Next Article in Journal
18F-FDG PET/CT and Urothelial Carcinoma: Impact on Management and Prognosis—A Multicenter Retrospective Study
Next Article in Special Issue
Deciphering The Potential Role of Hox Genes in Pancreatic Cancer
Previous Article in Journal
NECTIN4 (PVRL4) as Putative Therapeutic Target for a Specific Subtype of High Grade Serous Ovarian Cancer—An Integrative Multi-Omics Approach
Previous Article in Special Issue
The Role of HOX Transcription Factors in Cancer Predisposition and Progression
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Paralogous HOX13 Genes in Human Cancers

1
Scientific Direction, Istituto Nazionale Tumori IRCCS Fondazione G. Pascale, 80131 Naples, Italy
2
Department of Clinical Medicine and Surgery, School of Medicine, University of Naples Federico II, 80131 Naples, Italy
3
Pathology Unit, Istituto Nazionale Tumori IRCCS Fondazione G. Pascale, 80131 Naples, Italy
4
Pathological Anatomy Laboratory, Casa di Cura Maria Rosaria, 80045 Pompei, Italy
*
Author to whom correspondence should be addressed.
Cancers 2019, 11(5), 699; https://doi.org/10.3390/cancers11050699
Submission received: 1 April 2019 / Revised: 17 April 2019 / Accepted: 16 May 2019 / Published: 20 May 2019
(This article belongs to the Special Issue HOX Genes in Cancer)

Abstract

:
Hox genes (HOX in humans), an evolutionary preserved gene family, are key determinants of embryonic development and cell memory gene program. Hox genes are organized in four clusters on four chromosomal loci aligned in 13 paralogous groups based on sequence homology (Hox gene network). During development Hox genes are transcribed, according to the rule of “spatio-temporal collinearity”, with early regulators of anterior body regions located at the 3’ end of each Hox cluster and the later regulators of posterior body regions placed at the distal 5’ end. The onset of 3’ Hox gene activation is determined by Wingless-type MMTV integration site family (Wnt) signaling, whereas 5’ Hox activation is due to paralogous group 13 genes, which act as posterior-inhibitors of more anterior Hox proteins (posterior prevalence). Deregulation of HOX genes is associated with developmental abnormalities and different human diseases. Paralogous HOX13 genes (HOX A13, HOX B13, HOX C13 and HOX D13) also play a relevant role in tumor development and progression. In this review, we will discuss the role of paralogous HOX13 genes regarding their regulatory mechanisms during carcinogenesis and tumor progression and their use as biomarkers for cancer diagnosis and treatment.

1. Introduction

Homeobox genes, a superfamily of transcription factors, are considered as key determinants of embryonic development and body plan organization [1]. Among them, Hox genes (HOX in humans) represent a conserved gene family controlling antero-posterior axis and dorsum-ventral organization during development [2]. HOX genes are organized into the HOX gene network: four chromosomal loci (HOXA chr.7p15, HOXB chr.17q21, HOXC chr.12q13, HOXD chr.2q31-33) [3] for a total of 39 genes aligned into 13 vertical paralogous groups on the basis of sequence similarity and position in the loci [4]. The Hox gene network organization has evolved from a single ancestral proto-Hox gene through replications and transpositions [5].
The Hox gene network cooperates with the Polycomb and Trithorax gene families to mediate the cell memory program, with the Hox genes regulating mRNA transcription and Polycomb and Trithorax genes blocking or inducing DNA-chromatin interaction, respectively [6]. However, while Polycomb and Trithorax gene family members are dispersed throughout the genome [7], the HOX genes exists in tightly regulated clusters that form the largest physically and functionally identifiable network in the human genome as reported by Lander et al. [8].
During mammalian development, Hox genes begin to be expressed at gastrulation, controlling the identity of body regions, along the antero-posterior axis, according to the rules of spatio-temporal collinearity [9]. Hox genes are activated at the 3’ end loci in response to Wnt3 signals. Subsequent induction of Hox genes located at the 5’ end, following the same dynamics, generates the antero-posterior elongation. The activation of the 5’ end HOX genes is controlled by paralogous group 13 HOX genes, which play a crucial role in the retro-inhibition of the functions of more anterior Hox proteins, limiting axial elongation. This negative-dominant effect, called “posterior prevalence”, is due to the prevalent role of posterior Hox proteins on more anterior genes of the Hox network. The whole process induces the translation from temporal information to a series of spatial coordinates (spatio-temporal collinearity) [10]. Studies on the expression of the HOX network have shown unique patterns (HOX profiles) in different tissues and organs. These observations support the conclusion that: (i) the HOX gene network expression model for each organ is the sum of the different HOX profiles of the organ resident cells; (ii) the organ specific HOX profiles contribute to the anatomical organization of the organs and to the appropriate positioning of each organ along the antero-posterior axis of the body [11].
The post-genomic era has identified the HOX network as one of the richest areas of the genome for its content of non-coding RNAs (ncRNAs), including both microRNAs (miRNAs) and long-non-coding-RNAs (lncRNAs) [12]. Approximately 240 ncRNAs have been identified within the HOX network and less than 10% have been functionally characterized so far. Among them, key players are represented by HOTTIP, for its localization (adjacent to HOXA13)/interaction with posterior locus A HOX genes [13,14] and HOTAIR for its property to regulate in trans HOXD locus [15]. The understanding of the HOX network organization and activity has been profoundly boosted by the identification of the non-coding RNA world.
The comparison of HOX profiles in normal and neoplastic tissues has allowed the identification of HOX genes dysregulated in both tumor- and tissue-specific manner [16,17]. Their deregulation during tumor progression is associated with the tumor heterogeneity, epithelial mesenchymal transition and metastasis [18,19]. Since HOX genes also control important metabolic processes, the identification of unique HOX profiles associated to primary tumors has provided an important glimpse on the deregulation of primary metabolic processes contributing to specific tumor phenotypes and, more generally, to the biology of cancer cells [20,21]. Although the entire HOX network plays a central role in cancer development and progression, the most posterior genes of the network, HOX13 paralogues (HOXA13, HOXB13, HOXC13 and HOXD13), specifically, are crucial in modulating these processes, in cooperation with co-localizing lncRNAs.
During development, all HOX13 paralogues are involved in mediating the transition from the early to the late-distal limb program, controlling the spatial expression patterns of target genes [22] and in mediating gut and urogenital system formation [23,24,25]. However, many of them are still active in adult human organs and tissues [26] and frequently deregulated in human cancers [26,27,28,29].
Due to the crucial role played by paralogous HOX13 genes on the HOX gene network organization during embryonic development, as well as in tumor progression, in this review we will discuss: (i) their physiological and pathological regulatory mechanisms during carcinogenesis; (ii) their ability to modulate tumor progression also through the interaction with co-localized ncRNAs; (iii) their potential applications as biomarkers for cancer diagnosis and treatment.

2. HOXA13

During normal development, HOXA13 plays a leading role in the creation of posterior structures of the body, specifically in limb, gut and urogenital system development [22,24,30]. During abnormal development, modifications of HOXA13 homeoprotein have been associated with “Hand-foot-genital” syndrome [31].
HOXA13 deregulation in cancer has been recently validated by a meta-analysis study. In 844 tumor patient biopsies, enrolled in nine different studies, the aberrant expression of HOXA13 is significantly associated with poor histological grade, TNM stage and overall survival, suggesting that HOXA13 is a potentially valuable biomarker of poor prognosis and potential therapeutic target for human tumors. [32].
HOXA13 deregulation has been strongly associated with urological cancers (bladder and prostate). High levels of HOXA13 homeoprotein have been found in bladder cancer tissues and positively correlated with lymph nodes metastases, TNM stage, pathological grade and patient survival [33]. Moreover, since circulating bladder tumor cells can be carried in the urine, HOXA13 level has been also detected in this biological fluid. cDNA microarray analysis has highlighted that co-expression of HOXA13 and BLCA-4 could be able to discriminate low versus high grade bladder tumors [34]. These data have been then confirmed by STRING database screening which, in turn, has allowed the identification of a four-gene model, including IGF-1, hTERT, BLCA-4 and HOXA13 to stratify bladder cancers at different stages [35]. Aberrant HOXA13 expression has been reported in prostate cancer (PC), where nuclear HOXA13 expression is strongly associated with histological grade and Gleason score. Moreover, induced HOXA13 expression in PC cell models promoted cell proliferation, migration, invasion and inhibition of apoptosis [36]. HOXA13 is able to frequently form a fusion gene with nucleoporin NUP98, named NUP98-HOXA13, playing a key role in acute myeloid leukemia (AML) [37,38]. A chromosomal translocation between an upstream HOXA13 region and a downstream region of the BCL11B/CTIP2 locus has been described in T-cell acute lymphoblastic leukemia (T-ALL), resulting in a HOXA13 gene hyper-expression [39]. HOXA13 is strongly up-regulated in gastric cancer (GC) tissues compared to normal adjacent mucosa. HOXA13 aberrant expression correlates with GC tumor stage, histological differentiation and survival of the patient [40]. HOXA13 is also hyper-expressed in gastric stem cells [41] and its knockdown, in GC cell model, modulates epithelial-mesenchymal-transition (EMT) reducing cell invasion features [42]. A recent study has highlighted that HOXA13 over-expression significantly increases cadherin17 (CDH17) gene expression in GC cells and tissues. The simultaneous knockdown of both genes leads to a reduction of cell proliferation and invasion and promotes apoptosis in GC cell model [43]. Han and collaborators have lately shown that HOXA13 is strongly related to 5-fluorouracil (5-FU) resistance in GC patients. HOXA13 might confer resistance to GC cells by a p53-dependent pathway [44]. HOXA13 deregulation has been further associated with Disease Free Survival (DFS) in esophageal squamous cell carcinoma (ESCC) [45,46]. The knockdown of HOXA13 in ESCC cell model leads to a reduced number of colonies in vitro and tumor growth in nude mice [46]. The coordinated expression of HOXA13, ANXA2 and SOD2 strongly predicts poor prognosis in ESCC [47]. HOXA13 is also involved in the modulation of EMT in ESCC cells [48]. In ESCC patients treated with neoadjuvant chemotherapy, HOXA13 expression is associated with the worst tumor regression grade. The knockdown of HOXA13, in ESCC cells, increases cis-platinum-induced apoptosis, suggesting an essential role of HOXA13 in drug-resistance acquisition [48]. In a further investigation, an opposite trend of HOXA13 expression has been detected in oral squamous cell carcinoma (OSCC): a prevalent HOXA13 expression, in the superficial side of the lesions, is significantly associated to a better prognosis of OSCC patients [27].
There is also growing evidence that HOXA13 has a role in liver cancer. HOXA13 is over-expressed in primary hepatocellular carcinoma (HCC) and is strongly associated with hepatitis B and C virus infection. In addition, its expression has been detected in HCC cell lines originating from liver stem-like cells, suggesting the HOXA13 role in the differentiation and tumor evolution of hepatic stem cells [49]. The profile of the whole HOX network in a large cohort of paired liver biopsies, HCC versus their non-neoplastic counterparts, has identified the locus A HOX gene as the most dysregulated locus among the HOX loci and HOXA13 is systematically over-expressed in HCCs versus normal/non-neoplastic livers. The study has demonstrated that HCC samples with high HOXA13 expression manifest the dysregulation of a gene set associated to poor prognosis, according to HCC transcriptome classification. Furthermore, HOXA13 homeoprotein physically interacts with the cap-binding protein eIF4E, deregulated in HCC [50]. HOXA13 expression in HCC patients is also strongly correlated with the expression of angiogenic markers, such as VEGF, microvessels density and alpha-fetoprotein (AFP) serum levels. In addition, serum HOXA13 levels have been detected in 90 HCC patients suggesting that its circulating level could be used for early HCC diagnoses and prediction of the outcomes [51]. In HCC in vitro model, HOXA13 further correlates with poor differentiated HCC modulating sorafenib response [52]. The deregulation of HOXA13 has been also described in lung cancer. The expression data of HOXA13 have been collected from different databases, highlighting its aberrant expression mainly in lung adenocarcinoma progression [53]. In addition, Kang and collaborators have described a frequent gain of copies number on the short arm of chromosome 7 containing the whole locus HOXA, suggesting its critical role in lung adenocarcinoma evolution [54]. HOXA13 deregulation has been sporadically associated to other cancer phenotypes, such as ovarian cancer associated with poor clinical outcome [55], in glioma associated with tumor progression thought Wnt and TG-Beta pathways modulation [56] and thyroid cancers where HOXA13 nuclear expression is associated with different histotypes [29].
In recent studies, the aberrant role of HOXA13 in cancer is frequently associated with HOTTIP expression, suggesting that their interaction is strongly related to the modulation of tumor evolution and progression. LncRNA HOTTIP (HOX transcript at distal tip) has been functionally characterized in 2011 by Chang [56]. HOTTIP is located at the 5’ end of the locus HOXA on chromosome 7p15, adjacent to HOXA13. According to its genomic localization, HOTTIP is active, from development to adulthood, in lumbar-sacral body regions [57]. Through interaction with the activator Trithorax complex WDR5/MLL(H3K4me3), HOTTIP is able to promote the activation of a block at 5’ end of HOXA locus genes, from HOXA13 to HOXA9 [57]. In the mouse model, HOTTIP knock-out generates alterations very similar to HOXA11 and HOXA13 inactivation, supporting its role in the functional control of the locus HOXA lumbar-sacral region [57]. Based on the physical and functional interaction between HOTTIP and HOXA13, as well as the role played by HOXA13 in HCC, the role of HOTTIP in HCC has been also investigated accordingly. HOTTIP is significantly over-expressed in HCCs versus normal liver tissues, likewise HOXA13 [14]. The levels of HOXA13 and HOTTIP are able to predict HCC prognosis being associated to metastasis status, clinical outcome and patients’ survival [14]. LncRNA HOTTIP has been described as over-expressed in pancreatic cancers promoting tumor progression and EMT. HOTTIP is also highly active in pancreatic stem cells affecting stem cell factors (LIN28, NANOG, OCT4 and SOX2) and markers (ALDH1, CD44 and CD133) through a mechanism involving HOTTIP/WDR5/HOXA9/Wnt/beta-catenin axis. [58]. HOXA13-HOTTIP interaction has been also documented in GC progression: in GC cell model the knockdown of HOTTIP is strongly related to poor differentiated GC, TNM stage and lymph nodes metastasis [59]. HOXA13 is able to trans-activate IGFBP3 promoter in GC cells promoting their oncogenic potential. Its knockdown leads to a deregulation of both HOTTIP and IGFBP3, suggesting that HOXA13/HOTTIP/IGFBP3 cascade is strongly involved in GC carcinogenesis [60]. The main role of HOTTIP and HOXA13 in prostate cancer has been underlined by the recent finding that PC risk elements are mainly related to HOXA13 and HOTTIP expressions, but not to other HOXA locus genes [61]. In addition, the knockdown of HOXA13 and HOTTIP in PC cell models leads to a deregulation of different cell cycles and cell growth pathways [62]. The combined aberrant expression of HOXA13/HOTTIP related to the promotion of cell proliferation and in vivo/in vitro metastases, has been also reported in ESCC cells [63] and in non-small-cell-lung cancer (NSCLC) [64].
The coordinated dysregulation of HOTTIP and HOXA13 in different tumor types supports their strong interaction and involvement in general mechanisms of neoplastic transformation, regardless of specific tumor phenotypes.

3. HOXB13

HOXB13 plays an important role in dermis development, as well as being involved with other HOX13 paralogues in the formation and organization of posterior body structures [65,66] and in the modulation of prostate differentiation and function [67] by regulating the response to androgens [68].
The role of HOXB13 in human cancer has been mostly associated to breast cancer (BC) and PC. Initial analyses on BC have highlighted the over-expression of HOXB13 in tumor tissues compared to adjacent non-neoplastic mammary gland [69]. Later on, Ma and collaborators, through a gene expression profiling study on a case series of estrogen receptor positive (ER+) BC patients treated with adjuvant tamoxifen (TAM), have identified two genes differentially over-expressed, HOXB13 in TAM recurrences and IL17BR in non-recurrences patients, suggesting the existence of a HOXB13/IL17BR ratio to predict TAM response in BC patients [69]. Subsequent studies have validated the two-gene expression ratio for therapeutic stratification of ER+BC patients [70,71,72]. High HOXB13/IL17BR expression also represents a strong independent prognostic factor in ER+ node-negative (N-) BC patients [73], even regardless of TAM therapy [74]. In addition, a recent meta-analysis has enrolled 11 BC studies, with 2958 participants, concerning the use of HOXB13/IL17BR ratio in association with a worse outcome, particularly for (N-) patients [75]. However, other validation studies have shown that even HOXB13 alone is able to predict recurrence-free survival after endocrine therapy [76,77]. Ma and collaborators have further described the molecular grade index (MGI), a five cell-cycle gene test that, in combination with HOXB13/IL17BR ratio, identifies a subgroup of early ER+BC patients with a very poor clinic outcome, despite endocrine therapy [78]. This finding suggests the identification of a new predictive test, named Breast Cancer Index (BCI), a risk index based on a combination of MGI and HOXB13/Il17BR ratio. A case-control study, performed on an independent cohort of (N-) BC patients, in order to compare the predictive value of HOXB13/IL17BR, MGI and BCI index, has shown that the three tests are associated with the risk of BC death and display a strong prognostic value in BC management [79]. Over time, BCI has become one of the most widely used predictive tests in prognostic and therapeutic stratification of patients with early-stage (N-) and lymph nodes + (L+) BC treated with TAM alone, as well as with TAM+ocretide [80].
One of the epigenetic processes responsible for the regulation of HOXB13 is related to methylation of its promoter, responsible for a reduced expression in BC cell lines. Promoter hyper-methylation of HOXB13 is more frequent in ERα+ BC patients with lymph nodes metastases. This could explain why HOXB13 up-regulation has been described in BC patients undergoing TAM therapy [81]. Furthermore, HOXB13 would confer resistance to TAM by directly regulating ERα transcription and protein expression and it is known to be able to transcriptionally up-regulate IL-6 activating the mTOR pathway [82]. An alternative mechanism related to TAM resistance may be due to HOXB13 interaction with HBXIP, an oncogenic protein promoting cancer progression. The coordinated over-expression of HOXB13 and HBXIP induces TAM resistance in ERα BC cell models: HBXIP prevents chaperone-mediated-autophagy (CMA)-dependent degradation of HOXB13 through the acetylation of its K227 residue, causing HOXB13 accumulation. HBXIP further acts as co-activator of HOXB13 to stimulate IL-6 transcription and promoting TAM resistance [83].
The involvement of HOXB13 in prostate gland development dates back more than 20 years and its role in PC has been soon after demonstrated [84]. Induced HOXB13 expression in PC cell models leads to cell growth inhibition with G1 cell cycle arrest linked to cyclin D1 suppression, suggesting a central role of HOXB13 as PC tumor suppressor [85]. HOXB13 interacts directly with Androgen Receptor (AR) by suppressing the hormone-mediate AR activity in a dose-responsive manner, influencing growth regulation of PC cells [86]. HOXB13 has been also investigated in androgen-independent PC, resulting over-expressed in hormone-refractory tumors. The ability of HOXB13 to modulate PC cell growth in the absence of androgen is mediated by RB-E2F signaling and inhibition of p21waf tumor suppressors [87]. The expression of HOXB13 homeoprotein by immunohistochemistry in PC patients correlates with Gleason Score (GS) and pre-operative circulating PSA levels, but no correlation with clinic-pathological features has been detected [88]. HOXB13 represents a specific biomarker of PC cells and is useful for the differential diagnosis of tumor origin, prostate versus urothelium [89], and for distinguishing metastatic tumors of prostatic origin [90]. HOXB13 has been also recently reported as sensitive and specific biomarker in pleomorphic giant cell prostate adenocarcinoma [91]. On a case series of 12400 PC samples, the aberrant expression of HOXB13 is associated with pT stage, high GS, lymph node metastases, AR expression, high pre-operative PSA level and genetic alterations, such as PTEN deletion and TMPRSS2:ERG translocation [92]. A high expression of HOXB13, AR and PSA identifies a subset of patients with a worse PC prognosis. Combined analyses of HOXB13 and PSA, in metastatic PCs, shows that only HOXB13 is able to distinguish metastatic PCs with high sensitivity and specificity [92]. HOXB13 is able to interact with different molecular pathways during PC evolution: (i) suppresses Prostate Derived ETS Factor (PDEF) [93]; (ii) suppresses p21 in castration-resistant PC, stressing this important step in PC cell survival under no androgen-influence [94]; (iii) promotes PC cell invasion and metastasis by decreasing intracellular zinc levels, enhancing NF-kB [95]. Moreover, a direct biochemical and functional interaction has been described between HOXB13 and MEIS1 in PC cells: the corresponding two proteins are co-expressed on PC tissues with the consequence of modulating PC tumor progression by prolonging HOXB13 half-life [96].
The most numerous data in literature are related to identification of germline mutations in HOXB13 sequences strongly associated with hereditary PC. The same non-synonymous mutation, a change of guanine to adenosine (c251G—A) in the second position of codon 84 (GGA—GAA), resulting in a substitution of glycine for glutamic acid (G84E), has been observed in four families with PC subjects [97]. Several population studies have associated HOXB13 G84E variant to PC risk, especially in Europe [98,99], while in Asia, in addition to G84E, another mutation (G135E) would seem to be prevalent [100]. A study designed to analyze the distribution of the variant by ethnicity, has highlighted that G84E HOXB13 is more frequent among PC patients of European decent [101]. In order to validate the value of G84E variant in clinic PC management, 2443 PC families have been recruited by the International Consortium for Prostate Cancer Genetics (ICPCG) to genotype the mutation. This study has shown that HOXB13 G84E is present in approximately 5% of PC families, mainly of European descent, confirming its association with PC risk [102]. Other studies have shown that G84E variant does not specifically characterize men with a family history of PC, but it is rather strongly associated with PC in the general population [103]. The pathogenic mechanisms associated with PC patients with G84E variant, displayed that they are characterized by the alteration of specific molecular pathways, in particular a low prevalence of better documented EGR pathways and an increased prevalence of SPINK1 pathway [104]. Several other rare missense mutations of HOXB13 gene associated with a predisposition to PC have been identified (Y88D, L144P, G216C, R229G) [97]. In addition, the analyses of the entire HOXB13 gene in 462 Portuguese familial PC subjects have revealed the presence of two novel germline mutations, supporting the concept that different rare HOXB13 mutations could be found in different ethnic groups [105].
HOXB13 appears to be down-regulated in about 60% of colorectal cancers (CRC). In CRC cell models, HOXB13 down-regulates the expression of T-cell-factor 4 (TCF4) and its target c-myc, inhibiting β−catenin/TCF mediated signaling. The induced expression of HOXB13 leads to the suppression of cell growth in CRC cells [106]. HOXB13 is strongly methylated in CRC cells and inhibits growth and clonogenic survival in vitro as well as in nude mice [107]. However, a recent study has shown the aberrant expression of HOXB13 in proximal colon cancers and a strong relation with lncRNA HOTAIR deregulation [108]. G84E HOXB13 mutation is very frequent also in CRC patients, suggesting an association of G84E variant with CRC risk, as occur in PC [109].
Abnormal HOXB13 expression has been detected in uro-genital cancers. Ovarian, cervical and endometrial cancers display HOXB13 over-expression promoting EMT, cell invasion and tumor progression [110,111,112]. In bladder cancer, HOXB13 is able to discriminate between non-muscle and muscle invasive transitional BC and its cytoplasmic de-localization represents an important prognostic value in BC patients [113]. In renal cancer HOXB13 acts as a tumor suppressor and its methylation status positively correlates with tumor grade and micro-vessels invasion [114]. The role of HOXB13 tumor suppressor has been also demonstrated in gastric cancer [114] in which HOXB13 mRNA is significantly lower in primary tumors and its de-regulation is associated with a poorer differentiation, lymph node metastases, invasion and TNM stage. HOX B13 expression is increased by the treatment of GC cells with DNA methyltransferase inhibitor 5-aza-dC[115]. In HCC patients, aberrant expression of HOXB13 is strongly associated with clinic-pathological features, such as vascular invasion, tumor grade, TNM stage and a poorer survival. In addition, it is significantly correlated with VEGF expression and microvessels density, suggesting a central role of HOXB13 in HCC angiogenesis and tumor progression [116]. Finally, an altered HOXB13 expression has been described in oral cancers [117,118] and glioma patients [119].
HOXB13 appears to be under the control of a co-localizing lncRNA. HOXB13-AS1 is a 564 nucleotide- lncRNA localized on chromosome 17, adjacent to HOXB13 [120], highly expressed in several normal human tissues [121,122] and different tumor types [119,122]. HOXB13-AS1 is able to contribute to cancer cells proliferation by binding with the enhancer of zeste homolog 2 (EZH2), epigenetically suppressing HOXB13 expression of its neighbor gene [119].

4. HOXC13

During embryonal development, HOXC13 is involved in the formation of skin epithelia [123] and hair follicle generation [124]. HOXC13 is an important regulator of human keratin gene expression in early trichocyte differentiation [125,126,127,128]. Aberrant HOXC13 expression has been reported in pilomatricoma [129] together with a strong expression of K5, K14 and K17 cytokeratin [130]. It is worth noting that the genes of C HOX locus are in physical contiguity to one of the two clusters of keratin genes included in the human genome [131]. Over-expression of HOXC13 in transgenic mice is also able to develop alopecia and progressive pathological skin conditions [132].
The involvement of HOXC13 in cell cycle progression, cell growth and carcinogenesis has been well documented: knocking down HOXC13 in human cancer cell lines, such as colorectal, breast, prostate and cervical cancer, significantly affects the viability of cancer cells [133]. HOXC13 silencing further induces cell death leading to cell cycle arrest, at G0/G1 phase, and increasing apoptosis. Finally, HOXC13-induced-expression leads to 3D-colony-formation in soft agar, highlighting its role in cell proliferation and invasion [133]. In human metastatic melanoma cell lines, HOXC13 deregulation, along with the other C HOX genes of the posterior locus, is strongly related to the expression of IL-1α, IL-6, TNFα, VLA-2, VLA-5 and VLA-6 integrins and N-RAS mutation [134]. Maeda and collaborators have shown that expression levels of HOXC13 is are higher in nevi and pT1/pT2 than in patients with pT4 melanoma, decreasing in more advanced tumors [135]. However, this data appear to be in contrast with another study in which a series of human biological samples (tissues and cells), representative of malignant melanoma progression, display that HOX C13 expression significantly increases in metastases compared to primary tumors [19]. It is well documented the ability of posterior HOX gene to generate fusion transcripts with the nucleoporin NUP98. The chimera protein NUP98/HOXC13 has a pathogenic importance in acute myeloid leukemia (AML), which leads to the deletion of the mutual fusion of the gene [136,137,138]. Subsequent molecular analyses have highlighted the fusion in frame of exon 16 of NUP98 with exon2 of HOXC13. This translocation appears to coexist with an internal tandem duplication of the gene FLT3 (fms-related tyrosine-kinase 3). Both events are crucial for the leukemiogenesis process [139]. Rare cases of AML with NUP98 rearrangements without HOXC13 involvement have also been described [140]. Aberrant HOXC13 expression has also been reported in murine models of erythroleukemia (MEL). HOXC13 binds to ETS domain of the gene PU.1 enhancing its transactional activity. The induced expression of HOXC13 and PU.1 in MEL inhibits cell differentiation suggesting a primary role in tumor cell differentiation [141]. HOXC13 is highly expressed in ameloblastoma tissues compared to keratocystic odontogenic tumors and normal mucosa [142]. Furthermore, the whole HOXC locus of the HOX network appears to be deregulated in ameloblastoma together with the keratin genes that co-localize in the same 12q13.13 chromosomal area [143]. A large study enrolling different odontogenic tumors, such as ameloblastomas, calcifying cystic odontogenic tumors, ameloblastic fibromas, keratocystic odontogenic tumors and epithelial odontogenic tumors displayed an over-expression of HOXC13 in all lesions except in fibromas [144]. HOXC13 is also over-expressed in OSCC cell models in association with altered activity of the Polycomb Repressive Complex (PCR) responsible for epigenetic modifications [145]. DNA methylation and histones alterations are strongly associated with HOX gene expression in OSCC models [146]. HOXC13 is significantly up-regulated in ESCC in association with poorer clinic-pathological features and worse prognosis of the patients, and its knockdown decreases cell proliferation and induces apoptosis in ESCC cells [147]. In cervical cancer cell model, BMI-1, a gene which encodes a ring finger protein that is the major component of the polycomb group complex 1 (PRC1), is able to modulate HOXC13 expression. The knockdown of BMI-1 in this model leads to an over-expression of HOXC13 inducing cell-cycle arrest [148]. IHC HOXC13 over-expression has also been detected in well-differentiated and de-differentiated liposarcoma tissues, in which it is strongly related to 12q13-15 chromosomal amplification [149]. A four gene signature, including HOXC13, has been proposed in PC patients, to discriminate between recurrent versus non recurrent PC and to predict the outcome of the disease [150]. HOXC13 is significantly higher in tissues and cell lines of lung adenocarcinoma, in correlation with clinic-pathological features and poorer prognosis. Knockdown of HOXC13 in lung cancer cell models inhibits cell proliferation blocking G1 phase of cell-cycle [151]. HOXC13 is further down-regulated by miR141 in lung cancer cell lines [151]. Finally, it has been recently shown that HOXC13 is over-expressed in proximal colon cancer, which strongly correlates with lymph nodes metastases and aberrant expression of lncRNA HOTAIR [108].
The 5’ end HOXC region contains several lncRNAs, including HOXC13-AS (adjacent to HOXC13), HOXC-AS2, HOXC-AS3 transcripts [152], and HOTAIR. However, none of them plays a role in the interaction/regulation of HOXC13, both in normal and pathological conditions. HOXC13-AS is highly expressed in head and neck squamous carcinoma (HNSC) tissues and its aberrant expression is detectable in nasopharyngeal carcinoma (NPC) tissues and cell line. Knockdown of HOXC13-AS leads to an increase of NPC cell proliferation, migration and invasion [153].
LncRNA HOX Transcript Antisense Intergenic RNA (HOTAIR) located on chr.12q13.13 (between HOXC11 and HOXC12) is able to transcriptionally repress in trans the 5’ end of HOXD locus on chr. 2q32-33. [154,155]. HOTAIR acts as a regulator of chromatin states by binding PRC2, with its 5’end. HOTAIR further interacts, through its 3’ end, with LSD1 (lysine-specific demethylase 1), a central player in epigenetic regulation. HOTAIR is able to: (i) promote the epigenetic activation/repression of gene expression; (ii) affect the target suppression of gene expression through competitive binding to miRNAs; (iii) modify gene expression, at post-transcriptional level, interacting with transcription factors and ribosomes or binding to splicing factors [15]. Aberrant HOTAIR expression has been detected in several human cancers associating its role with tumor proliferation, angiogenesis, progression, drug resistance and worse prognosis [156]. In addition, numerous experimental evidences have focused the attention on the potential role of HOTAIR as circulating marker in cancer patients and as potential therapeutic target [157]. A potential relation between HOXC13 and HOTAIR expression has been shown only in a recent study on colon cancer in which both genes are co-expressed in the right CRCs samples and are correlated with lymph nodes metastases [108].

5. HOXD13

During development, HOXD13 plays a central role in the formation of the limbs [158,159], in a part of the gut [160] and in genitourinary system function [25]. Mutations in its structure (expansions of a polyalanine stretch in the amino-terminal region) are able to generate synpolydactyly, an inherited human abnormality of the hands and feet [161].
HOXD13 deregulation in human cancers mostly concerns haematological malignancies, less frequently than in other types of tumor. HOXD13, as the other posterior HOX genes, is involved in a chromosomal translocation in AML with the generation of a chimeric protein between HOXD13 and NUP98 [162]. The analyses of the fusion gene, in a murine hematopoietic model, show its involvement in the growth and differentiation of early hematopoietic progenitor cells. In addition, co-transduction of NUP98-HOXD13 transcript, plus Meis1 cofactor, induces lethal AML in mice models, highlighting their fundamental role in leukemic transformation [163]. During leukemic evolution, NUP98-HOXD13 interacts, besides Meis1, with MN1, a transcriptional co-activator forming fusion transcripts with TEL, GATA2, ERG, Epor and miR291/miR29b1 genes [164]. However, co-transduction of an activated NUP98-HOXD13 fusion gene and MN1 alone is able to induce AML in engrafted mice [165]. More recently, it has been demonstrated that the loss of Toll-like receptor 2 too is able to accelerate leukemic progression in NUP98-HOXD13 mouse model [166]. During leukemic progression, the presence of the fusion transcript NUP98-HOXD13 is associated with other hematopoietic disorders, such as myelodysplastic syndrome (MDS), chronic myeloid leukemia and blast crisis. Transgenic mice, expressing the fusion transcript, develop MDS; more than half of them progress to acute leukemia or display megakaryocytic differentiation and increase bone marrow apoptosis [167,168]. Other molecular alterations are associated with NUP98-HOXD13 in AML transformation. FLT3/ITD mutation is involved with leukemic transformation: its insertion in murine model induces only an MDS phenotype. In contrast, the co-expression of FLT3/ITD and NUP98-HOXD13, in the same model, induces AML with 100% penetration and short latency [169]. Furthermore, the loss of p15lnk4b together with NUP98-HOXD13 trans-gene leads to the development of Myeloid neoplasia, AML, Myelo-proliferative disease and MDS [170]. NUP98-HOXD13 fusion gene occurs also in non-lymphocytic-leukemia and it is coupled with NRAS, KRAS and Cbl gene mutations, in transgenic mice model [171]. Moreover, NUP98-HOXD13 drives the loss of one or both p53 alleles strengthening MDS phenotype and accelerating acute myeloid leukemia development [172].
A large IHC study performed on 4000 normal and neoplastic tissue samples, which included 79 different tumor categories, has shown that the over-expression of HOXD13 is prevalent in cancer tissues, compared to non-neoplastic samples, particularly in breast, colon and salivary glands cancers. However, in several tumor types, such as pancreatic and gastric cancers, HOXD13 has displayed the opposite trend being strongly down-regulated, which suggests also for this gene a dual role during tumor evolution, as oncogene and tumor suppressor [29]. Deregulation of HOXD13 displays a prognostic value in breast cancer. HOXD13 down-regulation is significantly associated with tumor size, lymph-node metastases and a poorer overall survival [173]. HOXD13 expression has been further proved to be a useful diagnostic tool in BC in combination with magnetic resonance imaging (MRI) [174]. HOXD13 promoter is methylated in about 60% of BC and the methylation status strongly correlates with clinic-pathological characteristics and a poorer survival of BC patients [175]. Moreover, the detection of HOXD13 methylation status in circulating free DNA (cfDNA) from serum, has proved to be a useful tool for BC patients management [176]. HOXD13 methylation status is associated with lung adenocarcinoma prognosis, suggesting, also for this tumor, its diagnostic value [177]. A microarray analysis reveals that posterior HOXD genes are involved in bone formation and hyper-expressed in primary Ewing sarcoma (ES). While posterior HOXD genes (from HOXD13 to HOXD9) promote chondrogenic differentiation and enhance bone-associated gene expression, HOXD11 and HOXD13 are specifically involved in cell growth and invasion of Ewing sarcoma. Their knockdown significantly suppresses lung metastases in mice models, suggesting a role in the metastatic potential of ES cells [178]. A cytogenetic analysis reveals the frequent chromosome copy gain at chr. 2q24, a region centromeric to HOXD13, in hepatoblastoma patients. Moreover, 2q24 gain is an independent factor able to predict poor outcome, suggesting the presence in this chromosomal area of a tumor suppressor gene involved in hepatoblastoma evolution [179].
The posterior HOXD locus, in addition to being regulated in trans by lncRNA HOTAIR, is under the control of two more lncRNAs, named Hotdog (HOG) and Twin of Hotdog (TOG), which are located in a desert zone of the gene adjacent to HOXD13 gene [180]. HOG/TOG is critical in the regulation of HOXD genes during the caecum development [180] but there is no information on its role during tumor transformation and progression.

6. Conclusions

The aim of this review was to summarize and comment on the numerous experimental evidences on the fundamental role of the most posterior genes of the HOX gene network in cancer diseases. Although the whole HOX gene network acts in a coordinated manner in body plan organization during development and in the maintenance of the phenotypic identity in human adult tissues and organs, deregulation of HOX13 paralogues has been strongly related to severe alterations of body structures [30,160] and associated with different molecular pathways promoting tumor diseases [32,181,182] (Figure 1). Since deregulation of HOX13 paralogues genes has been widely associated with invasion/metastasis [19,88] and drug resistance processes [83], their role has been suggested both as tumor diagnostic markers and as prognostic-predictive markers.
Even if HOX genes came from a common ancestral HOX gene, some of them appear to play different roles during cancer development and progression. In most human cancers, HOXB13 would seem to behave as a tumor suppressor gene, while for the other HOX13 paralogues genes a role as oncogenes is mainly described. HOX proteins act as transcription factors, therefore this remarkable functional difference may be due to their ability to interact with different target genes modulating many molecular pathways involved in proliferation, migration, and invasion. Further functional studies should be carried out to better define their mechanisms of action in order to modulate both their activation and inhibition. Blocking of HOX proteins activity by interfering with their binding to PBX co-factor, has been able to reduce tumor cell growth and induce apoptosis, supporting the therapeutic potential of inhibiting HOX/PBX dimer formation in cancer [183,184,185].
The ability to detect HOX13 proteins, as well as lncRNAs with which they co-localize/interact, in biological fluids [186,187,188], combined with therapeutic strategies that interfere with their activity, would open a new scenario in the management of cancer patients.

Author Contributions

Conceptualization, M.C. and C.C.; writing—original draft preparation, M.C. and M.G.M.; writing—review and editing, M.C. and R.D.C.; Supervision, G.B. and M.C.

Funding

This research was funded by the Italian Ministry of Health.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gehring, W.; Hiromi, Y. Homeotic genes and the homeobox. Annu. Rev. Genet. 1986, 20, 147–173. [Google Scholar] [CrossRef]
  2. Krumlauf, R. Hox genes in vertebrate development. Cell 1994, 78, 191–201. [Google Scholar] [CrossRef]
  3. Apiou, F.; Flagiello, D.; Cillo, C.; Malfoy, B.; Poupon, M.F.; Dutrillaux, B. Fine mapping of human HOX gene clusters. Cytogenet Cell Genet. 1996, 73, 114–115. [Google Scholar] [CrossRef] [PubMed]
  4. Graham, A.; Papalopulu, N.; Krumlauf, R. The murine and Drosophila homeobox gene complexes have common features of organization and expression. Cell 1989, 57, 367–378. [Google Scholar] [CrossRef]
  5. Gehring, W.J.; Kloter, U.; Suga, H. Evolution of the Hox gene complex from an evolutionary ground state. Curr. Top. Dev. Biol. 2009, 88, 35–61. [Google Scholar] [PubMed]
  6. Schuettengruber, B.; Bourbon, H.M.; Di Croce, L.; Cavalli, G. Genome Regulation by Polycomb and Trithorax: 70 Years and Counting. Cell 2017, 171, 34–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Schuettengruber, B.; Chourrout, D.; Vervoort, M.; Leblanc, B.; Cavalli, G. Genome regulation by polycomb and trithorax proteins. Cell 2007, 128, 735–745. [Google Scholar] [CrossRef]
  8. Lander, E.S. The new genomics: Global views of biology. Science 1996, 274, 536–539. [Google Scholar] [CrossRef]
  9. Durston, A.; Wacker, S.; Bardine, N.; Jansen, H. Time space translation: A hox mechanism for vertebrate a-p patterning. Curr. Genom. 2012, 13, 300–307. [Google Scholar] [CrossRef]
  10. Durston, A.J. Global posterior prevalence is unique to vertebrates: A dance to the music of time? Dev. Dyn. 2012, 241, 1799–1807. [Google Scholar] [CrossRef] [Green Version]
  11. Luo, Z.; Rhie, S.K.; Farnham, P.J. The Enigmatic HOX Genes: Can We Crack Their Code? Cancers 2019, 11, 323. [Google Scholar] [CrossRef] [PubMed]
  12. Botti, G.; De Chiara, A.; Di Bonito, M.; Cerrone, M.; Malzone, M.G.; Collina, F.; Cantile, M. Noncoding RNAs within the HOX gene network in tumor pathogenesis and progression. J. Cell. Physiol. 2018, 234, 395–413. [Google Scholar] [CrossRef] [PubMed]
  13. Burgess, D.J. Non-coding RNA: HOTTIP goes the distance. Nat. Rev. Genet. 2011, 12, 300. [Google Scholar] [CrossRef] [PubMed]
  14. Quagliata, L.; Matter, M.S.; Piscuoglio, S.; Arabi, L.; Ruiz, C.; Procino, A.; Kovac, M. Long noncoding RNA HOTTIP/HOXA13 expression is associated with disease progression and predicts outcome in hepatocellular carcinoma patients. Hepatology 2014, 59, 911–923. [Google Scholar] [CrossRef]
  15. Woo, C.J.; Kingston, R.E. HOTAIR lifts noncoding RNAs to new levels. Cell 2007, 129, 1257–1259. [Google Scholar] [CrossRef] [PubMed]
  16. Cillo, C.; Faiella, A.; Cantile, M.; Boncinelli, E. Homeobox genes and cancer. Exp. Cell Res. 1999, 248, 1–9. [Google Scholar] [CrossRef]
  17. Cillo, C.; Cantile, M.; Faiella, A.; Boncinelli, E. Homeobox genes in normal and malignant cells. J. Cell. Physiol. 2001, 188, 161–169. [Google Scholar] [CrossRef] [Green Version]
  18. Lv, X.; Li, L.; Lv, L.; Qu, X.; Li, K.; Deng, X.; Cheng, L.; He, H.; Dong, L. HOXD9 promotes epithelial–mesenchymal transition and cancer metastasis by ZEB1 regulation in hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 2015, 34, 133. [Google Scholar] [CrossRef] [Green Version]
  19. Cantile, M.; Scognamiglio, G.; Anniciello, A.; Farina, M.; Gentilcore, G.; Santonastaso, C.; Fulciniti, F.; Cillo, C.; Franco, R.; Ascierto, P.A.; et al. Increased HOX C13 expression in metastatic melanoma progression. J. Transl. Med. 2012, 10, 91. [Google Scholar] [CrossRef]
  20. Haria, D.; Naora, H. Homeobox Gene Deregulation: Impact on the Hallmarks of Cancer. Cancer Hallm. 2013, 1, 67–76. [Google Scholar] [CrossRef] [Green Version]
  21. Procino, A.; Cillo, C. The HOX genes network in metabolic diseases. Cell Biol. Int. 2013, 37, 1145–1148. [Google Scholar] [CrossRef]
  22. Sheth, R.; Barozzi, I.; Langlais, D.; Osterwalder, M.; Nemec, S.; Carlson, H.L.; Stadler, H.S.; Visel, A.; Drouin, J.; Kmita, M. Distal Limb Patterning Requires Modulation of cis-Regulatory Activities by HOX13. Cell Rep. 2016, 17, 2913–2926. [Google Scholar] [CrossRef]
  23. Beck, F. Homeobox genes in gut development. Gut 2002, 51, 450–454. [Google Scholar] [CrossRef] [Green Version]
  24. de Santa Barbara, P.; Roberts, D.J. Tail gut endoderm and gut/genitourinary/tail development: A new tissue-specific role for Hoxa13. Development 2002, 129, 551–561. [Google Scholar] [PubMed]
  25. Scott, V.; Morgan, E.A.; Stadler, H.S. Genitourinary functions of Hoxa13 and Hoxd13. J. Biochem. 2005, 137, 671–676. [Google Scholar] [CrossRef]
  26. Javed, S.; Langley, S.E. Importance of HOX genes in normal prostate gland formation, prostate cancer development and its early detection. BJU Int. 2014, 113, 535–540. [Google Scholar] [CrossRef] [PubMed]
  27. Aquino, G.; Franco, R.; Sabatino, R.; Mantia, E.L.; Scognamiglio, G.; Collina, F.; Longo, F.; Ionna, F.; Losito, N.S.; Liguori, G.; et al. Deregulation of paralogous 13 HOX genes in oral squamous cell carcinoma. Am. J. Cancer Res. 2015, 5, 3042–3055. [Google Scholar]
  28. Cantile, M.; Scognamiglio, G.; La Sala, L.; La Mantia, E.; Scaramuzza, V.; Valentino, E.; Tatangelo, F.; Losito, S.; Pezzullo, L.; Chiofalo, M.G.; et al. Aberrant expression of posterior HOX genes in well differentiated histotypes of thyroid cancers. Int. J. Mol. Sci. 2013, 14, 21727–21740. [Google Scholar] [CrossRef]
  29. Cantile, M.; Franco, R.; Tschan, A.; Baumhoer, D.; Zlobec, I.; Schiavo, G.; Forte, I.; Bihl, M.; Liguori, G.; Botti, G.; et al. HOX D13 expression across 79 tumor tissue types. Int. J. Cancer 2009, 125, 1532–1541. [Google Scholar] [CrossRef] [Green Version]
  30. Roux, M.; Bouchard, M.; Kmita, M. Multifaceted Hoxa13 function in urogenital development underlies the Hand-Foot-Genital Syndrome. Hum. Mol. Genet. 2019, 28, 1671–1681. [Google Scholar] [CrossRef]
  31. Mortlock, D.P.; Innis, J.W. Mutation of HOXA13 in hand-foot-genital syndrome. Nat. Genet. 1997, 15, 179–180. [Google Scholar] [CrossRef]
  32. Wen, Y.; Shu, F.; Chen, Y.; Chen, Y.; Lan, Y.; Duan, X.; Zhao, S.C.; Zeng, G. The prognostic value of HOXA13 in solid tumors: A meta-analysis. Clin. Chim. Acta 2018, 483, 64–68. [Google Scholar] [CrossRef]
  33. Hu, H.; Chen, Y.; Cheng, S.; Li, G.; Zhang, Z. Dysregulated expression of homeobox gene HOXA13 is correlated with the poor prognosis in bladder cancer. Wien. Klin. Wochenschr. 2017, 129, 391–397. [Google Scholar] [CrossRef]
  34. Guo, B.; Che, T.; Shi, B.; Guo, L.; Yin, Y.; Li, L.; Wang, J.; Yan, D.; Chen, Y. Screening and identification of specific markers for bladder transitional cell carcinoma from urine urothelial cells with suppressive subtractive hybridization and cDNA microarray. Can. Urol. Assoc. J. 2011, 5, 129–137. [Google Scholar] [CrossRef]
  35. Guo, B.; Che, T.; Shi, B.; Guo, L.; Zhang, Z.; Li, L.; Cai, C.; Chen, Y. Interaction network analysis of differentially expressed genes and screening of cancer marker in the urine of patients with invasive bladder cancer. Int. J. Clin. Exp. Med. 2015, 8, 3619–3628. [Google Scholar]
  36. Dong, Y.; Cai, Y.; Liu, B.; Jiao, X.; Li, Z.T.; Guo, D.Y.; Li, X.W.; Wang, Y.J.; Yang, D.K. HOXA13 is associated with unfavorable survival and acts as a novel oncogene in prostate carcinoma. Future Oncol. 2017, 13, 1505–1516. [Google Scholar] [CrossRef]
  37. Taketani, T.; Taki, T.; Ono, R.; Kobayashi, Y.; Ida, K.; Hayashi, Y. The chromosome translocation t(7;11)(p15;p15) in acute myeloid leukemia results in fusion of the NUP98 gene with a HOXA cluster gene, HOXA13, but not HOXA9. Genes Chromosomes Cancer 2002, 34, 437–443. [Google Scholar] [CrossRef]
  38. Fujino, T.; Suzuki, A.; Ito, Y.; Ohyashiki, K.; Hatano, Y.; Miura, I.; Nakamura, T. Single-translocation and double-chimeric transcripts: Detection of NUP98-HOXA9 in myeloid leukemias with HOXA11 or HOXA13 breaks of the chromosomal translocation t(7;11)(p15;p15). Blood 2002, 99, 1428–1433. [Google Scholar] [CrossRef]
  39. Su, X.; Drabkin, H.; Clappier, E.; Morgado, E.; Busson, M.; Romana, S.; Soulier, J.; Berger, R.; Bernard, O.A.; Lavau, C. Transforming potential of the T-cell acute lymphoblastic leukemia-associated homeobox genes HOXA13, TLX1, and TLX3. Genes Chromosomes Cancer 2006, 45, 846–855. [Google Scholar] [CrossRef]
  40. Han, Y.; Tu, W.W.; Wen, Y.G.; Li, D.P.; Qiu, G.Q.; Tang, H.M.; Peng, Z.H.; Zhou, C.Z. Identification and validation that up-expression of HOXA13 is a novel independent prognostic marker of a worse outcome in gastric cancer based on immunohistochemistry. Med. Oncol. 2013, 30, 564. [Google Scholar] [CrossRef]
  41. Yang, Y.C.; Wang, S.W.; Wu, I.C.; Chang, C.C.; Huang, Y.L.; Lee, O.K.; Chang, J.G.; Chen, A.; Kuo, F.C.; Wang, W.M.; et al. A tumorigenic homeobox (HOX) gene expressing human gastric cell line derived from putative gastric stem cell. Eur. J. Gastroenterol. Hepatol. 2009, 21, 1016–1023. [Google Scholar] [CrossRef]
  42. He, Y.X.; Song, X.H.; Zhao, Z.Y.; Zhao, H. HOXA13 upregulation in gastric cancer is associated with enhanced cancer cell invasion and epithelial-to-mesenchymal transition. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 258–265. [Google Scholar]
  43. Qu, L.P.; Zhong, Y.M.; Zheng, Z.; Zhao, R.X. CDH17 is a downstream effector of HOXA13 in modulating the Wnt/β-catenin signaling pathway in gastric cancer. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 1234–1241. [Google Scholar]
  44. Han, Y.; Song, C.; Wang, J.; Tang, H.; Peng, Z.; Lu, S. HOXA13 contributes to gastric carcinogenesis through DHRS2 interacting with MDM2 and confers 5-FU resistance by a p53-dependent pathway. Mol. Carcinog. 2018, 57, 722–734. [Google Scholar] [CrossRef]
  45. Yan, W.P.; Shen, L.Y.; Gu, Z.D.; Chen, K.N. Up-regulation of HOXA13 in esophageal squamous cell carcinoma of stage IIa and its effect on the prognosis. Chin. J. Gastrointest. Surg. 2009, 12, 20–23. [Google Scholar]
  46. Gu, Z.D.; Shen, L.Y.; Wang, H.; Chen, X.M.; Li, Y.; Ning, T.; Chen, K.N. HOXA13 promotes cancer cell growth and predicts poor survival of patients with esophageal squamous cell carcinoma. Cancer Res. 2009, 69, 4969–4973. [Google Scholar] [CrossRef]
  47. Ma, R.L.; Shen, L.Y.; Chen, K.N. Coexpression of ANXA2, SOD2 and HOXA13 predicts poor prognosis of esophageal squamous cell carcinoma. Oncol. Rep. 2014, 31, 2157–2164. [Google Scholar] [CrossRef]
  48. Shi, Q.; Shen, L.; Dong, B.; Fu, H.; Kang, X.; Dai, L.; Yang, Y.; Yan, W.; Chen, K.N. Downregulation of HOXA13 sensitizes human esophageal squamous cell carcinoma to chemotherapy. Thorac. Cancer 2018, 9, 836–846. [Google Scholar] [CrossRef] [Green Version]
  49. Huang, T.; Chesnokov, V.; Yokoyama, K.K.; Carr, B.I.; Itakura, K. Expression of the Hoxa-13 Gene Correlates to Hepatitis B and C Virus Associated HCC. Biochem Biophys Res. Commun. 2001, 281, 1041–1044. [Google Scholar] [CrossRef]
  50. Cillo, C.; Schiavo, G.; Cantile, M.; Bihl, M.P.; Sorrentino, P.; Carafa, V.; D’Armiento, M.; Roncalli, M.; Sansano, S.; Vecchione, R.; et al. The HOX gene network in hepatocellular carcinoma. Int. J. Cancer 2011, 129, 2577–2587. [Google Scholar] [CrossRef] [Green Version]
  51. Pan, T.T.; Jia, W.D.; Yao, Q.Y.; Sun, Q.K.; Ren, W.H.; Huang, M.; Ma, J.; Li, J.S.; Ma, J.L.; Yu, J.H.; et al. Overexpression of HOXA13 as a potential marker for diagnosis and poor prognosis of hepatocellular carcinoma. Tohoku J. Exp. Med. 2014, 234, 209–219. [Google Scholar] [CrossRef] [PubMed]
  52. Quagliata, L.; Quintavalle, C.; Lanzafame, M.; Matter, M.S.; Novello, C.; di Tommaso, L.; Pressiani, T.; Rimassa, L.; Tornillo, L.; Roncalli, M.; et al. High expression of HOXA13 correlates with poorly differentiated hepatocellular carcinomas and modulates sorafenib response in in vitro models. Lab. Investig. 2018, 98, 95–105. [Google Scholar] [CrossRef] [PubMed]
  53. Deng, Y.; He, R.; Zhang, R.; Gan, B.; Zhang, Y.; Chen, G.; Hu, X. The expression of HOXA13 in lung adenocarcinoma and its clinical significance: A study based on The Cancer Genome Atlas, Oncomine and reverse transcription-quantitative polymerase chain reaction. Oncol. Lett. 2018, 15, 8556–8572. [Google Scholar] [CrossRef] [Green Version]
  54. Kang, J.U. Characterization of amplification patterns and target genes on the short arm of chromosome 7 in early-stage lung adenocarcinoma. Mol. Med. Rep. 2013, 8, 1373–1378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Kelly, Z.; Moller-Levet, C.; McGrath, S.; Butler-Manuel, S.; Kavitha Madhuri, T.; Kierzek, A.M.; Pandha, H.; Morgan, R.; Michael, A. The prognostic significance of specific HOX gene expression patterns in ovarian cancer. Int. J. Cancer 2016, 139, 1608–1617. [Google Scholar] [CrossRef] [PubMed]
  56. Duan, R.; Han, L.; Wang, Q.; Wei, J.; Chen, L.; Zhang, J.; Kang, C.; Wang, L. HOXA13 is a potential GBM diagnostic marker and promotes glioma invasion by activating the Wnt and TGF-β pathways. Oncotarget 2015, 6, 27778–27793. [Google Scholar] [CrossRef] [Green Version]
  57. Wang, K.C.; Yang, Y.W.; Liu, B.; Sanyal, A.; Corces-Zimmerman, R.; Chen, Y.; Lajoie, B.R.; Protacio, A.; Flynn, R.A.; Gupta, R.A.; et al. A long noncoding RNA maintains active chromatin to coordinate homeotic gene expression. Nature 2011, 472, 120–124. [Google Scholar] [CrossRef]
  58. Fu, Z.; Chen, C.; Zhou, Q.; Wang, Y.; Zhao, Y.; Zhao, X.; Li, W.; Zheng, S.; Ye, H.; Wang, L.; et al. LncRNA HOTTIP modulates cancer stem cell properties in human pancreatic cancer by regulating HOXA9. Cancer Lett. 2017, 410, 68–81. [Google Scholar] [CrossRef]
  59. Chang, S.; Liu, J.; Guo, S.; He, S.; Qiu, G.; Lu, J.; Wang, J.; Fan, L.; Zhao, W.; Che, X. HOTTIP and HOXA13 are oncogenes associated with gastric cancer progression. Oncol. Rep. 2016, 35, 3577–3585. [Google Scholar] [CrossRef]
  60. Wang, S.S.; Wuputra, K.; Liu, C.J.; Lin, Y.C.; Chen, Y.T.; Chai, C.Y.; Lin, C.S.; Kuo, K.K.; Tsai, M.H.; Wang, S.W.; et al. Oncogenic function of the homeobox A13-long noncoding RNA HOTTIP-insulin growth factor-binding protein 3 axis in human gastric cancer. Oncotarget 2016, 7, 36049–36064. [Google Scholar] [CrossRef] [Green Version]
  61. Luo, Z.; Rhie, S.K.; Lay, F.D.; Farnham, P.J. A Prostate Cancer Risk Element Functions as a Repressive Loop that Regulates HOXA13. Cell Rep. 2017, 21, 1411–1417. [Google Scholar] [CrossRef] [PubMed]
  62. Zhang, S.R.; Yang, J.K.; Xie, J.K.; Zhao, L.C. Long noncoding RNA HOTTIP contributes to the progression of prostate cancer by regulating HOXA13. Cell Mol. Biol. 2016, 62, 84–88. [Google Scholar]
  63. Lin, C.; Wang, Y.; Wang, Y.; Zhang, S.; Yu, L.; Guo, C.; Xu, H. Transcriptional and posttranscriptional regulation of HOXA13 by lncRNA HOTTIP facilitates tumorigenesis and metastasis in esophageal squamous carcinoma cells. Oncogene 2017, 36, 5392–5406. [Google Scholar] [CrossRef]
  64. Sang, Y.; Zhou, F.; Wang, D.; Bi, X.; Liu, X.; Hao, Z.; Li, Q.; Zhang, W. Up-regulation of long non-coding HOTTIP functions as an oncogene by regulating HOXA13 in non-small cell lung cancer. Am. J. Transl. Res. 2016, 8, 2022–2032. [Google Scholar]
  65. Stelnicki, E.J.; Arbeit, J.; Cass, D.L.; Saner, C.; Harrison, M.; Largman, C. Modulation of the human homeobox genes PRX-2 and HOXB13 in scarless fetal wounds. J. Investig. Dermatol. 1998, 111, 57–63. [Google Scholar] [CrossRef]
  66. Kömüves, L.G.; Ma, X.K.; Stelnicki, E.; Rozenfeld, S.; Oda, Y.; Largman, C. HOXB13 homeodomain protein is cytoplasmic throughout fetal skin development. Dev. Dyn. 2003, 227, 192–202. [Google Scholar] [CrossRef] [Green Version]
  67. Economides, K.D.; Capecchi, M.R. Hoxb13 is required for normal differentiation and secretory function of the ventral prostate. Development 2003, 130, 2061–2069. [Google Scholar] [CrossRef] [Green Version]
  68. Norris, J.D.; Chang, C.Y.; Wittmann, B.M.; Kunder, R.S.; Cui, H.; Fan, D.; Joseph, J.D.; McDonnell, D.P. The homeodomain protein HOXB13 regulates the cellular response to androgens. Mol. Cell 2009, 36, 405–416. [Google Scholar] [CrossRef] [PubMed]
  69. Cantile, M.; Pettinato, G.; Procino, A.; Feliciello, I.; Cindolo, L.; Cillo, C. In vivo expression of the whole HOX gene network in human breast cancer. Eur. J. Cancer 2003, 39, 257–264. [Google Scholar] [CrossRef]
  70. Ma, X.J.; Wang, Z.; Ryan, P.D.; Isakoff, S.J.; Barmettler, A.; Fuller, A.; Muir, B.; Mohapatra, G.; Salunga, R.; Tuggle, J.T.; et al. A two-gene expression ratio predicts clinical outcome in breast cancer patients treated with tamoxifen. Cancer Cell 2004, 5, 607–616. [Google Scholar] [CrossRef] [Green Version]
  71. Goetz, M.P.; Suman, V.J.; Ingle, J.N.; Nibbe, A.M.; Visscher, D.W.; Reynolds, C.A.; Lingle, W.L.; Erlander, M.; Ma, X.J.; Sgroi, D.C.; et al. A two-gene expression ratio of homeobox 13 and interleukin-17B receptor for prediction of recurrence and survival in women receiving adjuvant tamoxifen. Clin. Cancer Res. 2006, 12, 2080–2087. [Google Scholar] [CrossRef]
  72. Wang, Z.; Dahiya, S.; Provencher, H.; Muir, B.; Carney, E.; Coser, K.; Shioda, T.; Ma, X.J.; Sgroi, D.C. The prognostic biomarkers HOXB13, IL17BR, and CHDH are regulated by estrogen in breast cancer. Clin. Cancer Res. 2007, 13, 6327–6334. [Google Scholar] [CrossRef]
  73. Jansen, M.P.; Sieuwerts, A.M.; Look, M.P.; Ritstier, K.; Meijer-van Gelder, M.E.; van Staveren, I.L.; Klijn, J.G.; Foekens, J.A.; Berns, E.M. HOXB13-to-IL17BR expression ratio is related with tumor aggressiveness and response to tamoxifen of recurrent breast cancer: A retrospective study. J. Clin. Oncol. 2007, 25, 662–668. [Google Scholar] [CrossRef]
  74. Jerevall, P.L.; Jansson, A.; Fornander, T.; Skoog, L.; Nordenskjöld, B.; Stål, O. Predictive relevance of HOXB13 protein expression for tamoxifen benefit in breast cancer. Breast Cancer Res. 2010, 12, R53. [Google Scholar] [CrossRef]
  75. Ma, X.J.; Hilsenbeck, S.G.; Wang, W.; Ding, L.; Sgroi, D.C.; Bender, R.A.; Osborne, C.K.; Allred, D.C.; Erlander, M.G. The HOXB13:IL17BR expression index is a prognostic factor in early-stage breast cancer. J. Clin. Oncol. 2006, 24, 4611–4619. [Google Scholar] [CrossRef]
  76. Jerevall, P.L.; Brommesson, S.; Strand, C.; Gruvberger-Saal, S.; Malmström, P.; Nordenskjöld, B.; Wingren, S.; Söderkvist, P.; Fernö, M.; Stål, O. Exploring the two-gene ratio in breast cancer--independent roles for HOXB13 and IL17BR in prediction of clinical. Breast Cancer Res. Treat. 2008, 107, 225–234. [Google Scholar] [CrossRef]
  77. Zhao, L.; Zhu, S.; Gao, Y.; Wang, Y. Two-gene expression ratio as predictor for breast cancer treated with tamoxifen: Evidence from meta-analysis. Tumour Biol. 2014, 35, 3113–3117. [Google Scholar] [CrossRef]
  78. Ma, X.J.; Salunga, R.; Dahiya, S.; Wang, W.; Carney, E.; Durbecq, V.; Harris, A.; Goss, P.; Sotiriou, C.; Erlander, M.; et al. A five-gene molecular grade index and HOXB13:IL17BR are complementary prognostic factors in early stage breast cancer. Clin. Cancer Res. 2008, 14, 2601–2608. [Google Scholar] [CrossRef]
  79. Habel, L.A.; Sakoda, L.C.; Achacoso, N.; Ma, X.J.; Erlander, M.G.; Sgroi, D.C.; Fehrenbacher, L.; Greenberg, D.; Quesenberry, C.P., Jr. HOXB13:IL17BR and molecular grade index and risk of breast cancer death among patients with lymph node-negative invasive disease. Breast Cancer Res. 2013, 15, R24. [Google Scholar] [CrossRef]
  80. Sgroi, D.C.; Chapman, J.A.; Badovinac-Crnjevic, T.; Zarella, E.; Binns, S.; Zhang, Y.; Schnabel, C.A.; Erlander, M.G.; Pritchard, K.I.; Han, L.; et al. Assessment of the prognostic and predictive utility of the Breast Cancer Index (BCI): An NCIC CTG MA.14 study. Breast Cancer Res. 2016, 18, 1. [Google Scholar] [CrossRef]
  81. Rodriguez, B.A.; Cheng, A.S.; Yan, P.S.; Potter, D.; Agosto-Perez, F.J.; Shapiro, C.L.; Huang, T.H. Epigenetic repression of the estrogen-regulated Homeobox B13 gene in breast cancer. Carcinogenesis 2008, 29, 1459–1465. [Google Scholar] [CrossRef] [Green Version]
  82. Shah, N.; Jin, K.; Cruz, L.A.; Park, S.; Sadik, H.; Cho, S.; Goswami, C.P.; Nakshatri, H.; Gupta, R.; Chang, H.Y.; et al. HOXB13 mediates tamoxifen resistance and invasiveness in human breast cancer by suppressing ERα and inducing IL-6 expression. Cancer Res. 2013, 73, 5449–5458. [Google Scholar] [CrossRef]
  83. Liu, B.; Wang, T.; Wang, H.; Zhang, L.; Xu, F.; Fang, R.; Li, L.; Cai, X.; Wu, Y.; Zhang, W.; et al. Oncoprotein HBXIP enhances HOXB13 acetylation and co-activates HOXB13 to confer tamoxifen resistance in breast cancer. J. Hematol. Oncol. 2018, 11, 26. [Google Scholar] [CrossRef] [Green Version]
  84. Sreenath, T.; Orosz, A.; Fujita, K.; Bieberich, C.J. Androgen-independent expression of hoxb-13 in the mouse prostate. Prostate 1999, 41, 203–207. [Google Scholar] [CrossRef]
  85. Jung, C.; Kim, R.S.; Lee, S.J.; Wang, C.; Jeng, M.H. HOXB13 homeodomain protein suppresses the growth of prostate cancer cells by the negative regulation of T-cell factor 4. Cancer Res. 2004, 64, 3046–3051. [Google Scholar] [CrossRef]
  86. Jung, C.; Kim, R.S.; Zhang, H.J.; Lee, S.J.; Jeng, M.H. HOXB13 induces growth suppression of prostate cancer cells as a repressor of hormone-activated androgen receptor signaling. Cancer Res. 2004, 64, 9185–9192. [Google Scholar] [CrossRef]
  87. Kim, Y.R.; Oh, K.J.; Park, R.Y.; Xuan, N.T.; Kang, T.W.; Kwon, D.D.; Choi, C.; Kim, M.S.; Nam, K.I.; Ahn, K.Y.; et al. HOXB13 promotes androgen independent growth of LNCaP prostate cancer cells by the activation of E2F signaling. Mol. Cancer 2010, 9, 124. [Google Scholar] [CrossRef]
  88. Jeong, T.O.; Oh, K.J.; Nguyen, X.; Thi, N.; Kim, Y.R.; Kim, M.S.; Lee, S.D.; Ryu, S.B.; Jung, C. Evaluation of HOXB13 as a molecular marker of recurrent prostate cancer. Mol. Med. Rep. 2012, 5, 901–904. [Google Scholar] [CrossRef] [Green Version]
  89. Varinot, J.; Cussenot, O.; Roupret, M.; Conort, P.; Bitker, M.O.; Chartier-Kastler, E.; Cheng, L.; Compérat, E. HOXB13 is a sensitive and specific marker of prostate cells, useful in distinguishing between carcinomas of prostatic and urothelial origin. Virchows Arch. 2013, 463, 803–809. [Google Scholar] [CrossRef]
  90. Barresi, V.; Ieni, A.; Cardia, R.; Licata, L.; Vitarelli, E.; Reggiani Bonetti, L.; Tuccari, G. HOXB13 as an immunohistochemical marker of prostatic origin in metastatic tumors. APMIS 2016, 124, 188–193. [Google Scholar] [CrossRef]
  91. Larnaudie, L.; Compérat, E.; Conort, P.; Varinot, J. HOXB13 a useful marker in pleomorphic giant cell adenocarcinoma of the prostate: A case report and review of the literature. Virchows Arch. 2017, 471, 133–136. [Google Scholar] [CrossRef]
  92. Zabalza, C.V.; Adam, M.; Burdelski, C.; Wilczak, W.; Wittmer, C.; Kraft, S.; Krech, T.; Steurer, S.; Koop, C.; Hube-Magg, C.; et al. HOXB13 overexpression is an independent predictor of early PSA recurrence in prostate cancer treated by radical prostatectomy. Oncotarget 2015, 6, 12822–12834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Kim, I.J.; Kang, T.W.; Jeong, T.; Kim, Y.R.; Jung, C. HOXB13 regulates the prostate-derived Ets factor: Implications for prostate cancer cell invasion. Int. J. Oncol. 2014, 45, 869–876. [Google Scholar] [CrossRef]
  94. Kim, Y.R.; Kang, T.W.; To, P.K.; Xuan Nguyen, N.T.; Cho, Y.S.; Jung, C.; Kim, M.S. HOXB13-mediated suppression of p21WAF1/CIP1 regulates JNK/c-Jun signaling in prostate cancer cells. Oncol. Rep. 2016, 35, 2011–2016. [Google Scholar] [CrossRef]
  95. Kim, Y.R.; Kim, I.J.; Kang, T.W.; Choi, C.; Kim, K.K.; Kim, M.S.; Nam, K.I.; Jung, C. HOXB13 downregulates intracellular zinc and increases NF-κB signaling to promote prostate cancer metastasis. Oncogene 2014, 33, 4558–4567. [Google Scholar] [CrossRef]
  96. Johng, D.; Torga, G.; Ewing, C.M.; Jin, K.; Norris, J.D.; McDonnell, D.P.; Isaacs, W.B. HOXB13 interaction with MEIS1 modifies proliferation and gene expression in prostate cancer. Prostate 2019, 79, 414–424. [Google Scholar] [CrossRef] [PubMed]
  97. Ewing, C.M.; Ray, A.M.; Lange, E.M.; Zuhlke, K.A.; Robbins, C.M.; Tembe, W.D.; Wiley, K.E.; Isaacs, S.D.; Johng, D.; Wang, Y.; et al. Germline mutations in HOXB13 and prostate-cancer risk. N. Engl. J. Med. 2012, 366, 141–149. [Google Scholar] [CrossRef] [PubMed]
  98. Breyer, J.P.; Avritt, T.G.; McReynolds, K.M.; Dupont, W.D.; Smith, J.R. Confirmation of the HOXB13 G84E germline mutation in familial prostate cancer. Cancer Epidemiol. Biomark. Prev. 2012, 21, 1348–1353. [Google Scholar] [CrossRef]
  99. Karlsson, R.; Aly, M.; Clements, M.; Zheng, L.; Adolfsson, J.; Xu, J.; Grönberg, H.; Wiklund, F. A population-based assessment of germline HOXB13 G84E mutation and prostate cancer risk. Eur. Urol. 2014, 65, 169–176. [Google Scholar] [CrossRef]
  100. Lin, X.; Qu, L.; Chen, Z.; Xu, C.; Ye, D.; Shao, Q.; Wang, X.; Qi, J.; Chen, Z.; Zhou, F.; et al. A novel germline mutation in HOXB13 is associated with prostate cancer risk in Chinese men. Prostate 2013, 73, 169–175. [Google Scholar] [CrossRef]
  101. Akbari, M.R.; Trachtenberg, J.; Lee, J.; Tam, S.; Bristow, R.; Loblaw, A.; Narod, S.A.; Nam, R.K. Association between germline HOXB13 G84E mutation and risk of prostate cancer. J. Natl. Cancer Inst. 2012, 104, 1260–1262. [Google Scholar] [CrossRef]
  102. Xu, J.; Lange, E.M.; Lu, L.; Zheng, S.L.; Wang, Z.; Thibodeau, S.N.; Cannon-Albright, L.A.; Teerlink, C.C.; Camp, N.J.; Johnson, A.M.; et al. International Consortium for Prostate Cancer Genetics.HOXB13 is a susceptibility gene for prostate cancer: Results from the International Consortium for Prostate Cancer Genetics (ICPCG). Hum. Genet. 2013, 132, 5–14. [Google Scholar] [CrossRef]
  103. Stott-Miller, M.; Karyadi, D.M.; Smith, T.; Kwon, E.M.; Kolb, S.; Stanford, J.L.; Ostrander, E.A. HOXB13 mutations in a population-based, case-control study of prostate cancer. Prostate 2013, 73, 634–641. [Google Scholar] [CrossRef]
  104. Smith, S.C.; Palanisamy, N.; Zuhlke, K.A.; Johnson, A.M.; Siddiqui, J.; Chinnaiyan, A.M.; Kunju, L.P.; Cooney, K.A.; Tomlins, S.A. HOXB13 G84E-related familial prostate cancers: A clinical, histologic, and molecular survey. Am. J. Surg. Pathol. 2014, 38, 615–626. [Google Scholar] [CrossRef]
  105. Maia, S.; Cardoso, M.; Pinto, P.; Pinheiro, M.; Santos, C.; Peixoto, A.; Bento, M.J.; Oliveira, J.; Henrique, R.; Jerónimo, C.; et al. Identification of Two Novel HOXB13 Germline Mutations in Portuguese Prostate Cancer Patients. PLoS ONE 2015, 10, e0132728. [Google Scholar] [CrossRef] [PubMed]
  106. Jung, C.; Kim, R.S.; Zhang, H.; Lee, S.J.; Sheng, H.; Loehrer, P.J.; Gardner, T.A.; Jeng, M.H.; Kao, C. HOXB13 is downregulated in colorectal cancer to confer TCF4-mediated transactivation. Br. J. Cancer 2005, 92, 2233–2239. [Google Scholar] [CrossRef] [Green Version]
  107. Ghoshal, K.; Motiwala, T.; Claus, R.; Yan, P.; Kutay, H.; Datta, J.; Majumder, S.; Bai, S.; Majumder, A.; Huang, T.; et al. HOXB13, a target of DNMT3B, is methylated at an upstream CpG island, and functions as a tumor suppressor in primary colorectal tumors. PLoS ONE 2010, 5, e10338. [Google Scholar] [CrossRef] [PubMed]
  108. Tatangelo, F.; Di Mauro, A.; Scognamiglio, G.; Aquino, G.; Lettiero, A.; Delrio, P.; Avallone, A.; Cantile, M.; Botti, G. Posterior HOX genes and HOTAIR expression in the proximal and distal colon cancer pathogenesis. J. Transl. Med. 2018, 16, 350. [Google Scholar] [CrossRef]
  109. Akbari, M.R.; Anderson, L.N.; Buchanan, D.D.; Clendenning, M.; Jenkins, M.A.; Win, A.K.; Hopper, J.L.; Giles, G.G.; Nam, R.; Narod, S.; et al. Germline HOXB13 p. Gly84Glu mutation and risk of colorectal cancer. Cancer Epidemiol. 2013, 37, 424–427. [Google Scholar] [CrossRef]
  110. Yuan, H.; Kajiyama, H.; Ito, S.; Chen, D.; Shibata, K.; Hamaguchi, M.; Kikkawa, F.; Senga, T. HOXB13 promotes ovarian cancer progression, HOXB13 and ALX4 induce SLUG expression for the promotion of EMT and cell invasion in ovarian cancer cells. Oncotarget 2015, 6, 13359–13370. [Google Scholar] [CrossRef] [PubMed]
  111. Saha, S.S.; Chowdhury, R.R.; Mondal, N.R.; Roy, S.; Sengupta, S. Expression signatures of HOX cluster genes in cervical cancer pathogenesis: Impact of human papillomavirus type 16 oncoprotein E7. Oncotarget 2017, 8, 36591–36602. [Google Scholar] [CrossRef] [Green Version]
  112. Zhao, Y.; Yamashita, T.; Ishikawa, M. Regulation of tumor invasion by HOXB13 gene overexpressed in human endometrial cancer. Oncol. Rep. 2005, 13, 721–726. [Google Scholar] [CrossRef]
  113. Marra, L.; Cantile, M.; Scognamiglio, G.; Perdonà, S.; La Mantia, E.; Cerrone, M.; Gigantino, V.; Cillo, C.; Caraglia, M.; Pignata, S.; et al. Deregulation of HOX B13 expression in urinary bladder cancer progression. Curr. Med. Chem. 2013, 20, 833–839. [Google Scholar]
  114. Okuda, H.; Toyota, M.; Ishida, W.; Furihata, M.; Tsuchiya, M.; Kamada, M.; Tokino, T.; Shuin, T. Epigenetic inactivation of the candidate tumor suppressor gene HOXB13 in human renal cell carcinoma. Oncogene 2006, 25, 1733–1742. [Google Scholar] [CrossRef]
  115. Sui, B.Q.; Zhang, C.D.; Liu, J.C.; Wang, L.; Dai, D.Q. HOXB13 expression and promoter methylation as a candidate biomarker in gastric cancer. Oncol. Lett. 2018, 15, 8833–8840. [Google Scholar] [CrossRef] [Green Version]
  116. Zhu, J.Y.; Sun, Q.K.; Wang, W.; Jia, W.D. High-level expression of HOXB13 is closely associated with tumor angiogenesis and poor prognosis of hepatocellular carcinoma. Int. J. Clin. Exp. Pathol. 2014, 7, 2925–2933. [Google Scholar]
  117. Zhang, E.; Han, L.; Yin, D.; He, X.; Hong, L.; Si, X.; Qiu, M.; Xu, T.; De, W.; Xu, L.; et al. H3K27 acetylation activated-long non-coding RNA CCAT1 affects cell proliferation and migration by regulating SPRY4 and HOXB13 expression in esophageal squamous cell carcinoma. Nucleic Acids Res. 2017, 45, 3086–3101. [Google Scholar] [CrossRef]
  118. Cazal, C.; Sobral, A.P.; de Almeida, F.C.; das Graças Silva-Valenzuela, M.; Durazzo, M.D.; Nunes, F.D. The homeobox HOXB13 is expressed in human minor salivary gland. Oral Dis. 2006, 12, 424–427. [Google Scholar] [CrossRef]
  119. Xiong, Y.; Kuang, W.; Lu, S.; Guo, H.; Wu, M.; Ye, M.; Wu, L. Long noncoding RNA HOXB13-AS1 regulates HOXB13 gene methylation by interacting with EZH2 in glioma. Cancer Med. 2018, 7, 4718–4728. [Google Scholar] [CrossRef]
  120. Liu, X.F.; Olsson, P.; Wolfgang, C.D.; Bera, T.K.; Duray, P.; Lee, B.; Pastan, I. PRAC: A novel small nuclear protein that is specifically expressed in human prostate and colon. Prostate 2001, 47, 125–131. [Google Scholar] [CrossRef]
  121. Olsson, P.; Motegi, A.; Bera, T.K.; Lee, B.; Pastan, I. PRAC2: A new gene expressed in human prostate and prostate cancer. Prostate 2003, 56, 123–130. [Google Scholar] [CrossRef]
  122. Jiang, R.; Zhao, C.; Gao, B.; Shao, N.; Wang, S.; Song, W. IL-22 promotes the progression of breast cancer through regulating HOXB-AS5. Oncotarget 2017, 8, 103601–103612. [Google Scholar] [Green Version]
  123. Jonker, L.; Kist, R.; Aw, A.; Wappler, I.; Peters, H. Pax9 is required for filiform papilla development and suppresses skin-specific differentiation of the mammalian tongue epithelium. Mech. Dev. 2004, 121, 1313–1322. [Google Scholar] [CrossRef]
  124. Godwin, A.R.; Capecchi, M.R. Hoxc13 mutant mice lack external hair. Genes Dev. 1998, 12, 11–20. [Google Scholar] [CrossRef] [Green Version]
  125. Magli, M.C.; Barba, P.; Celetti, A.; De Vita, G.; Cillo, C.; Boncinelli, E. Coordinate regulation of HOX genes in human hematopoietic cells. Proc. Natl. Acad. Sci. USA 1991, 88, 6348–6352. [Google Scholar] [CrossRef]
  126. Jave-Suarez, L.F.; Winter, H.; Langbein, L.; Rogers, M.A.; Schweizer, J. HOXC13 is involved in the regulation of human hair keratin gene expression. J. Biol. Chem. 2002, 277, 3718–3726. [Google Scholar] [CrossRef]
  127. Jave-Suárez, L.F.; Schweizer, J. The HOXC13-controlled expression of early hair keratin genes in the human hair follicle does not involve TALE proteins MEIS and PREP as cofactors. Arch. Dermatol. Res. 2006, 297, 372–376. [Google Scholar] [CrossRef]
  128. Luan, L.; Shi, J.; Yu, Z.; Andl, T. The major miR-31 target genes STK40 and LATS2 and their implications in the regulation of keratinocyte growth and hair differentiation. Exp. Dermatol. 2017, 26, 497–504. [Google Scholar] [CrossRef] [PubMed]
  129. Cribier, B.; Peltre, B.; Grosshans, E.; Langbein, L.; Schweizer, J. On the regulation of hair keratin expression: Lessons from studies in pilomatricomas. J. Investig. Dermatol. 2004, 122, 1078–1083. [Google Scholar] [CrossRef] [PubMed]
  130. Cribier, B.; Worret, W.I.; Braun-Falco, M.; Peltre, B.; Langbein, L.; Schweizer, J. Expression patterns of hair and epithelial keratins and transcription factors HOXC13, LEF1, and beta-catenin in a malignant pilomatricoma: A histological and immunohistochemical study. J. Cutan Pathol. 2006, 33, 1–9. [Google Scholar] [CrossRef]
  131. Yoon, S.J.; LeBlanc-Straceski, J.; Ward, D.; Krauter, K.; Kucherlapati, R. Organization of the human keratin type II gene cluster at 12q13. Genomics 1994, 24, 502–508. [Google Scholar] [CrossRef]
  132. Godwin, A.R.; Capecchi, M.R. Hair defects in Hoxc13 mutant mice. J. Investig. Dermatol. Symp. Proc. 1999, 4, 244–247. [Google Scholar] [CrossRef]
  133. Kasiri, S.; Ansari, K.I.; Hussain, I.; Bhan, A.; Mandal, S.S. Antisense oligonucleotide mediated knockdown of HOXC13 affects cell growth and induces apoptosis in tumor cells and over expression of HOXC13 induces 3D-colony formation. RSC Adv. 2013, 3, 3260–3269. [Google Scholar] [CrossRef]
  134. Cillo, C.; Cantile, M.; Mortarini, R.; Barba, P.; Parmiani, G.; Anichini, A. Differential patterns of HOX gene expression are associated with specific integrin and ICAM profiles in clonal populations isolated from a single human melanoma metastasis. Int. J. Cancer 1996, 66, 692–697. [Google Scholar] [CrossRef]
  135. Maeda, K.; Hamada, J.; Takahashi, Y.; Tada, M.; Yamamoto, Y.; Sugihara, T.; Moriuchi, T. Altered expressions of HOX genes in human cutaneous malignant melanoma. Int. J. Cancer 2005, 114, 436–441. [Google Scholar] [CrossRef] [PubMed]
  136. Panagopoulos, I.; Isaksson, M.; Billström, R.; Strömbeck, B.; Mitelman, F.; Johansson, B. Fusion of the NUP98 gene and the homeobox gene HOXC13 in acute myeloid leukemia with t(11;12)(p15;q13). Genes Chromosomes Cancer 2003, 36, 107–112. [Google Scholar] [CrossRef] [PubMed]
  137. La Starza, R.; Trubia, M.; Crescenzi, B.; Matteucci, C.; Negrini, M.; Martelli, M.F.; Pelicci, P.G.; Mecucci, C. Human homeobox gene HOXC13 is the partner of NUP98 in adult acute myeloid leukemia with t(11;12)(p15;q13). Genes Chromosomes Cancer 2003, 36, 420–423. [Google Scholar] [CrossRef]
  138. Kobzev, Y.N.; Martinez-Climent, J.; Lee, S.; Chen, J.; Rowley, J.D. Analysis of translocations that involve the NUP98 gene in patients with 11p15 chromosomal rearrangements. Genes Chromosomes Cancer 2004, 41, 339–352. [Google Scholar] [CrossRef] [PubMed]
  139. Tosić, N.; Stojiljković, M.; Colović, N.; Colović, M.; Pavlović, S. Acute myeloid leukemia with NUP98-HOXC13 fusion and FLT3 internal tandem duplication mutation: Case report and literature review. Cancer Genet. Cytogenet. 2009, 193, 98–103. [Google Scholar] [CrossRef] [PubMed]
  140. La Starza, R.; Brandimarte, L.; Pierini, V.; Nofrini, V.; Gorello, P.; Crescenzi, B.; Berchicci, L.; Matteucci, C.; Romoli, S.; Beacci, D.; et al. A NUP98-positive acute myeloid leukemia with a t(11;12)(p15;q13) without HOXC cluster gene involvement. Cancer Genet. Cytogenet. 2009, 193, 109–111. [Google Scholar] [CrossRef]
  141. Yamada, T.; Shimizu, T.; Suzuki, M.; Kihara-Negishi, F.; Nanashima, N.; Sakurai, T.; Fan, Y.; Akita, M.; Oikawa, T.; Tsuchida, S. Interaction between the homeodomain protein HOXC13 and ETS family transcription factor PU.1 and its implication in the differentiation of murine erythroleukemia cells. Exp. Cell Res. 2008, 314, 847–858. [Google Scholar] [CrossRef]
  142. Zhong, M.; Wang, J.; Gong, Y.B.; Li, J.C.; Zhang, B.; Hou, L. Expression of HOXC13 in ameloblastoma. Chin. J. Stomatol. 2007, 42, 43–46. [Google Scholar]
  143. Schiavo, G.; D’Antò, V.; Cantile, M.; Procino, A.; Di Giovanni, S.; Valletta, R.; Terracciano, L.; Baumhoer, D.; Jundt, G.; Cillo, C. Deregulated HOX genes in ameloblastomas are located in physical contiguity to keratin genes. J. Cell Biochem. 2011, 112, 3206–3215. [Google Scholar] [CrossRef] [PubMed]
  144. Hong, Y.S.; Wang, J.; Liu, J.; Zhang, B.; Hou, L.; Zhong, M. Expression of HOX C13 in odontogenic tumors. Shanghai J. Stomatol. 2007, 16, 587–591. [Google Scholar]
  145. Marcinkiewicz, K.M.; Gudas, L.J. Altered epigenetic regulation of homeobox genes in human oral squamous cell carcinoma cells. Exp. Cell Res. 2014, 320, 128–143. [Google Scholar] [CrossRef]
  146. Marcinkiewicz, K.M.; Gudas, L.J. Altered histone mark deposition and DNA methylation at homeobox genes in human oral squamous cell carcinoma. J. Cell. Physiol. 2014, 229, 1405–1416. [Google Scholar] [CrossRef]
  147. Luo, J.; Wang, Z.; Huang, J.; Yao, Y.; Sun, Q.; Wang, J.; Shen, Y.; Xu, L.; Ren, B. HOXC13 promotes proliferation of esophageal squamous cell carcinoma via repressing transcription of CASP3. Cancer Sci. 2018, 109, 317–329. [Google Scholar] [CrossRef]
  148. Chen, F.; Li, Y.; Wang, L.; Hu, L. Knockdown of BMI-1 causes cell-cycle arrest and derepresses p16INK4a, HOXA9 and HOXC13 mRNA expression in HeLa cells. Med. Oncol. 2011, 28, 1201–1209. [Google Scholar] [CrossRef]
  149. Cantile, M.; Galletta, F.; Franco, R.; Aquino, G.; Scognamiglio, G.; Marra, L.; Cerrone, M.; Malzone, G.; Manna, A.; Apice, G.; et al. Hyperexpression of HOXC13, located in the 12q13 chromosomal region, in well-differentiated and dedifferentiated human liposarcomas. Oncol. Rep. 2013, 30, 2579–2586. [Google Scholar] [CrossRef] [Green Version]
  150. Komisarof, J.; McCall, M.; Newman, L.; Bshara, W.; Mohler, J.L.; Morrison, C.; Land, H. A four gene signature predictive of recurrent prostate cancer. Oncotarget 2017, 8, 3430–3440. [Google Scholar] [CrossRef]
  151. Yao, Y.; Luo, J.; Sun, Q.; Xu, T.; Sun, S.; Chen, M.; Lin, X.; Qian, Q.; Zhang, Y.; Cao, L.; et al. HOXC13 promotes proliferation of lung adenocarcinoma via modulation of CCND1 and CCNE1. Am. J. Cancer Res. 2017, 7, 1820–1834. [Google Scholar]
  152. Alvarado, D.M.; McCall, K.; Hecht, J.T.; Dobbs, M.B.; Gurnett, C.A. Deletions of 5′ HOXC genes are associated with lower extremity malformations including clubfoot and vertical talus. J. Med. Genet. 2016, 53, 250–255. [Google Scholar] [CrossRef]
  153. Gao, C.; Lu, W.; Lou, W.; Wang, L.; Xu, Q. Long noncoding RNA HOXC13-AS positively affects cell proliferation and invasion in nasopharyngeal carcinoma via modulating miR-383-3p/HMGA2 axis. J. Cell. Physiol. 2018, 234, 12809–12820. [Google Scholar] [CrossRef]
  154. Rinn, J.L.; Kertesz, M.; Wang, J.K.; Squazzo, S.L.; Xu, X.; Brugmann, S.A.; Goodnough, L.H.; Helms, J.A.; Farnham, P.J.; Segal, E.; et al. Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs. Cell 2007, 129, 1311–1323. [Google Scholar] [CrossRef]
  155. Li, L.; Liu, B.; Wapinski, O.L.; Tsai, M.C.; Qu, K.; Zhang, J.; Carlson, J.C.; Lin, M.; Fang, F.; Gupta, R.A.; et al. Targeted disruption of Hotair leads to homeotic transformation and gene derepression. Cell Rep. 2013, 5, 3–12. [Google Scholar] [CrossRef]
  156. Tang, Q.; Hann, S.S. HOTAIR: An Oncogenic Long Non-Coding RNA in Human Cancer. Cell Physiol. Biochem. 2018, 47, 893–913. [Google Scholar] [CrossRef]
  157. Botti, G.; Marra, L.; Malzone, M.G.; Anniciello, A.; Botti, C.; Franco, R.; Cantile, M. LncRNA HOTAIR as Prognostic Circulating Marker and Potential Therapeutic Target in Patients with Tumor Diseases. Curr. Drug Targets 2017, 18, 27–34. [Google Scholar] [CrossRef]
  158. Davis, A.P.; Capecchi, M.R. A mutational analysis of the 5′ HoxD genes: Dissection of genetic interactions during limb development in the mouse. Development 1996, 122, 1175–1185. [Google Scholar]
  159. Fromental-Ramain, C.; Warot, X.; Messadecq, N.; LeMeur, M.; Dollé, P.; Chambon, P. Hoxa-13 and Hoxd-13 play a crucial role in the patterning of the limb autopod. Development 1996, 122, 2997–3011. [Google Scholar]
  160. Zákány, J.; Duboule, D. Hox genes in digit development and evolution. Cell Tissue Res. 1999, 296, 19–25. [Google Scholar] [CrossRef]
  161. Muragaki, Y.; Mundlos, S.; Upton, J.; Olsen, B.R. Altered growth and branching patterns in synpolydactyly caused by mutations in HOXD13. Science 1996, 272, 548–551. [Google Scholar] [CrossRef]
  162. Raza-Egilmez, S.Z.; Jani-Sait, S.N.; Grossi, M.; Higgins, M.J.; Shows, T.B.; Aplan, P.D. NUP98-HOXD13 gene fusion in therapy-related acute myelogenous leukemia. Cancer Res. 1998, 58, 4269–4273. [Google Scholar]
  163. Pineault, N.; Buske, C.; Feuring-Buske, M.; Abramovich, C.; Rosten, P.; Hogge, D.E.; Aplan, P.D.; Humphries, R.K. Induction of acute myeloid leukemia in mice by the human leukemia-specific fusion gene NUP98-HOXD13 in concert with Meis1. Blood 2003, 101, 4529–4538. [Google Scholar] [CrossRef] [Green Version]
  164. Slape, C.; Hartung, H.; Lin, Y.W.; Bies, J.; Wolff, L.; Aplan, P.D. Retroviral insertional mutagenesis identifies genes that collaborate with NUP98-HOXD13 during leukemic transformation. Cancer Res. 2007, 67, 5148–5155. [Google Scholar] [CrossRef]
  165. Imren, S.; Heuser, M.; Gasparetto, M.; Beer, P.A.; Norddahl, G.L.; Xiang, P.; Chen, L.; Berg, T.; Rhyasen, G.W.; Rosten, P.; et al. Modeling de novo leukemogenesis from human cord blood with MN1 and NUP98HOXD13. Blood 2014, 124, 3608–3612. [Google Scholar] [CrossRef] [Green Version]
  166. Monlish, D.A.; Bhatt, S.T.; Duncavage, E.J.; Greenberg, Z.J.; Keller, J.L.; Romine, M.P.; Yang, W.; Aplan, P.D.; Walter, M.J.; Schuettpelz, L.G. Loss of Toll-like receptor 2 results in accelerated leukemogenesis in the NUP98-HOXD13 mouse model of MDS. Blood 2018, 131, 1032–1035. [Google Scholar] [CrossRef]
  167. Slape, C.; Lin, Y.W.; Hartung, H.; Zhang, Z.; Wolff, L.; Aplan, P.D. NUP98-HOX translocations lead to myelodysplastic syndrome in mice and men. J. Natl. Cancer Inst. Monogr. 2008, 39, 64–68. [Google Scholar] [CrossRef]
  168. Lin, Y.W.; Slape, C.; Zhang, Z.; Aplan, P.D. NUP98-HOXD13 transgenic mice develop a highly penetrant, severe myelodysplastic syndrome that progresses to acute leukemia. Blood 2005, 106, 287–295. [Google Scholar] [CrossRef] [Green Version]
  169. Greenblatt, S.; Li, L.; Slape, C.; Nguyen, B.; Novak, R.; Duffield, A.; Huso, D.; Desiderio, S.; Borowitz, M.J.; Aplan, P.; et al. Knock-in of a FLT3/ITD mutation cooperates with a NUP98-HOXD13 fusion to generate acute myeloid leukemia in a mouse model. Blood 2012, 119, 2883–2894. [Google Scholar] [CrossRef] [Green Version]
  170. Humeniuk, R.; Koller, R.; Bies, J.; Aplan, P.; Wolff, L. Brief report: Loss of p15Ink4b accelerates development of myeloid neoplasms in Nup98-HoxD13 transgenic mice. Stem Cells 2014, 32, 1361–1366. [Google Scholar] [CrossRef]
  171. Slape, C.; Liu, L.Y.; Beachy, S.; Aplan, P.D. Leukemic transformation in mice expressing a NUP98-HOXD13 transgene is accompanied by spontaneous mutations in Nras, Kras, and Cbl. Blood 2008, 112, 2017–2019. [Google Scholar] [CrossRef] [Green Version]
  172. Xu, H.; Menendez, S.; Schlegelberger, B.; Bae, N.; Aplan, P.D.; Göhring, G.; Deblasio, T.R.; Nimer, S.D. Loss of p53 accelerates the complications of myelodysplastic syndrome in a NUP98-HOXD13-driven mouse model. Blood 2012, 120, 3089–3097. [Google Scholar] [CrossRef]
  173. Zhong, Z.B.; Shan, M.; Qian, C.; Liu, T.; Shi, Q.Y.; Wang, J.; Liu, Y.; Liu, Y.; Huang, Y.X.; Pang, D. Prognostic significance of HOXD13 expression in human breast cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 11407–11413. [Google Scholar]
  174. Ke, D.; Yang, R.; Jing, L. Combined diagnosis of breast cancer in the early stage by MRI and detection of gene expression. Exp. Ther. Med. 2018, 16, 467–472. [Google Scholar] [CrossRef]
  175. Zhong, Z.; Shan, M.; Wang, J.; Liu, T.; Xia, B.; Niu, M.; Ren, Y.; Pang, D. HOXD13 methylation status is a prognostic indicator in breast cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 10716–10724. [Google Scholar]
  176. Shan, M.; Yin, H.; Li, J.; Li, X.; Wang, D.; Su, Y.; Niu, M.; Zhong, Z.; Wang, J.; Zhang, X.; et al. Detection of aberrant methylation of a six-gene panel in serum DNA for diagnosis of breast cancer. Oncotarget 2016, 7, 18485–18494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Han, L.; Xu, G.; Xu, C.; Liu, B.; Liu, D. Potential prognostic biomarkers identified by DNA methylation profiling analysis for patients with lung adenocarcinoma. Oncol. Lett. 2018, 15, 3552–3557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. von Heyking, K.; Roth, L.; Ertl, M.; Schmidt, O.; Calzada-Wack, J.; Neff, F.; Lawlor, E.R.; Burdach, S.; Richter, G.H. The posterior HOXD locus: Its contribution to phenotype and malignancy of Ewing sarcoma. Oncotarget 2016, 7, 41767–41780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Kumon, K.; Kobayashi, H.; Namiki, T.; Tsunematsu, Y.; Miyauchi, J.; Kikuta, A.; Horikoshi, Y.; Komada, Y.; Hatae, Y.; Eguchi, H.; et al. Frequent increase of DNA copy number in the 2q24 chromosomal region and its association with a poor clinical outcome in hepatoblastoma: Cytogenetic and comparative genomic hybridization analysis. Jpn. J. Cancer Res. 2001, 92, 854–862. [Google Scholar] [CrossRef]
  180. Delpretti, S.; Montavon, T.; Leleu, M.; Joye, E.; Tzika, A.; Milinkovitch, M.; Duboule, D. Multiple enhancers regulate Hoxd genes and the Hotdog LncRNA during cecum budding. Cell Rep. 2013, 5, 137–150. [Google Scholar] [CrossRef]
  181. Shah, N.; Sukumar, S. The Hox genes and their roles in oncogenesis. Nat. Rev. Cancer 2010, 10, 361–371. [Google Scholar] [CrossRef]
  182. Alharbi, R.A.; Pettengell, R.; Pandha, H.S.; Morgan, R. The role of HOX genes in normal hematopoiesis and acute leukemia. Leukemia 2013, 27, 1000–1008. [Google Scholar] [CrossRef]
  183. Cantile, M.; Franco, R.; Schiavo, G.; Procino, A.; Cindolo, L.; Botti, G.; Cillo, C. The HOX Genes Network in Uro-Genital Cancers: Mechanisms and Potential Therapeutic Implications. Curr. Med. Chem. 2011, 18, 4872–4884. [Google Scholar] [CrossRef]
  184. Plowright, L.; Harrington, K.J.; Pandha, H.S.; Morgan, R. HOX transcription factors are potential therapeutic targets in non-small-cell lung cancer. Br. J. Cancer 2009, 100, 470–475. [Google Scholar] [CrossRef]
  185. Morgan, R.; El-Tanani, M.; Hunter, K.D.; Harrington, K.J.; Pandha, H.S. Targeting HOX/PBX dimers in cancer. Oncotarget 2017, 8, 32322–32331. [Google Scholar] [CrossRef] [Green Version]
  186. Li, Z.; Zhao, X.; Zhou, Y.; Liu, Y.; Zhou, Q.; Ye, H.; Wang, Y.; Zeng, J.; Song, Y.; Gao, W.; et al. The long non-coding RNA HOTTIP promotes progression and gemcitabine resistance by regulating HOXA13 in pancreatic cancer. J. Transl. Med. 2015, 13, 84. [Google Scholar] [CrossRef]
  187. Morgan, R.; El-Tanani, M. HOX Genes as Potential Markers of Circulating Tumour Cells. Curr. Mol. Med. 2016, 16, 322–327. [Google Scholar] [CrossRef]
  188. Botti, G.; Cantile, M. Circulating long non-coding RNAs: Could they be a useful tool for cancer therapy monitoring? Expert Rev. Anticancer Ther. 2018, 18, 1167–1168. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the paralogous HOX13 genes involvement in the main human cancers.
Figure 1. Schematic representation of the paralogous HOX13 genes involvement in the main human cancers.
Cancers 11 00699 g001

Share and Cite

MDPI and ACS Style

Botti, G.; Cillo, C.; De Cecio, R.; Malzone, M.G.; Cantile, M. Paralogous HOX13 Genes in Human Cancers. Cancers 2019, 11, 699. https://doi.org/10.3390/cancers11050699

AMA Style

Botti G, Cillo C, De Cecio R, Malzone MG, Cantile M. Paralogous HOX13 Genes in Human Cancers. Cancers. 2019; 11(5):699. https://doi.org/10.3390/cancers11050699

Chicago/Turabian Style

Botti, Gerardo, Clemente Cillo, Rossella De Cecio, Maria Gabriella Malzone, and Monica Cantile. 2019. "Paralogous HOX13 Genes in Human Cancers" Cancers 11, no. 5: 699. https://doi.org/10.3390/cancers11050699

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop