PAX8 is transcribed aberrantly in cervical tumors and derived cell lines due to complex gene rearrangements

  • Authors:
    • Eduardo López-Urrutia
    • Abraham Pedroza-Torres
    • Jorge Fernández-Retana
    • David Cantu De Leon
    • Fermín Morales-González
    • Nadia Jacobo-Herrera
    • Oscar Peralta-Zaragoza
    • Jorge García-Mendez
    • Verónica García-Castillo
    • Osvaldo Bautista-Isidro
    • Carlos Pérez-Plasencia
  • View Affiliations

  • Published online on: May 11, 2016     https://doi.org/10.3892/ijo.2016.3515
  • Pages: 371-380
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

The transcription factor PAX8, a member of the paired box-containing gene family with an important role in embryogenesis of the kidney, thyroid gland and nervous system, has been described as a biomarker in tumors of the thyroid, parathyroid, kidney and thymus. The PAX8 gene gives rise to four isoforms, through alternative mRNA splicing, but the splicing pattern in tumors is not yet established. Cervical cancer has a positive expression of PAX8; however, there is no available data determining which PAX8 isoform or isoforms are present in cervical cancer tissues as well as in cervical carcinoma-derived cell lines. Instead of a differential pattern of splicing isoforms, we found numerous previously unreported PAX8 aberrant transcripts ranging from 378 to 542 bases and present in both cervical carcinoma-derived cell lines and tumor samples. This is the first report of PAX8 aberrant transcript production in cervical cancer. Reported PAX8 isoforms possess differential transactivation properties; therefore, besides being a helpful marker for detection of cancer, PAX8 isoforms can plausibly exert differential regulation properties during carcinogenesis.

Introduction

The paired-box (PAX) family, initially described in Drosophila but highly conserved among vertebrates, encompasses nine DNA-binding proteins that function as transcription factors. PAX proteins share a common structure: a DNA-binding paired domain (PAired boX, hence the name of the family), a PST homeodomain and an octapeptide; genes from this family differ in the length of the first two and the presence or absence of the last domain. During development, PAX genes are expressed in different regions of the forming embryo, thus they constitute important gene expression regulators (1,2). PAX8 has a fairly well understood role in development: it is associated to morphogenesis of the kidney, thyroid gland and nervous system (3,4), and its expression is regulated by alternative splicing (5).

The PAX8 gene, located at 2q13 and spanning 63 kb, is transcribed, processed and translated to give rise to six reported mRNAs: the originally described PAX8A, B, C, and D, which are in turn translated to the corresponding protein isoforms (6), and a more recently described transcript, PAX8F, that shares the PAX8A ORF and bears an extended 5′UTR (7). The existence of PAX8B has been questioned and its GenBank entry (NM_013951.3) is currently suppressed. All PAX8 isoforms bind DNA with similar affinity; however, PAX8A and PAX8B display higher transactivation potential (6).

Recent studies suggest a role for PAX8 in carcinogenesis that makes it a biomarker candidate; still, more studies are necessary to completely understand this role. So far, PAX8 has been associated to important cancer processes such as Retinoblastoma/E2F1 transcription (8,9) and telomerase activity (10). Another study recently showed that by inhibiting PAX8 expression with shRNAs in ovarian cancer cell lines, their viability diminished (11); this suggests a direct involvement of PAX8 in cell proliferation which, in turn, could represent a potential therapeutic target.

An important fact that links PAX8 to carcinogenesis is the PAX8/PPARγ genetic rearrangement found in 30–35% of follicular thyroid carcinoma and up to 10% of follicular thyroid adenoma (1214). Hence, t(2;3)(q13;p25) fuses the promoter and 5-coding portion of the transcription factor PAX8 that involves DNA binding domains to the full-length coding sequence of the nuclear receptor peroxisome proliferator-activated receptor-γ 1 gene (15).

As a biomarker, PAX8 is frequently detected in epithelial tumors of the thyroid, parathyroid, kidney, thymus and female genital tract (16,17). In thyroid tumors, PAX8 overexpression has been found in follicular and papillary adenomas and carcinomas (16,18,19). Wilms' tumors and nephrogenic adenomas, among other renal tumors, show PAX8 overexpression owing to its role during renal organogenesis (2022). PAX8-expressing oviduct cells have been pointed out as the probable origin of ovarian and endometrial cancers, rendering PAX8 as an early gynecological cancer marker, associated with poor disease outcome (23,24). Only few groups have reported PAX8 overexpression in cervical cancer employing distinct methodologies (2527).

A genome-wide search for PAX8 binding sites using ChIP-Seq technology was recently carried out on thyroid cells. The authors identified binding sites that show that PAX8 regulates genes involved in cell proliferation and differentiation, as well as binding sites located in non-promoter regions, including introns, that hint to a role as a distal transcriptional regulator (28). However, none of the aforementioned reports of PAX8 as a tumor marker or functional PAX8 studies specify whether its results concern any particular PAX8 isoform; therefore, we set out to investigate which isoform or isoforms were expressed in a set of locally advanced cervical tumor samples compared to normal cervical tissues. PAX8 isoforms could possibly activate different genes contributing to carcinogenesis, given their different in vitro transactivation properties (6).

We had hypothesized that, since splicing is highly altered in cancer (29), we would find a particular PAX8 splicing pattern in cervical carcinoma tumor samples and cervical carcinoma-derived cell lines, but RACE experiments revealed the expression of previously unreported aberrant transcripts instead. We also found evidence hinting of a complex gene rearrangement in the genomic DNA from the cell lines, providing a possible explanation for these transcripts. To our knowledge, this is the first study showing that PAX8 gene undergoes complex rearrangement processes that could contribute to establish the CC tumor phenotype.

Materials and methods

Expression data Tissue samples

Patients were prospectively enrolled into the National Cancer Institute of Mexico (INCAN) tumor-banking protocol at the time of diagnosis. All patients included accepted and signed informed consent; institutional ethics and scientific board committees approved the protocol. Immediately after punch biopsy, tumor samples were split into three pieces, one for the pathologic confirmation of at least 80% of tumor cells, mandatory for this type of molecular profiles, and the remaining two for RNA and DNA isolation. RNA and DNA biopsies were frozen in liquid nitrogen until nucleic acid extraction. Eligibility criteria were i) patients with a confirmed pathologic diagnosis of CC staged IB2 up to IIIB (LACC); ii) biopsies with pathology report with >80% of tumors cells; hence, the genomic analysis is mainly addressed for tumor cells; iii) age greater to 20 and less than 60 years; iv) high-quality DNA and RNA; v) no presence of comor bidities; vi) and without previous oncological treatment.

Healthy cervical tissues were obtained from patients who had undergone hysterectomy by uterine myomatosis. Inclusion criteria were: i) no previous cervical surgery (such as the loop electrosurgical excision procedure or cone biopsy), ii) no HPV infection, iii) no hormonal treatment, and iv) at last three previous negative Pap smears.

PAX8 mRNA expression

We designed two sets of primers, one for detection of PAX8 exons 3–4 and another for exons 10–11 in the mRNA (NM_003466.3; see Fig. 1). RNA was extracted from 18 tumoral and 20 healthy, HPV-free, cervical tissue samples using the TRIzol reagent (Life Technologies cat. # 15596-026), following the manufacturer's recommendations. Its integrity was verified through agarose gel electrophoresis and it was quantitated by spectrophotometry (OD260/280>1.9). For first strand synthesis, we mixed 500 ng RNA, 0.5 mM dNTPs, 10 mM DT, 1× First-strand buffer and 180 ng random hexamers, heated them to 42°C for 2 min; then we added 200 units of SuperScript II reverse transcriptase (Invitrogen cat. #18064-022) and incubated the reaction for 50 min at 42°C, followed by 15 min at 70°C. The resulting cDNAs was probed for DNA contamination by performing no-RT assays. Quantitative PCR (qPCR) reactions were performed in triplicate using LightCycler® 480 SYBR Green I Master Mix (Roche, cat. # 04 707 516 0081) in the LightCycler 480 instrument, following the manufacturer's recommendations (40 amplification cycles, Tm 58°C). End-point PCR was performed using the PCR Master mix 2× (Thermo Scientific, cat. #K0171) 0.5 μM de cada primer (35 cycles, Tm 56°C).

Western blotting

Total proteins were extracted from 80–85% confluent cell cultures. Culture media was removed and cells were rinsed twice with PBS, scraped off from culture dishes, and lysed using the RIPA lysis reagent (Santa Cruz Biotechnology) following the manufacturer's recommendations.

Total protein (50 μg) was mixed with Laemmli sample buffer, boiled, separated in 12% or 15% SDS-PAGE and transferred onto a Hybond-P PVDF membrane (Amersham-GE Healthcare). Membranes were probed overnight using a 1:500 (v/v) dilution of the appropriate antibody; for detection, 1:2500 (v/v) dilutions of HRP anti-rabbit or anti-mouse conjugate antibodies (Santa Cruz Biotechnology) were used. Finally, using the SuperSignal WestFemto chemiluminescent substrate (Thermo Scientific), the membranes were scanned in the C-Digit blot scanner (Li-Cor) and the images were analyzed in the associated ImageStudio software (LiCor). Membranes were stripped and re-probed for actin detection as a loading control. The commercial antibodies used were anti-PAX8 (Cell Signaling Technology #9857) and anti-β actin (SantaCruz Biotechnology sc-1616). A representative image from three independent experiments is shown.

RT-PCR, 5′ RACE and Exon PCR

DNA and RNA were isolated from tumor samples or from the CaSki (ATCC, CRL1550) and SiHa (ATCC, HTB35) cells grown to approximately 80–85% confluence, using the TRIzol reagent (Life technologies) following the manufacturer's recommendations.

First strand cDNA synthesis was carried out for 90 min at 42°C. Total RNA (5 μg) was mixed with 1 mM oligo dT, 0.25 mM of each dNTP, 1× First-strand buffer, 10 mM DTT and 200 units of M-MLV reverse transcriptase (Promega). This reaction (1 μl) was used as a template for 50 μl PCR reactions for amplification of PAX8 transcripts using the PAX8Fw and PAX8Rv primers (Table I) 0.2 μM each, 25 μM of each dNTP, 1× Herculase II buffer and 1 unit of Herculase II fusion DNA polymerase (Agilent).

Table I

Primers used for 5′ RACE and Exon PCR experiments.

Table I

Primers used for 5′ RACE and Exon PCR experiments.

Name5′-3′Sequence Nt position
Primers for 5′ RACE
 15′ RACE adapter GCU GAU GGC GAU GAA UGA ACA CUG CGU UUG CUG GCU UUG AUG AAANA
 2RACE Outer primer GCT GAT GGC GAT GAA TGA ACA CTGNA
 3RACE Inner primer CGC GGA TCC GAA CAC TGC GTT TGC TGG CTT TGA TGNA
 4PAX8UTR3′RV AGT CCT CCT GTT GCT CAG TCG CT1561–1539a
 5PAX8 RV CTA CAG ATG GTC AAA GGC CGT1519–1499a
Primers for RT-PCR
 1PAX8 FW ATG CCT CAC AAC TCC ATC AGA167–187a
 2PAX8 RV CTA CAG ATG GTC AAA GGC CGT1519–1499a
 3GAPDH FW CCT CAA GAT CAT CAG CAA TGC CT617–639b
 4GAPDH RV TCA CGC CAC AGT TTC CCG GAG781–761b
Primers for Exon PCR
 6PAX8EX1Fw GAT GCA GGC ATC GAA TCT C399–417c
 7PAX8EX1Rv ACG CTC TCG AGA TCC AAC C639–621c
 8PAX8EX2Fw ATC CCC ACC CAA ACT CCT AC64185–64204c
 9PAX8EX2Rv TCA GCT GGA GAA GTC AAG CC64468–64449c
 10PAX8EX3Fw TGT CTA AAG ACC CCA CCT GC66487–66506c
 11PAX8EX3Rv AGC CAG GCC TTT CTT GTC TC66824–66805c
 12PAX8EX4Fw GCC ATG AGT TCT CTT TCC TCC68332–68352c
 13PAX8EX4Rv GGT ATG CTG AAG GGG AGG TG68538–68519c
 14PAX8EX5Fw ACT ACC CCA GAG TCA CCC AG68961–68980c
 15PAX8EX5Rv AAA GCC TCA GCA AAC TGC TC69215–69196c
 15PAX8EX6Fw TTT GGC CTA GAG CAT GAA TAG69376–69396c
 16PAX8EX6Rv GAG CAC AGG CTC ATT TGG AG69692–69673c
 17PAX8EX7Fw CCT AAG ACA CAG GCT CAG GG74363–74382c
 18PAX8EX7Rv AGC CAA GCT CTT CAG TCC C74620–74602c
 19PAX8EX8Fw CTT GTG CGT GTT CCC TCC75738–77555c
 20PAX8EX8Rv GTC TGC CCT GAG GAC CC76060–76044c
 21PAX8EX9Fw AGC CTC AGG AGA GTG AGA TG84030–84040c
 22PAX8EX9Rv GTC CCA CCT TGC TCC AAT AC84269–84250c
 23PAX8EX10Fw CCT GCA TTG ATG CCC TTC91115–91132c
 24PAX8EX10Rv AGG TAA CCT TTG ACC CAC CC91335–91316c

{ label (or @symbol) needed for fn[@id='tfn1-ijo-49-01-0371'] } NA, not applicable. Nucleotide positions: relative to BC001060,

a relative to NM_003466,

b relative to NM_002046.5,

c relative to NG_012384.

Amplification conditions were 5 min pre-incubation at 95°C followed by 35 cycles of 20 sec at 95°C, 20 sec at 53°C and 70 sec at 68°C, followed by a final step of 4 min at 68°C. GAPDH transcripts were detected as a loading control, using the corresponding primers (Table I) 0.25 μM each, 0.2 mM of each dNTP, 2.5 mM MgCl2 and 0.75 units of GoTaq DNA Polymerase (Promega) in a 25 μl volume. The amplification conditions were: 5 min pre-incubation at 95°C, 35 cycles of 30 sec at 95°C, 20 sec at 57°C and 20 sec at 72°C, followed by a final elongation step of 7 min at 72°C.

Rapid amplification of cDNA ends (RACE) was performed using the First Choice RLM-RACE kit (Life Technologies-Ambion) according to the manufacturer's protocol

Briefly, 1 μg of RNA from HeLa cells, SiHa cells or tumor samples was dephosphorylated with Calf intestine Acid Phosphatase (CIP) to remove the available phosphates from rRNA, tRNA, uncapped and partial transcripts; the cap structure was subsequently removed from mature mRNAs with Tobacco Acid Phosphatase (TAP) so that only these transcripts could acquire the 5′ RACE adapter (Table I) trough ligation with T4 RNA ligase. Next, a random-primed RT reaction was performed using M-MLV reverse transcriptase (Promega). The PAX8 transcripts were amplified from the cDNA pool through nested PCR, the first round was performed using the 5′ RACE Outer Primer from the kit in combination with the PAX8UTR3′Rv primer, while the 5′ RACE Inner Primer and the PAX8Rv primer were used for the second round (Table I). The resulting amplicons were resolved in ethidium bromide-stained 2% agarose gels.

PAX8 exons were amplified from genomic DNA with ad-hoc primers based on the PAX8 complete sequence (GenBank Gene ID: 7849; locus NG_012384). These primers were designed using the ExonPrimer server (ihg. helmholtz-muenchen.de/ihg/ExonPrimer.html) and are shown in Table I. PCR was performed using 25 ng gDNA template, 0.25 μM of each primer, 0.2 mM of each dNTP, 2.5 mM MgCl2 and 0.75 units of GoTaq DNA Polymerase (Promega) in a 25 μl volume. The amplification conditions were as follows: 5 min pre-incubation at 92°C followed by 35 cycles of 20 sec at 92°C, 20 sec at 67°C and 30 sec at 72°C, followed by a final step of 7 min at 72°C. All the resulting amplicons were resolved in ethidium bromide-stained 1.5% or 2% agarose gels.

Cloning and sequence analysis

The RACE product bands were excised from the agarose gels and purified using the QIAquick Spin kit (Qiagen) according to the manufacturer's protocol. After purification, the RACE products were cloned into the pGEM T-Easy plasmid (Promega). Plasmids from positive colonies were analyzed by restriction; each clone was sequenced in both chains using universal primers and the Big Dye Terminator Ready Reaction kit (Perkin-Elmer) and analyzed in the ABI PRISM 3130xl Genetic Analyzer System (Applied Biosystems). Sequence information was analyzed using the CLC Bio Main Workbench (CLC Bio; Qiagen) as well as the BLAT (genome.ucsc.edu) and BLAST (blast.ncbi.nlm.nih.gov) algorithms.

In order to search for sequence similarity among clones, we performed a CLUSTAL alignment and checked it manually. Based on this alignment, we searched for the most adequate model of evolution using the FindModel server (http://www.hiv.lanl.gov/content/sequence/findmodel/findmodel.html), and the GTR model was selected to perform a maximum likelihood phylogenetic analysis employing a 1000-replicate bootstrap analysis. Finally, a tree was constructed using the neighborjoining method.

Results

PAX8 mRNA expression

Since splicing events remove exons 8–10 (6), we designed the two sets of primers depicted in Fig. 1A to assess the PAX8 mRNA expression and integrity. We amplified independently exons 3–4 and exons 10–11 from 20 normal cervix samples and 18 cervical tumor samples and found interesting results: Although overall messenger expression was lower in tumors (Fig. 1B), the exon 3–4:10–11 ratio (Fig. 1C) was different in tumors when compared to normal cervical samples; which was more evident upon agarose-gel separation of the amplicons (Fig. 1D). This observation suggested the presence of transcripts containing only the 3′-most end of the PAX8 mRNA that may or may not correspond to reported splicing products.

PAX8 protein expression

In order to detect the PAX8 isoforms possibly expressed in cervical carcinoma-derived cell lines, we selected an antibody raised against the carboxi-terminal region of PAX8, shared by its isoforms. Previously, PAX8 has been detected through immunocytochemistry but current cell imaging techniques are yet to discriminate between isoforms sharing structural motifs. Consistently with reported detection in HEK cells, we observed a 50 kDa band, apparently corresponding to PAX8A in CaSki and SiHa cells; a 43 kDa band was detected, corresponding to PAX8C (Fig. 2). The reported PAX8A and C transcripts bear exons 3–4 and 10–11, therefore, their presence was not likely to account for the imbalance between these mRNA regions in cervical tumors.

PAX8 aberrant transcripts

We next set out to identify the PAX8 transcripts possibly expressed in cervical carcinoma-derived cell lines at the mRNA level. The 1372-nt PAX8A/F ORF was present in both CaSki and SiHa cells (Fig. 3), consistent with our protein detection results. The PAX8C and PAX8D transcripts were also detected in both cell lines. However, what really caught our interest were the 400–500 nt products that we found in both lines, more evident in the SiHa lane in Fig. 3.

Due to the fact that exons 10–11 were detected at higher levels in tumor samples, we reasoned there was a population of transcripts sharing the 3′ end of the PAX8 ORF; so we set up a 5′ RACE strategy using a reverse primer located 20 nt downstream from the end codon (PAX8UTR3′RV, Table I). We performed two independent 5′ RACE experiments from CaSki and SiHa cell lines (Fig. 4A and B). A single band of about 450 nt was observed as product from all these experiments. Furthermore, to rule out the possibility that the detected transcripts were exclusive to cell lines, we performed 5′ RACE experiments on three tumor samples and obtained similar ~450 bp products from two of them (Fig. 4C). All these 5′ RACE products were excised from the gel and cloned into a T-protruding vector.

We expected to obtain a number of positive colonies carrying a 450 bp insert but instead, we surprisingly obtained colonies carrying differently sized inserts. Plasmids from these colonies were purified and sequenced. In total, we obtained 33 sequences that are shown in Fig. 5A, aligned with the four reported PAX8 transcripts. Solid lines represent identical sequences, whereas sequences represented by dotted lines did not align to the PAX8 transcripts. The sequences of all transcripts were submitted to GenBank under accessions KJ545852 to KJ545885. The clones obtained form tumor samples and cervical carcinoma-derived cell lines form a single pool of PAX8 aberrant transcripts. The reported transcripts and our sequenced clones clustered separately from each other, and no evident clustering was found within our clone group (Fig. 5B; Table II).

Table II

Expected amplicon sizes from Exon PCR experiments.

Table II

Expected amplicon sizes from Exon PCR experiments.

Exon no.Exon PCR amplicon size (bp)
2241
3284
4338
5207
6255
7317
8258
9323
10240
11221

Upon analysis of the sequences from our collection of cloned transcripts we found that they only shared the 3′-most 170 nt with the PAX8 coding sequence. This region only represents two (11 and 12) of the ten exons that constitute the PAX8 ORF, encoded in exons 2–12 (see the representation in Fig. 6). Upstream from that region, the 33 clones share a 30-nt sequence from intron 10, which contains the poly pyrimidine tract and is located 5 nt upstream of the py-AG splicing acceptor site, adjacent to exon 11. The remaining 5′ region of the cloned transcripts was variable among them, comprising sequences that align to other regions of exon 10 with multiple single nucleotide mismatches. Performing a BLAST search with our clones produced variable results pointing at contigs or chromosome assemblies (data not shown). Taken together, the results from both analyses suggested genomic instability in the PAX8 gene, because the aberrant transcripts that we were able to isolate contain sequences from different non-contiguous genomic locations. Noteworthy, we did not find a definite sequence or pattern corresponding to either tumor samples or cell lines, reaffirming the idea that PAX8 produces similar aberrant transcripts in both cervical tumors and derived cell lines.

Some of the aberrant transcripts contain open reading frames, but none of them comprises its full length. If translated, these ORFs should lead to short (12–22 kDa) proteins sharing the carboxi-terminus and thus the PST transactivation domain from PAX8 (Fig. 6).

PAX8 exon amplification

Since all the detected transcripts conserved only the 3′-most introns, we wanted to know whether the missing introns were actually present in the genomic DNA of the cells. As an initial approach to assess the integrity of the PAX8 gene, we designed primers along the intronic regions flanking each exon (Fig. 7A) with the aid of the ExonPrimer software tool. We targeted only exons 2–11, since exon 1 is not translated in any reported PAX8 mRNA and exon 12, while protein coding, is more than 2.5 kb long. A validation experiment was performed using DNA isolated from peripheral blood lymphocytes from a healthy patient, it yielded the expected amplicon sizes, as seen in Fig. 7B.

The results using CaSki and SiHa cell DNA were very interesting: amplicons corresponding to exons 2–9 failed to amplify in both lines, save for exon 5 that was detected only in SiHa DNA (Fig. 7C). Notably, only exons 10 and 11, those identified in the aberrant transcripts, were detected through exon PCR. This suggests, together with the sequences present in the cloned transcripts, that the PAX8 gene may be altered in cervical carcinoma.

Discussion

In this study, we analyzed the expression levels of two regions of the PAX8 mRNA in a set of tumoral and healthy cervical tissue samples, and found an imbalance between their expression levels. This imbalance can be attributed to the presence of aberrant transcripts containing the 3′-most region of the mRNA. Previous work on female genital tract tumors has shown a consistent PAX8 expression in ovary tumors, such as non-mucinous carcinomas, including serous, endometroid, clear cell and transitional cell carcinomas (30,31). Only one investigation has reported PAX8 expression in carcinomas of uterine cervix, but the results are in the context of epithelial tumors (20).

We aimed to assess the possible differential splicing of the PAX8 transcripts in cervical carcinoma-derived cell lines and tumors. Upon performing these experiments, we found a completely different scenario, rather than the expected protein isoforms, we detected a single protein and several shorter putative isoforms. Consequently, we aimed to detect the corresponding shorter transcripts.

Notably, while we did detect a band of the expected molecular weight by western blotting, the full-length ORF was only amplified with a high sensitivity polymerase and was not detected through 5′ RACE experiments (compare Figs. 3 and 4). This suggests a very low abundance of this transcripts and, thus, high translation efficiency of this mRNA. A high translation efficiency has been already suggested for this transcript by Szczepanek-Parulska and co-workers (32), who amplified a 587/684 bp fragment to detect PAX8A/F and hypothesized that the extra 97 bases at the 5′UTR participate in translational regulation.

The 5′ RACE experiments yielded a collection of transcripts ranging from 378 to 542 (Fig. 4), out of the roughly 1300 bases of the full-length PAX8 ORF. Shorter-than-expected transcripts from a 10-intron gene suggest alternative splicing, so it was puzzling when the sequencing results only revealed the two 3′-most coding exons from the ORF preceded by unknown and intronic sequences. The fact that similar transcripts were obtained from different independent sources strongly suggests that the aberrant transcripts that we found are product of a recurring process, present in both cervical carcinoma-derived cell lines and tumor samples.

The production of novel transcripts through alterations of alternative splicing is a known trait of cancer cells (29, reviewed in ref. 33) but only a few genes have been reported to produce aberrant transcripts containing sequences other than exons, such as FHIT and WWOX; this phenomenon has been attributed to their proximity to the chromosome fragile sites FRA3B and FRA16D, respectively (34,35). PAX8, in turn, is located at the 2q13 locus (36,37), close to the FRA2B fragile site (38,39), which prompted us to try and assess the integrity of the PAX8 gene.

Rather than trying to establish the genomic status, our exon amplification experiments show that it is a worthwhile venture. These results suggest that the genomic sequence of the PAX8 gene is at least partially disrupted, or that it harbors a considerable amount of mutations that impede amplification of exons 2 to 10. Noteworthy, these results seem to correspond to the sequences found in the aberrant transcripts, since only exons 10 and 11 were present in them, only these same exons were detectable through Exon PCR. The PAX8 gene has previously proven to be unstable in tumor cells, the (2;3)(q13;p25) translocation, detected in both follicular thyroid carcinomas and adenomas (40), leads to the formation of a chimeric PAX8-peroxisome proliferator-activated receptor (PPAR-γ) oncogene (15). Moreover, cervical carcinoma is prone to chromosome abnormalities as human papillomavirus has been demonstrated to produce specific chromosomal imbalances in transfected keratinocytes (41). In any case, we are conducting further research to establish the nature of the suggested genomic alterations in the PAX8 gene.

Taking the data together, we find it reasonable to speculate that the cDNA aberrations that we found are the result of genomic modifications characteristic of cervical carcinoma. Even so, it is notable that these probable genomic modifications are likely not present in every cell from our samples; otherwise we would not have been able to detect the full-length PAX8 in total protein extracts.

Here we report the presence of a number of aberrant transcripts produced by the PAX8 gene cervical cancer tissues. These transcripts are present in cervical carcinoma-derived cell lines and tumors and likely encode shorter isoforms. Further studies would shed light on the implications of these transcripts in carcinogenesis and their role as cause and/or consequence of other molecular phenomena.

Acknowledgements

E.L.U. is indebted to CONACyT México for a Retention grant (RETENCION-191616). This work was partially supported by CONACyT (SALUD-2009-01-113948 and SALUD-2014-1-233733). We thank Jorge E. Campos for assistance in phylogenetic analysis.

References

1 

Gruss P and Walther C: Pax in development. Cell. 69:719–722. 1992. View Article : Google Scholar : PubMed/NCBI

2 

Goulding MD, Lumsden A and Gruss P: Signals from the notochord and floor plate regulate the region-specific expression of two Pax genes in the developing spinal cord. Development. 117:1001–1016. 1993.PubMed/NCBI

3 

Mansouri A, Hallonet M and Gruss P: Pax genes and their roles in cell differentiation and development. Curr Opin Cell Biol. 8:851–857. 1996. View Article : Google Scholar : PubMed/NCBI

4 

Mansouri A, Goudreau G and Gruss P: Pax genes and their role in organogenesis. Cancer Res. 59(Suppl): S1707–S1710. 1999.

5 

Poleev A, Fickenscher H, Mundlos S, Winterpacht A, Zabel B, Fidler A, Gruss P and Plachov D: PAX8, a human paired box gene: Isolation and expression in developing thyroid, kidney and Wilms' tumors. Development. 116:611–623. 1992.PubMed/NCBI

6 

Kozmik Z, Kurzbauer R, Dörfler P and Busslinger M: Alternative splicing of Pax-8 gene transcripts is developmentally regulated and generates isoforms with different transactivation properties. Mol Cell Biol. 13:6024–6035. 1993. View Article : Google Scholar : PubMed/NCBI

7 

Szczepanek-Parulska E, Szaflarski W, Piątek K, Budny B, Jaszczyńska-Nowinka K, Biczysko M, Wierzbicki T, Skrobisz J, Zabel M and Ruchała M: Alternative 3′acceptor site in the exon 2 of human PAX8 gene resulting in the expression of unknown mRNA variant found in thyroid hemiagenesis and some types of cancers. Acta Biochim Pol. 60:573–578. 2013.

8 

Miccadei S, Provenzano C, Mojzisek M, Natali PG and Civitareale D: Retinoblastoma protein acts as Pax 8 transcriptional coactivator. Oncogene. 24:6993–7001. 2005. View Article : Google Scholar : PubMed/NCBI

9 

Li CG, Nyman JE, Braithwaite AW and Eccles MR: PAX8 promotes tumor cell growth by transcriptionally regulating E2F1 and stabilizing RB protein. Oncogene. 30:4824–4834. 2011. View Article : Google Scholar : PubMed/NCBI

10 

Chen Y-J, Campbell HG, Wiles AK, Eccles MR, Reddel RR, Braithwaite AW and Royds JA: PAX8 regulates telomerase reverse transcriptase and telomerase RNA component in glioma. Cancer Res. 68:5724–5732. 2008. View Article : Google Scholar : PubMed/NCBI

11 

Cheung HW, Cowley GS, Weir BA, Boehm JS, Rusin S, Scott JA, East A, Ali LD, Lizotte PH, Wong TC, et al: Systematic investigation of genetic vulnerabilities across cancer cell lines reveals lineage-specific dependencies in ovarian cancer. Proc Natl Acad Sci USA. 108:12372–12377. 2011. View Article : Google Scholar : PubMed/NCBI

12 

Marques AR, Espadinha C, Catarino AL, Moniz S, Pereira T, Sobrinho LG and Leite V: Expression of PAX8-PPAR gamma 1 rearrangements in both follicular thyroid carcinomas and adenomas. J Clin Endocrinol Metab. 87:3947–3952. 2002.PubMed/NCBI

13 

French CA, Alexander EK, Cibas ES, Nose V, Laguette J, Faquin W, Garber J, Moore F Jr, Fletcher JA, Larsen PR, et al: Genetic and biological subgroups of low-stage follicular thyroid cancer. Am J Pathol. 162:1053–1060. 2003. View Article : Google Scholar : PubMed/NCBI

14 

Dwight T, Thoppe SR, Foukakis T, Lui WO, Wallin G, Höög A, Frisk T, Larsson C and Zedenius J: Involvement of the PAX8/peroxisome proliferator-activated receptor gamma rearrangement in follicular thyroid tumors. J Clin Endocrinol Metab. 88:4440–4445. 2003. View Article : Google Scholar : PubMed/NCBI

15 

Kroll TG, Sarraf P, Pecciarini L, Chen CJ, Mueller E, Spiegelman BM and Fletcher JA: PAX8-PPARgamma1 fusion oncogene in human thyroid carcinoma [corrected]. Science. 289:1357–1360. 2000. View Article : Google Scholar : PubMed/NCBI

16 

Laury AR, Perets R, Piao H, Krane JF, Barletta JA, French C, Chirieac LR, Lis R, Loda M, Hornick JL, et al: A comprehensive analysis of PAX8 expression in human epithelial tumors. Am J Surg Pathol. 35:816–826. 2011. View Article : Google Scholar : PubMed/NCBI

17 

Ordóñez NG: Value of PAX 8 immunostaining in tumor diagnosis: A review and update. Adv Anat Pathol. 19:140–151. 2012. View Article : Google Scholar : PubMed/NCBI

18 

Nonaka D, Tang Y, Chiriboga L, Rivera M and Ghossein R: Diagnostic utility of thyroid transcription factors Pax8 and TTF-2 (FoxE1) in thyroid epithelial neoplasms. Mod Pathol. 21:192–200. 2008.

19 

Scouten WT, Patel A, Terrell R, Burch HB, Bernet VJ, Tuttle RM and Francis GL: Cytoplasmic localization of the paired box gene, Pax-8, is found in pediatric thyroid cancer and may be associated with a greater risk of recurrence. Thyroid. 14:1037–1046. 2004. View Article : Google Scholar

20 

Ozcan A, Shen SS, Hamilton C, Anjana K, Coffey D, Krishnan B and Truong LD: PAX 8 expression in non-neoplastic tissues, primary tumors, and metastatic tumors: A comprehensive immuno histochemical study. Mod Pathol. 24:751–764. 2011. View Article : Google Scholar : PubMed/NCBI

21 

Tong G-X, Yu WM, Beaubier NT, Weeden EM, Hamele-Bena D, Mansukhani MM and O'Toole KM: Expression of PAX8 in normal and neoplastic renal tissues: An immunohistochemical study. Mod Pathol. 22:1218–1227. 2009. View Article : Google Scholar : PubMed/NCBI

22 

Albadine R, Schultz L, Illei P, Ertoy D, Hicks J, Sharma R, Epstein JI and Netto GJ: PAX8 (+)/p63 (−) immunostaining pattern in renal collecting duct carcinoma (CDC): A useful immunoprofile in the differential diagnosis of CDC versus urothelial carcinoma of upper urinary tract. Am J Surg Pathol. 34:965–969. 2010. View Article : Google Scholar : PubMed/NCBI

23 

Mhawech-Fauceglia P, Wang D, Samrao D, Godoy H, Pejovic T, Liu S and Lele S: Pair-Box (PAX8) protein-positive expression is associated with poor disease outcome in women with endometrial cancer. Br J Cancer. 107:370–374. 2012. View Article : Google Scholar : PubMed/NCBI

24 

Wang Y, Wang Y, Li J, Yuan Z, Yuan B, Zhang T, Cragun JM, Kong B and Zheng W: PAX8: A sensitive and specific marker to identify cancer cells of ovarian origin for patients prior to neoadjuvant chemotherapy. J Hematol Oncol. 6:602013. View Article : Google Scholar : PubMed/NCBI

25 

Tacha D, Zhou D and Cheng L: Expression of PAX8 in normal and neoplastic tissues: A comprehensive immunohistochemical study. Appl Immunohistochem Mol Morphol. 19:293–299. 2011. View Article : Google Scholar : PubMed/NCBI

26 

Kenny SL, McBride HA, Jamison J and McCluggage WG: Mesonephric adenocarcinomas of the uterine cervix and corpus: HPV-negative neoplasms that are commonly PAX8, CA125, and HMGA2 positive and that may be immunoreactive with TTF1 and hepatocyte nuclear factor 1-β. Am J Surg Pathol. 36:799–807. 2012. View Article : Google Scholar : PubMed/NCBI

27 

Shukla A, Thomas D and Roh MH: PAX8 and PAX2 expression in endocervical adenocarcinoma in situ and high-grade squamous dysplasia. Int J Gynecol Pathol. 32:116–121. 2013. View Article : Google Scholar

28 

Ruiz-Llorente S, Carrillo Santa de Pau E, Sastre-Perona A, Montero-Conde C, Gómez-López G, Fagin JA, Valencia A, Pisano DG and Santisteban P: Genome-wide analysis of Pax8 binding provides new insights into thyroid functions. BMC Genomics. 13:1472012. View Article : Google Scholar : PubMed/NCBI

29 

David CJ and Manley JL: Alternative pre-mRNA splicing regulation in cancer: Pathways and programs unhinged. Genes Dev. 24:2343–2364. 2010. View Article : Google Scholar : PubMed/NCBI

30 

Nonaka D, Chiriboga L and Soslow RA: Expression of pax8 as a useful marker in distinguishing ovarian carcinomas from mammary carcinomas. Am J Surg Pathol. 32:1566–1571. 2008. View Article : Google Scholar : PubMed/NCBI

31 

Zhao L, Guo M, Sneige N and Gong Y: Value of PAX8 and WT1 immunostaining in confirming the ovarian origin of metastatic carcinoma in serous effusion specimens. Am J Clin Pathol. 137:304–309. 2012. View Article : Google Scholar : PubMed/NCBI

32 

Szczepanek-Parulska E, Szaflarski W, Piątek K, Budny B, Jaszczyńska-Nowinka K, Biczysko M, Wierzbicki T, Skrobisz J, Zabel M and Ruchała M: Alternative 3′ acceptor site in the exon 2 of human PAX8 gene resulting in the expression of unknown mRNA variant found in thyroid hemiagenesis and some types of cancers. Acta Biochim Pol. 60:573–578. 2013.

33 

Kalnina Z, Zayakin P, Silina K and Linē A: Alterations of pre-mRNA splicing in cancer. Genes Chromosomes Cancer. 42:342–357. 2005. View Article : Google Scholar : PubMed/NCBI

34 

Ohta M, Inoue H, Cotticelli MG, Kastury K, Baffa R, Palazzo J, Siprashvili Z, Mori M, McCue P, Druck T, et al: The FHIT gene, spanning the chromosome 3p14.2 fragile site and renal carcinoma-associated t(3;8) breakpoint, is abnormal in digestive tract cancers. Cell. 84:587–597. 1996. View Article : Google Scholar : PubMed/NCBI

35 

Bednarek AK, Keck-Waggoner CL, Daniel RL, Laflin KJ, Bergsagel PL, Kiguchi K, Brenner AJ and Aldaz CM: WWOX, the FRA16D gene, behaves as a suppressor of tumor growth. Cancer Res. 61:8068–8073. 2001.PubMed/NCBI

36 

Stapleton P, Weith A, Urbánek P, Kozmik Z and Busslinger M: Chromosomal localization of seven PAX genes and cloning of a novel family member, PAX-9. Nat Genet. 3:292–298. 1993. View Article : Google Scholar : PubMed/NCBI

37 

Karolchik D, Barber GP, Casper J, Clawson H, Cline MS, Diekhans M, Dreszer TR, Fujita PA, Guruvadoo L, Haeussler M, et al: The UCSC Genome Browser database: 2014 update. Nucleic Acids Res. 42(D1): D764–D770. 2014. View Article : Google Scholar :

38 

Hecht F, Ramesh KH and Lockwood DH: A guide to fragile sites on human chromosomes. Cancer Genet Cytogenet. 44:37–45. 1990. View Article : Google Scholar : PubMed/NCBI

39 

IJdo JW, Baldini A, Wells RA, Ward DC and Reeders ST: FRA2B is distinct from inverted telomere repeat arrays at 2q13. Genomics. 12:833–835. 1992. View Article : Google Scholar : PubMed/NCBI

40 

Cheung L, Messina M, Gill A, Clarkson A, Learoyd D, Delbridge L, Wentworth J, Philips J, Clifton-Bligh R and Robinson BG: Detection of the PAX8-PPARγ fusion oncogene in both follicular thyroid carcinomas and adenomas. J Clin Endocrinol Metab. 88:354–357. 2003. View Article : Google Scholar : PubMed/NCBI

41 

Solinas-Toldo S, Dürst M and Lichter P: Specific chromosomal imbalances in human papillomavirus-transfected cells during progression toward immortality. Proc Natl Acad Sci USA. 94:3854–3859. 1997. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

July-2016
Volume 49 Issue 1

Print ISSN: 1019-6439
Online ISSN:1791-2423

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
López-Urrutia E, Pedroza-Torres A, Fernández-Retana J, De Leon DC, Morales-González F, Jacobo-Herrera N, Peralta-Zaragoza O, García-Mendez J, García-Castillo V, Bautista-Isidro O, Bautista-Isidro O, et al: PAX8 is transcribed aberrantly in cervical tumors and derived cell lines due to complex gene rearrangements. Int J Oncol 49: 371-380, 2016
APA
López-Urrutia, E., Pedroza-Torres, A., Fernández-Retana, J., De Leon, D.C., Morales-González, F., Jacobo-Herrera, N. ... Pérez-Plasencia, C. (2016). PAX8 is transcribed aberrantly in cervical tumors and derived cell lines due to complex gene rearrangements. International Journal of Oncology, 49, 371-380. https://doi.org/10.3892/ijo.2016.3515
MLA
López-Urrutia, E., Pedroza-Torres, A., Fernández-Retana, J., De Leon, D. C., Morales-González, F., Jacobo-Herrera, N., Peralta-Zaragoza, O., García-Mendez, J., García-Castillo, V., Bautista-Isidro, O., Pérez-Plasencia, C."PAX8 is transcribed aberrantly in cervical tumors and derived cell lines due to complex gene rearrangements". International Journal of Oncology 49.1 (2016): 371-380.
Chicago
López-Urrutia, E., Pedroza-Torres, A., Fernández-Retana, J., De Leon, D. C., Morales-González, F., Jacobo-Herrera, N., Peralta-Zaragoza, O., García-Mendez, J., García-Castillo, V., Bautista-Isidro, O., Pérez-Plasencia, C."PAX8 is transcribed aberrantly in cervical tumors and derived cell lines due to complex gene rearrangements". International Journal of Oncology 49, no. 1 (2016): 371-380. https://doi.org/10.3892/ijo.2016.3515