4530.81 Kundu TK, Palhan VB, Wang Z, An W, Cole PA &

Roeder RG (2000) Activator-dependent transcription

from chromatin in vitro involving targeted histone

acetylation by p300. Mol Cell 6, 551–561.82 Chen YJ, Wang YN & Chang WC (2007) ERK2-

mediated C-terminal serine phosphorylation of p300 is

vital to the regulation of epidermal growth factor-

induced keratin 16 gene expression. J Biol Chem 282,

27215–27228.83 Hoey T, Weinzierl RO, Gill G, Chen JL, Dynlacht BD

& Tjian R (1993) Molecular cloning and functional

analysis of Drosophila TAF110 reveal properties

expected of coactivators. Cell 72, 247–260.84 Chen JL, Attardi LD, Verrijzer CP, Yokomori K &

Tjian R (1994) Assembly of recombinant TFIID

reveals differential coactivator requirements for

distinct transcriptional activators. Cell 79, 93–105.85 Gill G, Pascal E, Tseng ZH & Tjian R (1994) A

glutamine-rich hydrophobic patch in transcription

factor Sp1 contacts the dTAFII110 component of the

Drosophila TFIID complex and mediates

transcriptional activation. Proc Natl Acad Sci USA 91,

192–196.86 Tanese N, Saluja D, Vassallo MF, Chen JL & Admon A

(1996) Molecular cloning and analysis of two subunits

of the human TFIID complex: hTAFII130 and

hTAFII100. Proc Natl Acad Sci USA 93, 13611–13616.87 Saluja D, Vassallo MF & Tanese N (1998) Distinct

subdomains of human TAFII130 are required for

interactions with glutamine-rich transcriptional

activators. Mol Cell Biol 18, 5734–5743.88 Rojo-Niersbach E, Furukawa T & Tanese N (1999)

Genetic dissection of hTAF(II)130 defines a

hydrophobic surface required for interaction with

glutamine-rich activators. J Biol Chem 274, 33778–33784.

89 Dunah AW, Jeong H, Griffin A, Kim YM, Standaert

DG, Hersch SM, Mouradian MM, Young AB, Tanese

N & Krainc D (2002) Sp1 and TAFII130

transcriptional activity disrupted in early Huntington’s

disease. Science 296, 2238–2243.90 Ma Z, Chang MJ, Shah R, Adamski J, Zhao X &

Benveniste EN (2004) Brg-1 is required for maximal

transcription of the human matrix metalloproteinase-2

gene. J Biol Chem 279, 46326–46334.

245FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

91 Kadam S, McAlpine GS, Phelan ML, Kingston RE,

Jones KA & Emerson BM (2000) Functional

selectivity of recombinant mammalian SWI/SNF

subunits. Genes Dev 14, 2441–2451.92 Kadam S & Emerson BM (2003) Transcriptional

specificity of human SWI/SNF BRG1 and BRM

chromatin remodeling complexes. Mol Cell 11, 377–389.

93 Liu H, Kang H, Liu R, Chen X & Zhao K (2002)

Maximal induction of a subset of interferon target

genes requires the chromatin-remodeling activity of the

BAF complex. Mol Cell Biol 22, 6471–6479.94 Xu YZ, Heravi M, Thuraisingam T, Di Marco S,

Muanza T & Radzioch D (2010) Brg-1 mediates the

constitutive and fenretinide-induced expression of

SPARC in mammary carcinoma cells via its

interaction with transcription factor Sp1. Mol Cancer

9, 210.

95 Chiang CM & Roeder RG (1995) Cloning of an

intrinsic human TFIID subunit that interacts with

multiple transcriptional activators. Science 267, 531–536.

96 De Cesare D, Fimia GM, Brancorsini S, Parvinen M

& Sassone-Corsi P (2003) Transcriptional control in

male germ cells: general factor TFIIA participates in

CREM-dependent gene activation. Mol Endocrinol 17,

2554–2565.97 Aiello A, Pandini G, Sarfstein R, Werner H,

Manfioletti G, Vigneri R & Belfiore A (2010) HMGA1

protein is a positive regulator of the insulin-like

growth factor-I receptor gene. Eur J Cancer 46, 1919–1926.

98 Li AY, Lin HH, Kuo CY, Shih HM, Wang CC, Yen

Y & Ann DK (2011) High-mobility group A2 protein

modulates hTERT transcription to promote

tumorigenesis. Mol Cell Biol 31, 2605–2617.99 Suzuki T, Muto S, Miyamoto S, Aizawa K, Horikoshi

M & Nagai R (2003) Functional interaction of the

DNA-binding transcription factor Sp1 through its

DNA-binding domain with the histone chaperone

TAF-I. J Biol Chem 278, 28758–28764.100 Esteve PO, Chin HG & Pradhan S (2007) Molecular

mechanisms of transactivation and doxorubicin-

mediated repression of survivin gene in cancer cells. J

Biol Chem 282, 2615–2625.101 Song J, Ugai H, Kanazawa I, Sun K & Yokoyama

KK (2001) Independent repression of a GC-rich

housekeeping gene by Sp1 and MAZ involves the

same cis-elements. J Biol Chem 276, 19897–19904.102 Doetzlhofer A, Rotheneder H, Lagger G, Koranda M,

Kurtev V, Brosch G, Wintersberger E & Seiser C

(1999) Histone deacetylase 1 can repress transcription

by binding to Sp1. Mol Cell Biol 19, 5504–5511.103 Zhang Y & Dufau ML (2002) Silencing of

transcription of the human luteinizing hormone

receptor gene by histone deacetylase-mSin3A complex.

J Biol Chem 277, 33431–33438.104 Chang YC, Illenye S & Heintz NH (2001)

Cooperation of E2F-p130 and Sp1-pRb complexes in

repression of the Chinese hamster dhfr gene. Mol Cell

Biol 21, 1121–1131.105 Hou M, Wang X, Popov N, Zhang A, Zhao X, Zhou

R, Zetterberg A, Bjorkholm M, Henriksson M,

Gruber A et al. (2002) The histone deacetylase

inhibitor trichostatin A derepresses the telomerase

reverse transcriptase (hTERT) gene in human cells.

Exp Cell Res 274, 25–34.106 Ohlsson C, Kley N, Werner H & LeRoith D (1998)

p53 regulates insulin-like growth factor-I (IGF-I)

receptor expression and IGF-I-induced tyrosine

phosphorylation in an osteosarcoma cell line:

interaction between p53 and Sp1. Endocrinology 139,

1101–1107.107 Lagger G, Doetzlhofer A, Schuettengruber B,

Haidweger E, Simboeck E, Tischler J, Chiocca S,

Suske G, Rotheneder H, Wintersberger E et al. (2003)

The tumor suppressor p53 and histone deacetylase 1

are antagonistic regulators of the cyclin-dependent

kinase inhibitor p21/WAF1/CIP1 gene. Mol Cell Biol

23, 2669–2679.108 Won J, Yim J & Kim TK (2002) Sp1 and Sp3 recruit

histone deacetylase to repress transcription of human

telomerase reverse transcriptase (hTERT) promoter in

normal human somatic cells. J Biol Chem 277, 38230–38238.

109 Sun JM, Chen HY, Moniwa M, Litchfield DW, Seto

E & Davie JR (2002) The transcriptional repressor

Sp3 is associated with CK2-phosphorylated histone

deacetylase 2. J Biol Chem 277, 35783–35786.110 Lin SY, Black AR, Kostic D, Pajovic S, Hoover CN

& Azizkhan JC (1996) Cell cycle-regulated association

of E2F1 and Sp1 is related to their functional

interaction. Mol Cell Biol 16, 1668–1675.111 Gartel AL, Ye X, Goufman E, Shianov P, Hay N,

Najmabadi F & Tyner AL (2001) Myc represses the

p21(WAF1/CIP1) promoter and interacts with Sp1/

Sp3. Proc Natl Acad Sci USA 98, 4510–4515.112 Ichimura T, Watanabe S, Sakamoto Y, Aoto T, Fujita

N & Nakao M (2005) Transcriptional repression and

heterochromatin formation by MBD1 and MCAF/

AM family proteins. J Biol Chem 280, 13928–13935.113 Kardassis D, Papakosta P, Pardali K & Moustakas A

(1999) c-Jun transactivates the promoter of the human

p21(WAF1/Cip1) gene by acting as a superactivator of

the ubiquitous transcription factor Sp1. J Biol Chem

274, 29572–29581.114 Gizard F, Robillard R, Barbier O, Quatannens B,

Faucompre A, Revillion F, Peyrat JP, Staels B &

Hum DW (2005) TReP-132 controls cell proliferation

by regulating the expression of the cyclin-dependent

246 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

kinase inhibitors p21WAF1/Cip1 and p27Kip1. Mol

Cell Biol 25, 4335–4348.115 Abramovitch S, Glaser T, Ouchi T & Werner H

(2003) BRCA1-Sp1 interactions in transcriptional

regulation of the IGF-IR gene. FEBS Lett 541, 149–154.

116 Dong J, Tsai-Morris CH & Dufau ML (2006) A novel

estradiol/estrogen receptor alpha-dependent

transcriptional mechanism controls expression of the

human prolactin receptor. J Biol Chem 281, 18825–18836.

117 Tapias A, Ciudad CJ, Roninson IB & Noe V (2008)

Regulation of Sp1 by cell cycle related proteins. Cell

Cycle 7, 2856–2867.118 Vallian S, Chin KV & Chang KS (1998) The

promyelocytic leukemia protein interacts with Sp1 and

inhibits its transactivation of the epidermal growth

factor receptor promoter. Mol Cell Biol 18, 7147–7156.119 Johnson-Pais T, Degnin C & Thayer MJ (2001)

pRB induces Sp1 activity by relieving inhibition

mediated by MDM2. Proc Natl Acad Sci USA 98,

2211–2216.120 Lim K & Chang HI (2009) O-GlcNAc inhibits

interaction between Sp1 and Elf-1 transcription

factors. Biochem Biophys Res Commun 380, 569–574.121 Pal S, Claffey KP, Cohen HT & Mukhopadhyay D

(1998) Activation of Sp1-mediated vascular

permeability factor/vascular endothelial growth factor

transcription requires specific interaction with protein

kinase C zeta. J Biol Chem 273, 26277–26280.122 Cohen HT, Zhou M, Welsh AM, Zarghamee S,

Scholz H, Mukhopadhyay D, Kishida T, Zbar B,

Knebelmann B & Sukhatme VP (1999) An important

von Hippel-Lindau tumor suppressor domain mediates

Sp1-binding and self-association. Biochem Biophys Res

Commun 266, 43–50.123 Rafty LA & Khachigian LM (2002) von Hippel-

Lindau tumor suppressor protein represses platelet-

derived growth factor B-chain gene expression via the

Sp1 binding element in the proximal PDGF-B

promoter. J Cell Biochem 85, 490–495.124 Lim K & Chang HI (2009) O-GlcNAc modification of

Sp1 inhibits the functional interaction between Sp1

and Oct1. FEBS Lett 583, 512–520.125 Lim K & Chang HI (2009) O-GlcNAcylation of Sp1

interrupts Sp1 interaction with NF-Y. Biochem

Biophys Res Commun 382, 593–597.126 Di Padova M, Bruno T, De Nicola F, Iezzi S,

D’Angelo C, Gallo R, Nicosia D, Corbi N, Biroccio

A, Floridi A et al. (2003) Che-1 arrests human colon

carcinoma cell proliferation by displacing HDAC1

from the p21WAF1/CIP1 promoter. J Biol Chem 278,

36496–36504.127 Zhang X & Liu Y (2003) Suppression of HGF

receptor gene expression by oxidative stress is

mediated through the interplay between Sp1 and Egr-

1. Am J Physiol Renal Physiol 284, F1216–F1225.128 Trinh BQ, Barengo N & Naora H (2011)

Homeodomain protein DLX4 counteracts key

transcriptional control mechanisms of the TGF-beta

cytostatic program and blocks the antiproliferative

effect of TGF-beta. Oncogene 30, 2718–2729.129 Koshiji M, To KK, Hammer S, Kumamoto K, Harris

AL, Modrich P & Huang LE (2005) HIF-1alpha

induces genetic instability by transcriptionally

downregulating MutSalpha expression. Mol Cell 17,

793–803.130 Gueven N, Keating K, Fukao T, Loeffler H, Kondo

N, Rodemann HP & Lavin MF (2003) Site-directed

mutagenesis of the ATM promoter: consequences for

response to proliferation and ionizing radiation. Genes

Chromosom Cancer 38, 157–167.131 Lee JS, Galvin KM & Shi Y (1993) Evidence for

physical interaction between the zinc-finger

transcription factors YY1 and Sp1. Proc Natl Acad

Sci USA 90, 6145–6149.132 Czuwara-Ladykowska J, Shirasaki F, Jackers P,

Watson DK & Trojanowska M (2001) Fli-1 inhibits

collagen type I production in dermal fibroblasts via an

Sp1-dependent pathway. J Biol Chem 276, 20839–20848.

133 Lim K & Chang HI (2010) O-GlcNAc inhibits

interaction between Sp1 and sterol regulatory element

binding protein 2. Biochem Biophys Res Commun 393,

314–318.134 Yieh L, Sanchez HB & Osborne TF (1995) Domains

of transcription factor Sp1 required for synergistic

activation with sterol regulatory element binding

protein 1 of low density lipoprotein receptor

promoter. Proc Natl Acad Sci USA 92, 6102–6106.135 Moustakas A & Kardassis D (1998) Regulation of the

human p21/WAF1/Cip1 promoter in hepatic cells by

functional interactions between Sp1 and Smad family

members. Proc Natl Acad Sci USA 95, 6733–6738.136 Feng XH, Lin X & Derynck R (2000) Smad2, Smad3

and Smad4 cooperate with Sp1 to induce p15(Ink4B)

transcription in response to TGF-beta. EMBO J 19,

5178–5193.137 Wang CH, Chang HC & Hung WC (2006) p16

inhibits matrix metalloproteinase-2 expression via

suppression of Sp1-mediated gene transcription. J Cell

Physiol 208, 246–252.138 Wei M, Liu B, Gu Q, Su L, Yu Y & Zhu Z (2013)

Stat6 cooperates with Sp1 in controlling breast cancer

cell proliferation by modulating the expression of p21

(Cip1/WAF1) and p27 (Kip1). Cell Oncol 36, 79–93.139 Lee JA, Suh DC, Kang JE, Kim MH, Park H, Lee

MN, Kim JM, Jeon BN, Roh HE, Yu MY et al.

(2005) Transcriptional activity of Sp1 is regulated by

molecular interactions between the zinc finger DNA

247FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

binding domain and the inhibitory domain with

corepressors, and this interaction is modulated by

MEK. J Biol Chem 280, 28061–28071.140 Han JA & Park SC (2000) Transglutaminase-

dependent modulation of transcription factor Sp1

activity. Mol Cells 10, 612–618.141 Tatsukawa H, Fukaya Y, Frampton G, Martinez-

Fuentes A, Suzuki K, Kuo TF, Nagatsuma K,

Shimokado K, Okuno M, Wu J et al. (2009) Role of

transglutaminase 2 in liver injury via cross-linking and

silencing of transcription factor Sp1. Gastroenterology

136, 1783–95 e10.

142 Li SH, Cheng AL, Zhou H, Lam S, Rao M, Li H &

Li XJ (2002) Interaction of Huntington disease protein

with transcriptional activator Sp1. Mol Cell Biol 22,

1277–1287.143 Singh AK, Battu A, Mohareer K, Hasnain SE &

Ehtesham NZ (2010) Transcription of human resistin

gene involves an interaction of Sp1 with peroxisome

proliferator-activating receptor gamma

(PPARgamma). PLoS One 5, e9912.

144 Jeang KT, Chun R, Lin NH, Gatignol A, Glabe CG

& Fan H (1993) In vitro and in vivo binding of

human immunodeficiency virus type 1 Tat protein and

Sp1 transcription factor. J Virol 67, 6224–6233.145 Curtin D, Jenkins S, Farmer N, Anderson AC,

Haisenleder DJ, Rissman E, Wilson EM & Shupnik

MA (2001) Androgen suppression of GnRH-

stimulated rat LHbeta gene transcription occurs

through Sp1 sites in the distal GnRH-responsive

promoter region. Mol Endocrinol 15, 1906–1917.146 deGraffenried LA, Hopp TA, Valente AJ, Clark RA

& Fuqua SA (2004) Regulation of the estrogen

receptor alpha minimal promoter by Sp1, USF-1 and

ERalpha. Breast Cancer Res Treat 85, 111–120.147 Han WD, Si YL, Zhao YL, Li Q, Wu ZQ, Hao HJ &

Song HJ (2008) GC-rich promoter elements maximally

confers estrogen-induced transactivation of LRP16

gene through ERalpha/Sp1 interaction in MCF-7 cells.

J Steroid Biochem Mol Biol 109, 47–56.148 Kang JH, Tsai-Morris CH & Dufau ML (2011)

Complex formation and interactions between

transcription factors essential for human prolactin

receptor gene transcription. Mol Cell Biol 31, 3208–3222.

149 Khan S, Barhoumi R, Burghardt R, Liu S, Kim K &

Safe S (2006) Molecular mechanism of inhibitory aryl

hydrocarbon receptor-estrogen receptor/Sp1 cross talk

in breast cancer cells. Mol Endocrinol 20, 2199–2214.150 Kim K, Barhoumi R, Burghardt R & Safe S (2005)

Analysis of estrogen receptor alpha-Sp1 interactions in

breast cancer cells by fluorescence resonance energy

transfer. Mol Endocrinol 19, 843–854.151 Li D, Mitchell D, Luo J, Yi Z, Cho SG, Guo J, Li X,

Ning G, Wu X & Liu M (2007) Estrogen regulates

KiSS1 gene expression through estrogen receptor

alpha and SP protein complexes. Endocrinology 148,

4821–4828.152 Panno ML, Mauro L, Marsico S, Bellizzi D, Rizza P,

Morelli C, Salerno M, Giordano F & Ando S (2006)

Evidence that the mouse insulin receptor substrate-1

belongs to the gene family on which the promoter is

activated by estrogen receptor alpha through its

interaction with Sp1. J Mol Endocrinol 36, 91–105.153 Sisci D, Middea E, Morelli C, Lanzino M, Aquila S,

Rizza P, Catalano S, Casaburi I, Maggiolini M &

Ando S (2010) 17beta-estradiol enhances alpha(5)

integrin subunit gene expression through ERalpha-Sp1

interaction and reduces cell motility and invasion of

ERalpha-positive breast cancer cells. Breast Cancer

Res Treat 124, 63–77.154 Suzuki A, Sanda N, Miyawaki Y, Fujimori Y,

Yamada T, Takagi A, Murate T, Saito H & Kojima T

(2010) Down-regulation of PROS1 gene expression by

17beta-estradiol via estrogen receptor alpha

(ERalpha)-Sp1 interaction recruiting receptor-

interacting protein 140 and the corepressor-HDAC3

complex. J Biol Chem 285, 13444–13453.155 Tu CC, Velmurugan BK, Day CH, Kuo WW, Yeh

SP, Chen RJ, Liao CR, Chen HY, Tsai FJ, Wu WJ

et al. (2013) Estrogen receptor-alpha over-expression

mediated apoptosis in Hep3B cells by binding with

Sp1 proteins. J Mol Endocrinol 51, 203–212.156 Chiefari E, Brunetti A, Arturi F, Bidart JM, Russo D,

Schlumberger M & Filetti S (2002) Increased

expression of AP2 and Sp1 transcription factors in

human thyroid tumors: a role in NIS expression

regulation? BMC Cancer 2, 35.

157 Wang L, Wei D, Huang S, Peng Z, Le X, Wu TT,

Yao J, Ajani J & Xie K (2003) Transcription factor

Sp1 expression is a significant predictor of survival in

human gastric cancer. Clin Cancer Res 9, 6371–6380.158 Jiang NY, Woda BA, Banner BF, Whalen GF,

Dresser KA & Lu D (2008) Sp1, a new biomarker that

identifies a subset of aggressive pancreatic ductal

adenocarcinoma. Cancer Epidemiol Biomarkers Prev

17, 1648–1652.159 Guan H, Cai J, Zhang N, Wu J, Yuan J, Li J & Li M

(2012) Sp1 is upregulated in human glioma, promotes

MMP-2-mediated cell invasion and predicts poor

clinical outcome. Int J Cancer 130, 593–601.160 Wang XB, Peng WQ, Yi ZJ, Zhu SL & Gan QH

(2007) Expression and prognostic value of

transcriptional factor sp1 in breast cancer. Ai Zheng

26, 996–1000.161 Seznec J, Silkenstedt B & Naumann U (2011)

Therapeutic effects of the Sp1 inhibitor mithramycin

A in glioblastoma. J Neurooncol 101, 365–377.162 Sun Y, Giacalone NJ & Lu B (2011) Terameprocol

(tetra-O-methyl nordihydroguaiaretic acid), an

248 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

inhibitor of Sp1-mediated survivin transcription,

induces radiosensitization in non-small cell lung

carcinoma. J Thorac Oncol 6, 8–14.163 Kim JY, Kim EH, Park SS, Lim JH, Kwon TK &

Choi KS (2008) Quercetin sensitizes human hepatoma

cells to TRAIL-induced apoptosis via Sp1-mediated

DR5 up-regulation and proteasome-mediated c-FLIPS

down-regulation. J Cell Biochem 105, 1386–1398.164 Chae JI, Cho JH, Lee KA, Choi NJ, Seo KS, Kim

SB, Lee SH & Shim JH (2012) Role of transcription

factor Sp1 in the quercetin-mediated inhibitory effect

on human malignant pleural mesothelioma. Int J Mol

Med 30, 835–841.165 Yuan H, Gong A & Young CY (2005) Involvement of

transcription factor Sp1 in quercetin-mediated

inhibitory effect on the androgen receptor in human

prostate cancer cells. Carcinogenesis 26, 793–801.166 Jia Z, Zhang J, Wei D, Wang L, Yuan P, Le X, Li Q,

Yao J & Xie K (2007) Molecular basis of the

synergistic antiangiogenic activity of bevacizumab and

mithramycin A. Cancer Res 67, 4878–4885.167 Yao JC, Wang L, Wei D, Gong W, Hassan M, Wu

TT, Mansfield P, Ajani J & Xie K (2004) Association

between expression of transcription factor Sp1 and

increased vascular endothelial growth factor

expression, advanced stage, and poor survival in

patients with resected gastric cancer. Clin Cancer Res

10, 4109–4117.168 Lee KE, Shin JA, Hong IS, Cho NP & Cho SD (2013)

Effect of methanol extracts of Cnidium officinale

Makino and Capsella bursa-pastoris on the apoptosis

of HSC-2 human oral cancer cells. Exp Ther Med 5,

789–792.169 Hsu TI, Wang MC, Chen SY, Yeh YM, Su WC,

Chang WC & Hung JJ (2012) Sp1 expression regulates

lung tumor progression. Oncogene 31, 3973–3988.170 Hanahan D & Weinberg RA (2011) Hallmarks of

cancer: the next generation. Cell 144, 646–674.171 Hanahan D & Weinberg RA (2000) The hallmarks of

cancer. Cell 100, 57–70.172 Endoh T, Tsuji N, Asanuma K, Yagihashi A &

Watanabe N (2005) Survivin enhances telomerase

activity via up-regulation of specificity protein 1- and

c-Myc-mediated human telomerase reverse

transcriptase gene transcription. Exp Cell Res 305,

300–311.173 Furuya M, Tsuji N, Kobayashi D & Watanabe N

(2009) Interaction between survivin and aurora-B

kinase plays an important role in survivin-mediated

up-regulation of human telomerase reverse

transcriptase expression. Int J Oncol 34, 1061–1068.174 Kyo S, Takakura M, Taira T, Kanaya T, Itoh H,

Yutsudo M, Ariga H & Inoue M (2000) Sp1

cooperates with c-Myc to activate transcription of the

human telomerase reverse transcriptase gene (hTERT).

Nucleic Acids Res 28, 669–677.175 Liu C, Fang X, Ge Z, Jalink M, Kyo S, Bjorkholm

M, Gruber A, Sjoberg J & Xu D (2007) The

telomerase reverse transcriptase (hTERT) gene is a

direct target of the histone methyltransferase SMYD3.

Cancer Res 67, 2626–2631.176 Marconett CN, Sundar SN, Tseng M, Tin AS, Tran

KQ, Mahuron KM, Bjeldanes LF & Firestone GL

(2011) Indole-3-carbinol downregulation of telomerase

gene expression requires the inhibition of estrogen

receptor-alpha and Sp1 transcription factor

interactions within the hTERT promoter and mediates

the G1 cell cycle arrest of human breast cancer cells.

Carcinogenesis 32, 1315–1323.177 Oh ST, Kyo S & Laimins LA (2001) Telomerase

activation by human papillomavirus type 16 E6

protein: induction of human telomerase reverse

transcriptase expression through Myc and GC-rich

Sp1 binding sites. J Virol 75, 5559–5566.178 Palumbo SL, Ebbinghaus SW & Hurley LH (2009)

Formation of a unique end-to-end stacked pair of G-

quadruplexes in the hTERT core promoter with

implications for inhibition of telomerase by G-

quadruplex-interactive ligands. J Am Chem Soc 131,

10878–10891.179 Takakura M, Kyo S, Kanaya T, Hirano H, Takeda J,

Yutsudo M & Inoue M (1999) Cloning of human

telomerase catalytic subunit (hTERT) gene promoter

and identification of proximal core promoter sequences

essential for transcriptional activation in immortalized

and cancer cells. Cancer Res 59, 551–557.180 Wick M, Zubov D & Hagen G (1999) Genomic

organization and promoter characterization of the

gene encoding the human telomerase reverse

transcriptase (hTERT). Gene 232, 97–106.181 Zhao JQ, Glasspool RM, Hoare SF, Bilsland A,

Szatmari I & Keith WN (2000) Activation of

telomerase rna gene promoter activity by NF-Y, Sp1,

and the retinoblastoma protein and repression by Sp3.

Neoplasia 2, 531–539.182 Bond GL, Hu W, Bond EE, Robins H, Lutzker SG,

Arva NC, Bargonetti J, Bartel F, Taubert H, Wuerl P

et al. (2004) A single nucleotide polymorphism in the

MDM2 promoter attenuates the p53 tumor suppressor

pathway and accelerates tumor formation in humans.

Cell 119, 591–602.183 Knappskog S & Lonning PE (2011) Effects of the

MDM2 promoter SNP285 and SNP309 on Sp1

transcription factor binding and cancer risk.

Transcription 2, 207–210.184 Knappskog S & Lonning PE (2011) MDM2 promoter

SNP285 and SNP309; phylogeny and impact on

cancer risk. Oncotarget 2, 251–258.

249FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

185 Knappskog S, Bjornslett M, Myklebust LM, Huijts PE,

Vreeswijk MP, Edvardsen H, Guo Y, Zhang X, Yang

M, Ylisaukko-Oja SK et al. (2011) The MDM2

promoter SNP285C/309G haplotype diminishes Sp1

transcription factor binding and reduces risk for breast

and ovarian cancer in Caucasians. Cancer Cell 19, 273–282.

186 Schutte M, Elstrodt F, Bralten LB, Nagel JH, Duijm

E, Hollestelle A, Vuerhard MJ, Wasielewski M,

Peeters JK, van der Spek P et al. (2008) Exon

expression arrays as a tool to identify new cancer

genes. PLoS One 3, e3007.

187 Yarden RI, Friedman E, Metsuyanim S, Olender T,

Ben-Asher E & Papa MZ (2008) MDM2 SNP309

accelerates breast and ovarian carcinogenesis in

BRCA1 and BRCA2 carriers of Jewish-Ashkenazi

descent. Breast Cancer Res Treat 111, 497–504.188 Wang X, Pan L, Feng Y, Wang Y, Han Q, Han L,

Han S, Guo J, Huang B & Lu J (2008) P300 plays a

role in p16(INK4a) expression and cell cycle arrest.

Oncogene 27, 1894–1904.189 Wang X, Feng Y, Pan L, Wang Y, Xu X, Lu J &

Huang B (2007) The proximal GC-rich region of p16

(INK4a) gene promoter plays a role in its

transcriptional regulation. Mol Cell Biochem 301, 259–266.

190 Wu J, Xue L, Weng M, Sun Y, Zhang Z, Wang W &

Tong T (2007) Sp1 is essential for p16 expression in

human diploid fibroblasts during senescence. PLoS

One 2, e164.

191 Xu W, Zhu Q, Wu Z, Guo H, Wu F, Mashausi DS,

Zheng C & Li D (2012) A novel evolutionarily

conserved element is a general transcriptional repressor

of p21WAF(1)/CIP(1). Cancer Res 72, 6236–6246.192 Biggs JR, Kudlow JE & Kraft AS (1996) The role of

the transcription factor Sp1 in regulating the

expression of the WAF1/CIP1 gene in U937 leukemic

cells. J Biol Chem 271, 901–906.193 Koutsodontis G, Moustakas A & Kardassis D (2002)

The role of Sp1 family members, the proximal GC-

rich motifs, and the upstream enhancer region in the

regulation of the human cell cycle inhibitor p21WAF-

1/Cip1 gene promoter. Biochemistry 41, 12771–12784.194 Kavurma MM & Khachigian LM (2004) Vascular

smooth muscle cell-specific regulation of cyclin-

dependent kinase inhibitor p21(WAF1/Cip1)

transcription by Sp1 is mediated via distinct cis-acting

positive and negative regulatory elements in the

proximal p21(WAF1/Cip1) promoter. J Cell Biochem

93, 904–916.195 Werner H, Stannard B, Bach MA, LeRoith D &

Roberts CT Jr (1990) Cloning and characterization of

the proximal promoter region of the rat insulin-like

growth factor I (IGF-I) receptor gene. Biochem

Biophys Res Commun 169, 1021–1027.

196 Jensen DE, Rich CB, Terpstra AJ, Farmer SR &

Foster JA (1995) Transcriptional regulation of the

elastin gene by insulin-like growth factor-I involves

disruption of Sp1 binding. Evidence for the role of Rb

in mediating Sp1 binding in aortic smooth muscle

cells. J Biol Chem 270, 6555–6563.197 Kaytor EN, Zhu JL, Pao CI & Phillips LS (2001)

Physiological concentrations of insulin promote

binding of nuclear proteins to the insulin-like growth

factor I gene. Endocrinology 142, 1041–1049.198 Maor S, Mayer D, Yarden RI, Lee AV, Sarfstein R,

Werner H & Papa MZ (2006) Estrogen receptor

regulates insulin-like growth factor-I receptor gene

expression in breast tumor cells: involvement of

transcription factor Sp1. J Endocrinol 191, 605–612.199 Maor S, Yosepovich A, Papa MZ, Yarden RI, Mayer

D, Friedman E & Werner H (2007) Elevated insulin-

like growth factor-I receptor (IGF-IR) levels in

primary breast tumors associated with BRCA1

mutations. Cancer Lett 257, 236–243.200 Maor SB, Abramovitch S, Erdos MR, Brody LC &

Werner H (2000) BRCA1 suppresses insulin-like

growth factor-I receptor promoter activity: potential

interaction between BRCA1 and Sp1. Mol Genet

Metab 69, 130–136.201 Shahrabani-Gargir L, Pandita TK & Werner H (2004)

Ataxia-telangiectasia mutated gene controls insulin-

like growth factor I receptor gene expression in a

deoxyribonucleic acid damage response pathway via

mechanisms involving zinc-finger transcription factors

Sp1 and WT1. Endocrinology 145, 5679–5687.202 Haley J, Whittle N, Bennet P, Kinchington D, Ullrich

A & Waterfield M (1987) The human EGF receptor

gene: structure of the 110 kb locus and identification

of sequences regulating its transcription. Oncogene Res

1, 375–396.203 Johnson AC, Ishii S, Jinno Y, Pastan I & Merlino GT

(1988) Epidermal growth factor receptor gene

promoter. Deletion analysis and identification of

nuclear protein binding sites. J Biol Chem 263, 5693–5699.

204 Kageyama R, Merlino GT & Pastan I (1988)

Epidermal growth factor (EGF) receptor gene

transcription. Requirement for Sp1 and an EGF

receptor-specific factor. J Biol Chem 263, 6329–6336.205 Kitadai Y, Yasui W, Yokozaki H, Kuniyasu H,

Haruma K, Kajiyama G & Tahara E (1992) The level

of a transcription factor Sp1 is correlated with the

expression of EGF receptor in human gastric

carcinomas. Biochem Biophys Res Commun 189, 1342–1348.

206 Salvatori L, Pallante P, Ravenna L, Chinzari P, Frati

L, Russo MA & Petrangeli E (2003) Oestrogens and

selective oestrogen receptor (ER) modulators regulate

EGF receptor gene expression through human ER

250 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

alpha and beta subtypes via an Sp1 site. Oncogene 22,

4875–4881.207 Carpentier C, Laigle-Donadey F, Marie Y, Auger N,

Benouaich-Amiel A, Lejeune J, Kaloshi G, Delattre

JY, Thillet J & Sanson M (2006) Polymorphism in

Sp1 recognition site of the EGF receptor gene

promoter and risk of glioblastoma. Neurology 67, 872–874.

208 Pascall JC & Brown KD (1997) Identification of a

minimal promoter element of the mouse epidermal

growth factor gene. Biochem J 324 (Pt 3), 869–875.209 Tiesman J & Rizzino A (1990) Nucleotide sequence of

the 50-flanking region of the mouse k-FGF oncogene

exhibits an alternating purine:pyrimidine motif with

the potential to form Z-DNA. Gene 96, 311–312.210 Luster TA, Johnson LR, Nowling TK, Lamb KA,

Philipsen S & Rizzino A (2000) Effects of three Sp1

motifs on the transcription of the FGF-4 gene. Mol

Reprod Dev 57, 4–15.211 Payson RA, Chotani MA & Chiu IM (1998)

Regulation of a promoter of the fibroblast growth

factor 1 gene in prostate and breast cancer cells. J

Steroid Biochem Mol Biol 66, 93–103.212 Jimenez SK, Sheikh F, Jin Y, Detillieux KA, Dhaliwal

J, Kardami E & Cattini PA (2004) Transcriptional

regulation of FGF-2 gene expression in cardiac

myocytes. Cardiovasc Res 62, 548–557.213 Zhu JL, Kaytor EN, Pao CI, Meng XP & Phillips LS

(2000) Involvement of Sp1 in the transcriptional

regulation of the rat insulin-like growth factor-1 gene.

Mol Cell Endocrinol 164, 205–218.214 Cadoret A, Baron-Delage S, Bertrand F, Kornprost

M, Groyer A, Gespach C, Capeau J & Cherqui G

(1998) Oncogene-induced up-regulation of Caco-2 cell

proliferation involves IGF-II gene activation through

a protein kinase C-mediated pathway. Oncogene 17,

877–887.215 Hyun SW, Kim SJ, Park K, Rho HM & Lee YI

(1993) Characterization of the P4 promoter region of

the human insulin-like growth factor II gene. FEBS

Lett 332, 153–158.216 Jiang Y, Wang L, Gong W, Wei D, Le X, Yao J,

Ajani J, Abbruzzese JL, Huang S & Xie K (2004) A

high expression level of insulin-like growth factor I

receptor is associated with increased expression of

transcription factor Sp1 and regional lymph node

metastasis of human gastric cancer. Clin Exp

Metastasis 21, 755–764.217 Lee YI, Lee S, Lee Y, Bong YS, Hyun SW, Yoo YD,

Kim SJ, Kim YW & Poo HR (1998) The human

hepatitis B virus transactivator X gene product

regulates Sp1 mediated transcription of an insulin-like

growth factor II promoter 4. Oncogene 16, 2367–2380.218 Li T, Chen YH, Liu TJ, Jia J, Hampson S, Shan YX,

Kibler D & Wang PH (2003) Using DNA microarray

to identify Sp1 as a transcriptional regulatory element

of insulin-like growth factor 1 in cardiac muscle cells.

Circ Res 93, 1202–1209.219 Asanuma K, Tsuji N, Endoh T, Yagihashi A &

Watanabe N (2004) Survivin enhances Fas ligand

expression via up-regulation of specificity protein 1-

mediated gene transcription in colon cancer cells. J

Immunol 172, 3922–3929.220 Kim YH, Lee DH, Jeong JH, Guo ZS & Lee YJ

(2008) Quercetin augments TRAIL-induced apoptotic

death: involvement of the ERK signal transduction

pathway. Biochem Pharmacol 75, 1946–1958.221 Lee KA, Lee YJ, Ban JO, Lee YJ, Lee SH, Cho MK,

Nam HS, Hong JT & Shim JH (2012) The flavonoid

resveratrol suppresses growth of human malignant

pleural mesothelioma cells through direct inhibition of

specificity protein 1. Int J Mol Med 30, 21–27.222 Lee SO, Abdelrahim M, Yoon K, Chintharlapalli S,

Papineni S, Kim K, Wang H & Safe S (2010)

Inactivation of the orphan nuclear receptor TR3/

Nur77 inhibits pancreatic cancer cell and tumor

growth. Cancer Res 70, 6824–6836.223 Li F & Altieri DC (1999) Transcriptional analysis of

human survivin gene expression. Biochem J 344 (Pt 2),

305–311.224 Li F & Altieri DC (1999) The cancer antiapoptosis

mouse survivin gene: characterization of locus and

transcriptional requirements of basal and cell cycle-

dependent expression. Cancer Res 59, 3143–3151.225 Kim YH, Park JW, Lee JY & Kwon TK (2004)

Sodium butyrate sensitizes TRAIL-mediated apoptosis

by induction of transcription from the DR5 gene

promoter through Sp1 sites in colon cancer cells.

Carcinogenesis 25, 1813–1820.226 Yoshida T, Maeda A, Tani N & Sakai T (2001)

Promoter structure and transcription initiation sites of

the human death receptor 5/TRAIL-R2 gene. FEBS

Lett 507, 381–385.227 Duan H, Heckman CA & Boxer LM (2005) Histone

deacetylase inhibitors down-regulate bcl-2 expression

and induce apoptosis in t(14;18) lymphomas. Mol Cell

Biol 25, 1608–1619.228 Azahri NS, Di Bartolo BA, Khachigian LM &

Kavurma MM (2012) Sp1, acetylated histone-3 and

p300 regulate TRAIL transcription: mechanisms of

PDGF-BB-mediated VSMC proliferation and

migration. J Cell Biochem 113, 2597–2606.229 Lee KA, Chae JI & Shim JH (2012) Natural

diterpenes from coffee, cafestol and kahweol induce

apoptosis through regulation of specificity protein 1

expression in human malignant pleural mesothelioma.

J Biomed Sci 19, 60.

230 Lee TJ, Jung EM, Lee JT, Kim S, Park JW, Choi KS

& Kwon TK (2006) Mithramycin A sensitizes cancer

cells to TRAIL-mediated apoptosis by down-

251FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

regulation of XIAP gene promoter through Sp1 sites.

Mol Cancer Ther 5, 2737–2746.231 Kavurma MM, Bobryshev Y & Khachigian LM

(2002) Ets-1 positively regulates Fas ligand

transcription via cooperative interactions with Sp1. J

Biol Chem 277, 36244–36252.232 Kavurma MM, Santiago FS, Bonfoco E &

Khachigian LM (2001) Sp1 phosphorylation regulates

apoptosis via extracellular FasL-Fas engagement. J

Biol Chem 276, 4964–4971.233 Savickiene J, Treigyte G, Magnusson KE &

Navakauskiene R (2005) p21 (Waf1/Cip1) and FasL

gene activation via Sp1 and NFkappaB is required for

leukemia cell survival but not for cell death induced

by diverse stimuli. Int J Biochem Cell Biol 37, 784–796.

234 Abdelrahim M & Safe S (2005) Cyclooxygenase-2

inhibitors decrease vascular endothelial growth factor

expression in colon cancer cells by enhanced

degradation of Sp1 and Sp4 proteins. Mol Pharmacol

68, 317–329.235 Abdelrahim M, Smith R III, Burghardt R & Safe S

(2004) Role of Sp proteins in regulation of vascular

endothelial growth factor expression and proliferation

of pancreatic cancer cells. Cancer Res 64, 6740–6749.236 Eisermann K, Broderick CJ, Bazarov A, Moazam

MM & Fraizer GC (2013) Androgen up-regulates

vascular endothelial growth factor expression in

prostate cancer cells via an Sp1 binding site. Mol

Cancer 12, 7.

237 Finkenzeller G, Sparacio A, Technau A, Marme D &

Siemeister G (1997) Sp1 recognition sites in the

proximal promoter of the human vascular endothelial

growth factor gene are essential for platelet-derived

growth factor-induced gene expression. Oncogene 15,

669–676.238 Kazi AA, Jones JM & Koos RD (2005) Chromatin

immunoprecipitation analysis of gene expression in the

rat uterus in vivo: estrogen-induced recruitment of

both estrogen receptor alpha and hypoxia-inducible

factor 1 to the vascular endothelial growth factor

promoter. Mol Endocrinol 19, 2006–2019.239 Koos RD, Kazi AA, Roberson MS & Jones JM

(2005) New insight into the transcriptional regulation

of vascular endothelial growth factor expression in the

endometrium by estrogen and relaxin. Ann N Y Acad

Sci 1041, 233–247.240 Milanini J, Vinals F, Pouyssegur J & Pages G (1998)

p42/p44 MAP kinase module plays a key role in the

transcriptional regulation of the vascular endothelial

growth factor gene in fibroblasts. J Biol Chem 273,

18165–18172.241 Motoyama K, Fukumoto S, Koyama H, Emoto M,

Shimano H, Maemura K & Nishizawa Y (2006)

SREBP inhibits VEGF expression in human smooth

muscle cells. Biochem Biophys Res Commun 342, 354–360.

242 Mukhopadhyay D, Knebelmann B, Cohen HT,

Ananth S & Sukhatme VP (1997) The von Hippel-

Lindau tumor suppressor gene product interacts with

Sp1 to repress vascular endothelial growth factor

promoter activity. Mol Cell Biol 17, 5629–5639.

243 Pages G & Pouyssegur J (2005) Transcriptional

regulation of the Vascular Endothelial Growth Factor

gene–a concert of activating factors. Cardiovasc Res

65, 564–573.244 Pore N, Liu S, Shu HK, Li B, Haas-Kogan D, Stokoe

D, Milanini-Mongiat J, Pages G, O’Rourke DM,

Bernhard E et al. (2004) Sp1 is involved in Akt-

mediated induction of VEGF expression through an

HIF–1-independent mechanism. Mol Biol Cell 15,

4841–4853.245 Reisinger K, Kaufmann R & Gille J (2003) Increased

Sp1 phosphorylation as a mechanism of hepatocyte

growth factor (HGF/SF)-induced vascular endothelial

growth factor (VEGF/VPF) transcription. J Cell Sci

116, 225–238.246 Santra M, Santra S, Zhang J & Chopp M (2008)

Ectopic decorin expression up-regulates VEGF

expression in mouse cerebral endothelial cells via

activation of the transcription factors Sp1, HIF1alpha,

and Stat3. J Neurochem 105, 324–337.247 Schafer G, Cramer T, Suske G, Kemmner W,

Wiedenmann B & Hocker M (2003) Oxidative stress

regulates vascular endothelial growth factor-A gene

transcription through Sp1- and Sp3-dependent

activation of two proximal GC-rich promoter

elements. J Biol Chem 278, 8190–8198.248 Shi Q, Le X, Abbruzzese JL, Peng Z, Qian CN, Tang

H, Xiong Q, Wang B, Li XC & Xie K (2001)

Constitutive Sp1 activity is essential for differential

constitutive expression of vascular endothelial growth

factor in human pancreatic adenocarcinoma. Cancer

Res 61, 4143–4154.249 Stoner M, Wormke M, Saville B, Samudio I, Qin C,

Abdelrahim M & Safe S (2004) Estrogen regulation of

vascular endothelial growth factor gene expression in

ZR-75 breast cancer cells through interaction of

estrogen receptor alpha and SP proteins. Oncogene 23,

1052–1063.250 Tischer E, Mitchell R, Hartman T, Silva M,

Gospodarowicz D, Fiddes JC & Abraham JA (1991)

The human gene for vascular endothelial growth factor.

Multiple protein forms are encoded through alternative

exon splicing. J Biol Chem 266, 11947–11954.251 Wei D, Wang L, He Y, Xiong HQ, Abbruzzese JL &

Xie K (2004) Celecoxib inhibits vascular endothelial

growth factor expression in and reduces angiogenesis

and metastasis of human pancreatic cancer via

252 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

suppression of Sp1 transcription factor activity. Cancer

Res 64, 2030–2038.252 Yu DC, Waby JS, Chirakkal H, Staton CA & Corfe

BM (2010) Butyrate suppresses expression of

neuropilin I in colorectal cell lines through inhibition

of Sp1 transactivation. Mol Cancer 9, 276.

253 Okamoto M, Ono M, Uchiumi T, Ueno H, Kohno K,

Sugimachi K & Kuwano M (2002) Up-regulation of

thrombospondin-1 gene by epidermal growth factor

and transforming growth factor beta in human cancer

cells–transcriptional activation and messenger RNA

stabilization. Biochim Biophys Acta 1574, 24–34.254 Liu MY, Eyries M, Zhang C, Santiago FS &

Khachigian LM (2006) Inducible platelet-derived

growth factor D-chain expression by angiotensin II

and hydrogen peroxide involves transcriptional

regulation by Ets-1 and Sp1. Blood 107, 2322–2329.255 Rafty LA & Khachigian LM (2001) Sp1

phosphorylation regulates inducible expression of

platelet-derived growth factor B-chain gene via atypical

protein kinase C-zeta. Nucleic Acids Res 29, 1027–1033.256 Rafty LA, Santiago FS & Khachigian LM (2002)

NF1/X represses PDGF A-chain transcription by

interacting with Sp1 and antagonizing Sp1 occupancy

of the promoter. EMBO J 21, 334–343.257 Santiago FS & Khachigian LM (2004) Ets-1 stimulates

platelet-derived growth factor A-chain gene

transcription and vascular smooth muscle cell growth

via cooperative interactions with Sp1. Circ Res 95,

479–487.258 Belaguli NS, Aftab M, Rigi M, Zhang M, Albo D &

Berger DH (2010) GATA6 promotes colon cancer cell

invasion by regulating urokinase plasminogen

activator gene expression. Neoplasia 12, 856–865.259 Benasciutti E, Pages G, Kenzior O, Folk W, Blasi F &

Crippa MP (2004) MAPK and JNK transduction

pathways can phosphorylate Sp1 to activate the uPA

minimal promoter element and endogenous gene

transcription. Blood 104, 256–262.260 Zannetti A, Del Vecchio S, Carriero MV, Fonti R,

Franco P, Botti G, D’Aiuto G, Stoppelli MP &

Salvatore M (2000) Coordinate up-regulation of Sp1

DNA-binding activity and urokinase receptor

expression in breast carcinoma. Cancer Res 60,

1546–1551.261 Hockings JK, Degner SC, Morgan SS, Kemp MQ &

Romagnolo DF (2008) Involvement of a specificity

proteins-binding element in regulation of basal and

estrogen-induced transcription activity of the BRCA1

gene. Breast Cancer Res 10, R29.

262 Maor S, Papa MZ, Yarden RI, Friedman E, Lerenthal

Y, Lee SW, Mayer D & Werner H (2007) Insulin-like

growth factor-I controls BRCA1 gene expression

through activation of transcription factor Sp1. Horm

Metab Res 39, 179–185.

263 Keating KE, Gueven N, Watters D, Rodemann HP &

Lavin MF (2001) Transcriptional downregulation of

ATM by EGF is defective in ataxia-telangiectasia cells

expressing mutant protein. Oncogene 20, 4281–4290.264 Truman JP, Gueven N, Lavin M, Leibel S, Kolesnick

R, Fuks Z & Haimovitz-Friedman A (2005) Down-

regulation of ATM protein sensitizes human prostate

cancer cells to radiation-induced apoptosis. J Biol

Chem 280, 23262–23272.265 Bu Y, Suenaga Y, Ono S, Koda T, Song F,

Nakagawara A & Ozaki T (2008) Sp1-mediated

transcriptional regulation of NFBD1/MDC1 plays a

critical role in DNA damage response pathway. Genes

Cells 13, 53–66.266 Dalvai M, Mondesert O, Bourdon JC, Ducommun B

& Dozier C (2011) Cdc25B is negatively regulated by

p53 through Sp1 and NF-Y transcription factors.

Oncogene 30, 2282–2288.267 Dalvai M, Mondesert O, Bugler B, Manenti S,

Ducommun B & Dozier C (2012) Doxorubicin

promotes transcriptional upregulation of Cdc25B in

cancer cells by releasing Sp1 from the promoter.

Oncogene 32, 5123–5128.268 Sengupta S, Shimamoto A, Koshiji M, Pedeux R,

Rusin M, Spillare EA, Shen JC, Huang LE, Lindor

NM, Furuichi Y et al. (2005) Tumor suppressor p53

represses transcription of RECQ4 helicase. Oncogene

24, 1738–1748.269 Oei SL, Herzog H, Hirsch-Kauffmann M, Schneider R,

Auer B & Schweiger M (1994) Transcriptional

regulation and autoregulation of the human gene for

ADP-ribosyltransferase. Mol Cell Biochem 138, 99–104.270 Laniel MA, Poirier GG & Guerin SL (2004) A

conserved initiator element on the mammalian poly

(ADP-ribose) polymerase-1 promoters, in combination

with flanking core elements, is necessary to obtain

high transcriptional activity. Biochim Biophys Acta

1679, 37–46.271 Hao B, Miao X, Li Y, Zhang X, Sun T, Liang G,

Zhao Y, Zhou Y, Wang H, Chen X et al. (2006) A

novel T-77C polymorphism in DNA repair gene

XRCC1 contributes to diminished promoter activity

and increased risk of non-small cell lung cancer.

Oncogene 25, 3613–3620.272 Zhou ZQ & Walter CA (1998) Cloning and

characterization of the promoter of baboon XRCC1, a

gene involved in DNA strand-break repair. Somat Cell

Mol Genet 24, 23–39.273 Iso T, Futami K, Iwamoto T & Furuichi Y (2007)

Modulation of the expression of bloom helicase by

estrogenic agents. Biol Pharm Bull 30, 266–271.274 Wang H, Wang S, Shen L, Chen Y, Zhang X, Zhou J,

Wang Z, Hu C, Yue W & Wang H (2010) Chk2

down-regulation by promoter hypermethylation in

human bulk gliomas. Life Sci 86, 185–191.

253FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

275 Zhang S, Lu J, Zhao X, Wu W, Wang H, Lu J, Wu

Q, Chen X, Fan W, Chen H et al. (2010) A variant in

the CHEK2 promoter at a methylation site relieves

transcriptional repression and confers reduced risk of

lung cancer. Carcinogenesis 31, 1251–1258.276 Xu XS, Wang L, Abrams J & Wang G (2011) Histone

deacetylases (HDACs) in XPC gene silencing and

bladder cancer. J Hematol Oncol 4, 17.

277 Jaitovich-Groisman I, Benlimame N, Slagle BL, Perez

MH, Alpert L, Song DJ, Fotouhi-Ardakani N,

Galipeau J & Alaoui-Jamali MA (2001)

Transcriptional regulation of the TFIIH transcription

repair components XPB and XPD by the hepatitis B

virus x protein in liver cells and transgenic liver tissue.

J Biol Chem 276, 14124–14132.278 Ma L, Weeda G, Jochemsen AG, Bootsma D,

Hoeijmakers JH & van der Eb AJ (1992) Molecular

and functional analysis of the XPBC/ERCC-3

promoter: transcription activity is dependent on the

integrity of an Sp1-binding site. Nucleic Acids Res 20,

217–224.279 Nichols AF, Itoh T, Zolezzi F, Hutsell S & Linn S

(2003) Basal transcriptional regulation of human

damage-specific DNA-binding protein genes DDB1

and DDB2 by Sp1, E2F, N-myc and NF1 elements.

Nucleic Acids Res 31, 562–569.280 Hosoi Y, Watanabe T, Nakagawa K, Matsumoto Y,

Enomoto A, Morita A, Nagawa H & Suzuki N (2004)

Up-regulation of DNA-dependent protein kinase

activity and Sp1 in colorectal cancer. Int J Oncol 25,

461–468.281 Wang S, Wang M, Yin S, Fu G, Li C, Chen R, Li A,

Zhou J, Zhang Z & Liu Q (2008) A novel variable

number of tandem repeats (VNTR) polymorphism

containing Sp1 binding elements in the promoter of

XRCC5 is a risk factor for human bladder cancer.

Mutat Res 638, 26–36.282 Lin Z, Zhang X, Tuo J, Guo Y, Green B, Chan

CC, Tan W, Huang Y, Ling W, Kadlubar FF et al.

(2008) A variant of the Cockayne syndrome B gene

ERCC6 confers risk of lung cancer. Hum Mutat 29,

113–122.283 Yamabe Y, Shimamoto A, Goto M, Yokota J,

Sugawara M & Furuichi Y (1998) Sp1-mediated

transcription of the Werner helicase gene is modulated

by Rb and p53. Mol Cell Biol 18, 6191–6200.284 Hung WC, Tseng WL, Shiea J & Chang HC (2010)

Skp2 overexpression increases the expression of MMP-

2 and MMP-9 and invasion of lung cancer cells.

Cancer Lett 288, 156–161.285 Sroka IC, Nagle RB & Bowden GT (2007)

Membrane-type 1 matrix metalloproteinase is

regulated by sp1 through the differential activation of

AKT, JNK, and ERK pathways in human prostate

tumor cells. Neoplasia 9, 406–417.

286 Yun S, Dardik A, Haga M, Yamashita A, Yamaguchi

S, Koh Y, Madri JA & Sumpio BE (2002)

Transcription factor Sp1 phosphorylation induced by

shear stress inhibits membrane type 1-matrix

metalloproteinase expression in endothelium. J Biol

Chem 277, 34808–34814.287 Chang HC, Liu LT & Hung WC (2004) Involvement

of histone deacetylation in ras-induced down-

regulation of the metastasis suppressor RECK. Cell

Signal 16, 675–679.288 Sasahara RM, Takahashi C & Noda M (1999)

Involvement of the Sp1 site in ras-mediated

downregulation of the RECK metastasis suppressor

gene. Biochem Biophys Res Commun 264, 668–675.

289 Sasahara RM, Takahashi C, Sogayar MC & Noda M

(1999) Oncogene-mediated downregulation of RECK,

a novel transformation suppressor gene. Braz J Med

Biol Res 32, 891–895.290 Apt D, Watts RM, Suske G & Bernard HU (1996)

High Sp1/Sp3 ratios in epithelial cells during epithelial

differentiation and cellular transformation correlate

with the activation of the HPV-16 promoter. Virology

224, 281–291.291 D’Costa ZJ, Jolly C, Androphy EJ, Mercer A,

Matthews CM & Hibma MH (2012) Transcriptional

repression of E–cadherin by human papillomavirus

type 16 E6. PLoS One 7, e48954.

292 Faraldo ML, Rodrigo I, Behrens J, Birchmeier W &

Cano A (1997) Analysis of the E–cadherin and P-

cadherin promoters in murine keratinocyte cell lines

from different stages of mouse skin carcinogenesis.

Mol Carcinog 20, 33–47.293 Graff JR, Herman JG, Myohanen S, Baylin SB &

Vertino PM (1997) Mapping patterns of CpG island

methylation in normal and neoplastic cells implicates

both upstream and downstream regions in de novo

methylation. J Biol Chem 272, 22322–22329.294 Liu YN, Lee WW, Wang CY, Chao TH, Chen Y &

Chen JH (2005) Regulatory mechanisms controlling

human E–cadherin gene expression. Oncogene 24,

8277–8290.295 Han S & Roman J (2005) COX-2 inhibitors

suppress integrin alpha5 expression in human lung

carcinoma cells through activation of Erk:

involvement of Sp1 and AP-1 sites. Int J Cancer

116, 536–546.296 Nam EH, Lee Y, Park YK, Lee JW & Kim S (2012)

ZEB2 upregulates integrin alpha5 expression through

cooperation with Sp1 to induce invasion during

epithelial-mesenchymal transition of human cancer

cells. Carcinogenesis 33, 563–571.297 Kuo L, Chang HC, Leu TH, Maa MC & Hung WC

(2006) Src oncogene activates MMP-2 expression via

the ERK/Sp1 pathway. J Cell Physiol 207, 729–734.

254 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

298 Pan MR & Hung WC (2002) Nonsteroidal anti-

inflammatory drugs inhibit matrix metalloproteinase-2

via suppression of the ERK/Sp1-mediated

transcription. J Biol Chem 277, 32775–32780.299 Qin H, Sun Y & Benveniste EN (1999) The

transcription factors Sp1, Sp3, and AP-2 are required

for constitutive matrix metalloproteinase-2 gene

expression in astroglioma cells. J Biol Chem 274,

29130–29137.300 Pollak M (2012) The insulin and insulin-like growth

factor receptor family in neoplasia: an update. Nat

Rev Cancer 12, 159–169.301 Chitnis MM, Yuen JS, Protheroe AS, Pollak M &

Macaulay VM (2008) The type 1 insulin-like growth

factor receptor pathway. Clin Cancer Res 14, 6364–6370.

302 Kaytor EN, Zhu JL, Pao CI & Phillips LS (2001)

Insulin-responsive nuclear proteins facilitate Sp1

interactions with the insulin-like growth factor-I gene.

J Biol Chem 276, 36896–36901.303 Mitsudomi T & Yatabe Y (2010) Epidermal growth

factor receptor in relation to tumor development:

EGFR gene and cancer. FEBS J 277, 301–308.

304 Li J, Zou WX & Chang KS (2014) Inhibition of Sp1

Functions by Its Sequestration into PML Nuclear

Bodies. PLoS One 9, e94450.

305 Mart�ınez-Balb�as MA, Dey A, Rabindran SK, Ozato

K & Wu C (1995) Displacement of sequence-specific

transcription factors from mitotic chromatin. Cell 83,

29–38.306 Jin W, Chen BB, Li JY, Zhu H, Huang M, Gu SM,

Wang QQ, Chen JY, Yu S, Wu J et al. (2012) TIEG1

inhibits breast cancer invasion and metastasis by

inhibition of epidermal growth factor receptor

(EGFR) transcription and the EGFR signaling

pathway. Mol Cell Biol 32, 50–63.307 Salvatori L, Ravenna L, Caporuscio F, Principessa L,

Coroniti G, Frati L, Russo MA & Petrangeli E (2011)

Action of retinoic acid receptor on EGFR gene

transactivation and breast cancer cell proliferation:

interplay with the estrogen receptor. Biomed

Pharmacother 65, 307–312.308 Salvatori L, Ravenna L, Felli MP, Cardillo MR,

Russo MA, Frati L, Gulino A & Petrangeli E (2000)

Identification of an estrogen-mediated

deoxyribonucleic acid-binding independent

transactivation pathway on the epidermal growth

factor receptor gene promoter. Endocrinology 141,

2266–2274.309 Koutsodontis G, Vasilaki E, Chou WC, Papakosta P

& Kardassis D (2005) Physical and functional

interactions between members of the tumour

suppressor p53 and the Sp families of transcription

factors: importance for the regulation of genes

involved in cell-cycle arrest and apoptosis. Biochem J

389, 443–455.310 Lin RK, Wu CY, Chang JW, Juan LJ, Hsu HS, Chen

CY, Lu YY, Tang YA, Yang YC, Yang PC et al.

(2010) Dysregulation of p53/Sp1 control leads to

DNA methyltransferase-1 overexpression in lung

cancer. Cancer Res 70, 5807–5817.311 Pietrzak M & Puzianowska-Kuznicka M (2008) p53-

dependent repression of the human MCL-1 gene

encoding an anti-apoptotic member of the BCL-2

family: the role of Sp1 and of basic transcription

factor binding sites in the MCL-1 promoter. Biol

Chem 389, 383–393.312 Krekac D, Brozkova K, Knoflickova D, Hrstka R,

Muller P, Nenutil R & Vojtesek B (2008)

MDM2SNP309 does not associate with elevated

MDM2 protein expression or breast cancer risk.

Oncology 74, 84–87.313 Liu W, Innocenti F, Wu MH, Desai AA, Dolan ME,

Cook EH Jr & Ratain MJ (2005) A functional

common polymorphism in a Sp1 recognition site of

the epidermal growth factor receptor gene promoter.

Cancer Res 65, 46–53.314 Rayess H, Wang MB & Srivatsan ES (2012) Cellular

senescence and tumor suppressor gene p16. Int J

Cancer 130, 1715–1725.315 Esteller M, Corn PG, Baylin SB & Herman JG (2001)

A gene hypermethylation profile of human cancer.

Cancer Res 61, 3225–3229.316 Romagosa C, Simonetti S, Lopez-Vicente L, Mazo A,

Lleonart ME, Castellvi J & Ramon y Cajal S (2011)

p16(Ink4a) overexpression in cancer: a tumor

suppressor gene associated with senescence and high-

grade tumors. Oncogene 30, 2087–2097.317 Oh JE, Han JA & Hwang ES (2007) Downregulation

of transcription factor, Sp1, during cellular senescence.

Biochem Biophys Res Commun 353, 86–91.318 Zhao J, Bilsland A, Hoare SF & Keith WN (2003)

Involvement of NF-Y and Sp1 binding sequences in

basal transcription of the human telomerase RNA

gene. FEBS Lett 536, 111–119.319 Roder K, Wolf SS, Larkin KJ & Schweizer M

(1999) Interaction between the two ubiquitously

expressed transcription factors NF-Y and Sp1. Gene

234, 61–69.320 Liu L, Ishihara K, Ichimura T, Fujita N, Hino S,

Tomita S, Watanabe S, Saitoh N, Ito T & Nakao M

(2009) MCAF1/AM is involved in Sp1-mediated

maintenance of cancer-associated telomerase activity. J

Biol Chem 284, 5165–5174.321 Athanasoula K, Gogas H, Polonifi K, Vaiopoulos

AG, Polyzos A & Mantzourani M (2014) Survivin

beyond physiology: orchestration of multistep

carcinogenesis and therapeutic potentials. Cancer Lett

347, 175–182.

255FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

322 Knight JS, Cotter MA II & Robertson ES (2001) The

latency-associated nuclear antigen of Kaposi’s

sarcoma-associated herpesvirus transactivates the

telomerase reverse transcriptase promoter. J Biol

Chem 276, 22971–22978.323 Bochman ML, Paeschke K & Zakian VA (2012) DNA

secondary structures: stability and function of G-

quadruplex structures. Nat Rev Genet 13, 770–780.324 Todd AK & Neidle S (2008) The relationship of

potential G-quadruplex sequences in cis-upstream

regions of the human genome to SP1-binding

elements. Nucleic Acids Res 36, 2700–2704.325 Kumar P, Yadav VK, Baral A, Kumar P, Saha D &

Chowdhury S (2011) Zinc-finger transcription factors

are associated with guanine quadruplex motifs in

human, chimpanzee, mouse and rat promoters

genome-wide. Nucleic Acids Res 39, 8005–8016.326 Raiber EA, Kranaster R, Lam E, Nikan M &

Balasubramanian S (2012) A non-canonical DNA

structure is a binding motif for the transcription factor

SP1 in vitro. Nucleic Acids Res 40, 1499–1508.327 Cazzalinia O, Scovassib AI, Savioa M, Stivalaa LA &

Prosperi E (2010) Multiple roles of the cell cycle

inhibitor p21CDKN1A in the DNA damage response.

Mutation Res 704, 12–20.328 Abbas T & Dutta A (2009) p21 in cancer: intricate

networks and multiple activities. Nat Rev Cancer 9,

400–414.329 Macleod KF, Sherry N, Hannon G, Beach D, Tokino

T, Kinzler K, Vogelstein B & Jacks T (1995) p53-

dependent and independent expression of p21 during

cell growth, differentiation, and DNA damage. Genes

Dev 9, 935–944.330 Owen GI, Richer JK, Tung L, Takimoto G &

Horwitz KB (1998) Progesterone regulates

transcription of the p21(WAF1) cyclin- dependent

kinase inhibitor gene through Sp1 and CBP/p300. J

Biol Chem 273, 10696–10701.331 Wei Z, Jiang X, Qiao H, Zhai B, Zhang L, Zhang Q,

Wu Y, Jiang H & Sun X (2013) STAT3 interacts with

Skp2/p27/p21 pathway to regulate the motility and

invasion of gastric cancer cells. Cell Signal 25, 931–938.332 Pagliuca A, Gallo P & Lania L (2000) Differential role

for Sp1/Sp3 transcription factors in the regulation of

the promoter activity of multiple cyclin-dependent

kinase inhibitor genes. J Cell Biochem 76, 360–367.333 Zhu WG, Srinivasan K, Dai Z, Duan W, Druhan LJ,

Ding H, Yee L, Villalona-Calero MA, Plass C &

Otterson GA (2003) Methylation of adjacent CpG

sites affects Sp1/Sp3 binding and activity in the p21

(Cip1) promoter. Mol Cell Biol 23, 4056–4065.334 Ganapathy M, Ghosh R, Jianping X, Zhang X,

Bedolla R, Schoolfield J, Yeh IT, Troyer DA, Olumi

AF & Kumar AP (2009) Involvement of FLIP in 2-

methoxyestradiol-induced tumor regression in

transgenic adenocarcinoma of mouse prostate model.

Clin Cancer Res 15, 1601–1611.335 Ulrich E, Kauffmann-Zeh A, Hueber AO, Williamson

J, Chittenden T, Ma A & Evan G (1997) Gene

structure, cDNA sequence, and expression of murine

Bak, a proapoptotic Bcl-2 family member. Genomics

44, 195–200.336 Maksimovic-Ivanic D, Stosic-Grujicic S, Nicoletti F &

Mijatovic S (2012) Resistance to TRAIL and how to

surmount it. Immunol Res 52, 157–168.337 French LE & Tschopp J (2002) Defective death

receptor signaling as a cause of tumor immune escape.

Semin Cancer Biol 12, 51–55.338 Athanasoula K, Gogas H, Polonifi K, Vaiopoulos

AG, Polyzos A & Mantzourani M (2014) Survivin

beyond physiology: Orchestration of multistep

carcinogenesis and therapeutic potentials. Cancer

Letters 347, 175–182.339 Xu R, Zhang P, Huang J, Ge S, Lu J & Qian G (2007)

Sp1 and Sp3 regulate basal transcription of the survivin

gene. Biochem Biophys Res Commun 356, 286–292.340 Wu J, Ling X, Pan D, Apontes P, Song L, Liang P,

Altieri DC, Beerman T & Li F (2005) Molecular

mechanism of inhibition of survivin transcription by

the GC-rich sequence-selective DNA binding

antitumor agent, hedamycin: evidence of survivin

down-regulation associated with drug sensitivity. J

Biol Chem 280, 9745–9751.341 Billon N, Carlisi D, Datto MB, van Grunsven LA,

Watt A, Wang XF & Rudkin BB (1999) Cooperation

of Sp1 and p300 in the induction of the CDK

inhibitor p21WAF1/CIP1 during NGF-mediated

neuronal differentiation. Oncogene 18, 2872–2882.342 Xiao H, Hasegawa T & Isobe K (2000) p300

collaborates with Sp1 and Sp3 in p21(waf1/cip1)

promoter activation induced by histone deacetylase

inhibitor. J Biol Chem 275, 1371–1376.343 Ramos YF, Hestand MS, Verlaan M, Krabbendam E,

Ariyurek Y, van Galen M, van Dam H, van Ommen

GJ, den Dunnen JT, Zantema A et al. (2010)

Genome-wide assessment of differential roles for p300

and CBP in transcription regulation. Nucleic Acids Res

38, 5396–5408.344 Cheng Q, Ling X, Haller A, Nakahara T, Yamanaka

K, Kita A, Koutoku H, Takeuchi M, Brattain MG &

Li F (2012) Suppression of survivin promoter activity

by YM155 involves disruption of Sp1-DNA

interaction in the survivin core promoter. Int J

Biochem Mol Biol 3, 179–197.345 Trisciuoglio D, Iervolino A, Candiloro A, Fibbi G,

Fanciulli M, Zangemeister-Wittke U, Zupi G & Del

Bufalo D (2004) bcl-2 induction of urokinase

plasminogen activator receptor expression in human

cancer cells through Sp1 activation: involvement of

ERK1/ERK2 activity. J Biol Chem 279, 6737–6745.

256 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford

346 Shibuya M (2013) Vascular endothelial growth factor

and its receptor system: physiological functions in

angiogenesis and pathological roles in various diseases.

J Biochem 153, 13–19.347 Kieran MW, Kalluri R & Cho YJ (2012) The VEGF

pathway in cancer and disease: responses, resistance,

and the path forward. Cold Spring Harb Perspect Med

2, a006593.

348 Donovan K, Alekseev O, Qi X, Cho W & Azizkhan-

Clifford J (2014) O-GlcNAc modification of

transcription factor Sp1 mediates hyperglycemia-

induced VEGF–A upregulation in retinal cells. Invest

Ophthalmol Vis Sci 91(s252).

349 Lynch TP, Ferrer CM, Jackson SR, Shahriari KS,

Vosseller K & Reginato MJ (2012) Critical role of O-

Linked beta-N-acetylglucosamine transferase in

prostate cancer invasion, angiogenesis, and metastasis.

J Biol Chem 287, 11070–11081.350 Caldwell SA, Jackson SR, Shahriari KS, Lynch TP,

Sethi G, Walker S, Vosseller K & Reginato MJ (2010)

Nutrient sensor O-GlcNAc transferase regulates breast

cancer tumorigenesis through targeting of the

oncogenic transcription factor FoxM1. Oncogene 29,

2831–2842.351 Issad T, Masson E & Pagesy P (2010) O-GlcNAc

modification, insulin signaling and diabetic

complications. Diabetes Metab 36, 423–435.352 Hart GW, Slawson C, Ramirez-Correa G & Lagerlof

O (2011) Cross talk between O-GlcNAcylation and

phosphorylation: roles in signaling, transcription, and

chronic disease. Annu Rev Biochem 80, 825–858.353 Wang SS, Chen LW, Chen LY, Tsai HH, Shih YC,

Yang CT & Chang PJ (2010) Transcriptional

regulation of the ORF61 and ORF60 genes of

Kaposi’s sarcoma-associated herpesvirus. Virology 397,

311–321.354 Haas TL, Stitelman D, Davis SJ, Apte SS & Madri

JA (1999) Egr-1 mediates extracellular matrix-driven

transcription of membrane type 1 matrix

metalloproteinase in endothelium. J Biol Chem 274,

22679–22685.355 Carneiro P, Fernandes MS, Figueiredo J, Caldeira J,

Carvalho J, Pinheiro H, Leite M, Melo S, Oliveira P,

Simoes-Correia J et al. (2012) E–cadherin dysfunction

in gastric cancer–cellular consequences, clinicalapplications and open questions. FEBS Lett 586,

2981–2989.356 Paredes J, Figueiredo J, Albergaria A, Oliveira P,

Carvalho J, Ribeiro AS, Caldeira J, Costa AM,

Simoes-Correia J, Oliveira MJ et al. (2012) Epithelial

E– and P-cadherins: role and clinical significance in

cancer. Biochim Biophys Acta 1826, 297–311.357 David JM & Rajasekaran AK (2012) Dishonorable

discharge: the oncogenic roles of cleaved E–cadherinfragments. Cancer Res 72, 2917–2923.

358 Banine F, Bartlett C, Gunawardena R, Muchard C,

Yaniv M, Knudsen ES, Weissman BE & Sherman LS

(2005) SWI/SNF chromatin-remodeling factors induce

changes in DNA methylation to promote

transcriptional activation. Cancer Res 65, 3542–3547.359 Basha R, Ingersoll SB, Sankpal UT, Ahmad S, Baker

CH, Edwards JR, Holloway RW, Kaja S & Abdelrahim

M (2011) Tolfenamic acid inhibits ovarian cancer cell

growth and decreases the expression of c-Met and

survivin through suppressing specificity protein

transcription factors. Gynecol Oncol 122, 163–170.360 Pathi S, Jutooru I, Chadalapaka G, Nair V, Lee SO &

Safe S (2012) Aspirin inhibits colon cancer cell and

tumor growth and downregulates specificity protein

(Sp) transcription factors. PLoS One 7, e48208.

361 Pathi S, Li X & Safe S (2013) Tolfenamic acid inhibits

colon cancer cell and tumor growth and induces

degradation of specificity protein (Sp) transcription

factors. Mol Carcinog 53(Suppl 1), E53–E61.362 Ulrich CM, Whitton J, Yu JH, Sibert J, Sparks R,

Potter JD & Bigler J (2005) PTGS2 (COX-2) -765G >C promoter variant reduces risk of colorectal adenoma

among nonusers of nonsteroidal anti-inflammatory

drugs. Cancer Epidemiol Biomarkers Prev 14, 616–619.363 Lian S, Potula HH, Pillai MR, Van Stry M, Koyanagi

M, Chung L, Watanabe M & Bix M (2013)

Transcriptional activation of mina by sp1/3 factors.

PLoS One 8, e80638.

364 Yu J, Wei M, Boyd Z, Lehmann EB, Trotta R, Mao

H, Liu S, Becknell B, Jaung MS, Jarjoura D et al.

(2007) Transcriptional control of human T-BET

expression: the role of Sp1. Eur J Immunol 37, 2549–2561.

365 Kim SH, Shanware NP, Bowler MJ & Tibbetts RS

(2010) Amyotrophic lateral sclerosis-associated

proteins TDP-43 and FUS/TLS function in a common

biochemical complex to co-regulate HDAC6 mRNA. J

Biol Chem 285, 34097–34105.366 Laskow TC, Buchser WJ, Pitt BR, Basse PH,

Butterfield LH, Kalinski P & Lotze MT (2010) Zinc in

innate and adaptive tumor immunity. J Transl Med 8,

118.

367 Astrinidis A, Kim J, Kelly CM, Olofsson BA, Torabi B,

Sorokina EM & Azizkhan-Clifford J (2010) The

transcription factor SP1 regulates centriole function and

chromosomal stability through a functional interaction

with the mammalian target of rapamycin/raptor

complex. Genes Chromosom Cancer 49, 282–297.368 Jensen DE, Black AR, Swick AG & Azizkhan JC

(1997) Distinct roles for Sp1 and E2F sites in the

growth/cell cycle regulation of the DHFR promoter. J

Cell Biochem 67, 24–31.369 Black AR, Jensen D, Lin SY & Azizkhan JC (1999)

Growth/cell cycle regulation of Sp1 phosphorylation. J

Biol Chem 274, 1207–1215.

257FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

K. Beishline and J. Azizkhan-Clifford Sp1 and the hallmarks of cancer

370 Gueven N, Fukao T, Luff J, Paterson C, Kay G,

Kondo N & Lavin MF (2006) Regulation of the Atm

promoter in vivo. Genes Chromosom Cancer 45, 61–71.371 Gardiner-Garden M & Frommer M (1987) CpG islands

in vertebrate genomes. J Mol Biol 196, 261–282.372 Jiang Y, Liu S, Chen X, Cao Y & Tao Y (2013)

Genome-wide distribution of DNA methylation and

DNA demethylation and related chromatin regulators

in cancer. Biochim Biophys Acta 1835, 155–163.373 Brandeis M, Frank D, Keshet I, Siegfried Z,

Mendelsohn M, Nemes A, Temper V, Razin A &

Cedar H (1994) Sp1 elements protect a CpG island

from de novo methylation. Nature 371, 435–438.374 Siegfried Z, Eden S, Mendelsohn M, Feng X, Tsuberi

BZ & Cedar H (1999) DNA methylation represses

transcription in vivo. Nat Genet 22, 203–206.375 Clark SJ, Harrison J & Molloy PL (1997) Sp1 binding

is inhibited by (m)Cp(m)CpG methylation. Gene 195,

67–71.

376 Strunnikova M, Schagdarsurengin U, Kehlen A,

Garbe JC, Stampfer MR & Dammann R (2005)

Chromatin inactivation precedes de novo DNA

methylation during the progressive epigenetic silencing

of the RASSF1A promoter. Mol Cell Biol 25, 3923–3933.

377 Liu S, Liu Z, Xie Z, Pang J, Yu J, Lehmann E,

Huynh L, Vukosavljevic T, Takeki M, Klisovic RB

et al. (2008) Bortezomib induces DNA

hypomethylation and silenced gene transcription by

interfering with Sp1/NF-kappaB-dependent DNA

methyltransferase activity in acute myeloid leukemia.

Blood 111, 2364–2373.378 Cruickshanks HA, McBryan T, Nelson DM,

Vanderkraats ND, Shah PP, van Tuyn J, Singh Rai T,

Brock C, Donahue G, Dunican DS et al. (2013)

Senescent cells harbour features of the cancer

epigenome. Nat Cell Biol 15, 1495–1506.

258 FEBS Journal 282 (2015) 224–258 ª 2014 FEBS

Sp1 and the hallmarks of cancer K. Beishline and J. Azizkhan-Clifford