Abstract
Chromosomal translocations in cancer as well as benign neoplasias typically lead to the formation of fusion genes. Such genes may encode chimeric proteins when two protein-coding regions fuse in-frame, or they may result in deregulation of genes via promoter swapping or translocation of the gene into the vicinity of a highly active regulatory element. A less studied consequence of chromosomal translocations is the fusion of two breakpoint genes resulting in an out-of-frame chimera. The breaks then occur in one or both protein-coding regions forming a stop codon in the chimeric transcript shortly after the fusion point. Though the latter genetic events and mechanisms at first awoke little research interest, careful investigations have established them as neither rare nor inconsequential. In the present work, we review and discuss the truncation of genes in neoplastic cells resulting from chromosomal rearrangements, especially from seemingly balanced translocations.
Keywords: Chromosome aberration, translocation, gene truncation, out-of-frame fusion transcript, review
Translocations are genetic rearrangements in which chromosomes exchange material. Translocations may be balanced (no net loss or gain of genetic material occurs) or unbalanced. They sometimes involve more than two chromosomes in chains of exchanges or are complex in even more intricate ways, and they may coexist with additional numerical as well as structural genetic aberrations such as inversions, insertions, and deletions.
Acquired chromosome translocations are common in neoplasias, benign as well as malignant, testifying to their pathogenetic importance in such processes (1,2). They may be found as sole cytogenetic aberrations or as a part of a complex karyotype, in which case they may be secondary cytogenetic events. In the last update (June 6, 2022) of the “Mitelman Database of Chromosome Aberrations and Gene Fusions in Cancer”, 72.718 karyotypically abnormal cases were listed, of which 40.981 (56%) had one or more translocations (2).
Chromosomal translocations in cancer may lead to fusion genes when the breakpoints are in gene loci; in fact, translocations are the most common mechanism whereby such hybrids are generated (1,3). The swapping of chromosome segments results in joining of material from the loci in such a manner that a chimera is formed consisting of the head (5′-end) of one gene and the tail (3′-end) of the other. The new chimera acts as a single unit affecting various cellular processes, and it may be tumorigenic (1). The chimeric gene may encode a protein when two protein-coding regions are fused in-frame. The resulting protein has functional domains derived from each of the two original proteins and may be of importance early in cancer development (3-6). A chimera may also demonstrate promoter swapping when the translocation breakpoints occur in the 5′, non-coding end of the genes. This leads to deregulation of the involved genes, again possibly resulting in oncogenesis (7-15). In B- and T-lineage lymphatic leukemias and lymphomas, one partner of the fusion gene typically codes for an immunoglobulin or T-cell receptor; the other partner gene becomes quantitatively deregulated when it comes under the control of elements whose normal function it is to maintain steady transcription of the immunoglobulin and/or T-cell receptor genes (1). The role of chromosomal translocations in generating the above-mentioned types of fusion gene was unraveled in the 1980s and has been repeatedly and extensively reviewed (1-6,16-22).
Clinically, the detection of chromosome translocations leading to the generation of fusion genes may play a key role in the diagnosis and sub-classification of cancers, may have prognostic significance, and the generated fusion genes (or their chimeric proteins) may even be the target for molecular therapies (1). t(9;22)(q34;q11) is the classic example: The translocation is pathognomonic for chronic myeloid leukemia (and Ph-positive acute leukemias), results in fusion of BCR activator of RhoGEF and GTPase (BCR on 22q11) with the ABL proto-oncogene 1, non-receptor tyrosine kinase (ABL1 on 9q34), and codes for a chimeric BCR::ABL1 tyrosine kinase which can be blocked by the tyrosine kinase inhibitor imatinib mesylate (23,24).
Apart from translocations, also other structural chromosome rearrangements such as deletions (25), inversions (26,27), insertions (28-31), and tandem duplications (32) may generate highly consequential fusion genes in the oncological context. For reasons of simplicity, in the present text they are all included in the term “chromosomal translocations”.
A much less studied consequence of chromosomal translocations in cancer is truncation of genes and the expression of their aberrant truncated proteins. The breakpoints occur in protein-coding regions but a stop codon is formed after the fusion point in the resulting transcripts. Thus, a translocation may join the 5′-end coding part of a gene with an intergenic genomic sequence, it may fuse the coding part of a 5′-end gene to the antisense strand (most often intronic) of the 3′-partner gene, or it may generate out-of-frame fusion transcripts if the reading frame of the 3’-partner gene is shifted and a stop codon formed after the fusion point in the chimeric transcript. These genetic events originally came across as only marginally important, but careful investigations have shown that gene abrogation or the generation of out-of-frame chimeras for genes such as colony stimulating factor 1 (CSF1), ETS variant transcription factor 6 (ETV6), Fos proto-oncogene, AP-1 transcription factor subunit (FOS), high mobility group AT-hook 2 (HMGA2), and RUNX family transcription factor 1 (RUNX1) is not rare (33-59). In addition, functional studies have shown that truncated genes can act oncogenically (60-66). To the best of our knowledge, such events have not been extensively reviewed which is why we here summarize existing knowledge and thinking about the role gene truncations resulting from balanced chromosome rearrangements plays in tumorigenesis.
Methodological considerations. Up until and through the 1990s, detection of rearranged/fused genes in cancer almost always began with a discovery made by chromosome banding analysis of neoplastic cells. A structural chromosome aberration would be found with breakpoints mapping to distinguishable bands on the rearranged chromosome(s). The second step towards finding a pathogenetically important gene fusion brought about by cytogenetic changes mostly involved investigation of abnormal metaphase plates by means of fluorescence in situ hybridization (FISH) techniques using probes such as yeast (YAC), bacterial (BAC), and P1 (PAC) artificial chromosomes, fosmids or cosmids looking for the smallest possible probe that spanned the chromosomal break. A third step would then follow in which molecular cloning techniques were used, typically Southern blot and various types of polymerase chain reaction (PCR) amplification, that helped localize the breakpoint more precisely and identified the rearranged genes whose fusion was the result of the chromosomal rearrangement. The above-mentioned sequential procedure was laborious, but robust and reliable and resulted in the identification of many both “canonical” and “non-canonical” cancer-associated fusions (1). The final molecular product of a “canonical” fusion gene was a chimeric protein or the detection of promoter swapping between two genes (see above). In “non-canonical” fusion genes, the chimeric transcript did not produce any chimeric protein but rather a truncated peptide encoded by exons from the 5′-end partner gene that had been fused out-of-frame with a coding sequence, alternatively an intronic or antisense sequence, from the 5′-end partner gene. Shortly after the fusion point, a stop codon had been introduced. The sequence of the 3′-end partner gene might code for a novel amino acid string not present in the normal protein produced by the 3′-end partner gene. By way of illustration, and using the above-mentioned step-by-step procedure on a case of pediatric pre-B acute lymphoblastic leukemia, a cryptic recombination of chromosome bands 12p13 and 12q13 was found resulting in fusion of the first two exons of ETV6 with a sequence from intron 1 of the bromodomain adjacent to zinc finger domain 2A gene (BAZ2A) (40). Likewise, examining an atypical chronic myeloid leukemia carrying a t(3;21)(q21;q22), Micci and co-workers (67) found that the coding sequence of the receptor-like tyrosine kinase (RYK) gene had been fused to sequences from chromosome 21 that included the ATP synthase peripheral stalk subunit OSCP (ATP5O) gene coding for a mitochondrial ATP synthase; the translocation led to truncation of the RYK gene.
However, sometimes genes are recombined through a chromosomal rearrangement without the generation of a functional gene fusion. The breakpoint on one of the chromosomes may occur within a gene locus whereas the other breakpoint lies in an intergenic region. An example of this phenomenon was the t(19;21)(q13;q22) chromosome translocation detected by Hromas and coworkers (68,69) in a radiation-associated leukemia. The translocation resulted in recombination of exon 6 of RUNX1 (also known as AML1) in 21q22 with an intergenic sequence in 19q13 they called “AMP19” (GenBank sequence AY004251.2). The aberrant RUNX1::AMP19 fusion transcript produced a truncated RUNX1 protein containing only the DNA binding domain which, consequently, inhibited the function of normal RUNX1 (69).
The introduction of next generation sequencing (NGS, also called high throughput sequencing) has changed the preferred approach to the detection of fusion genes in neoplastic cells (70-74). Isolated RNA (and less frequently DNA) from tumor samples is massively parallel sequenced whereupon the data are analyzed for fusion transcripts/genes using specific algorithms resulting in the detection of many fusion transcripts/genes (74-76). But also the NGS approach has its issues of which the considerable cost of performing extensive analyses is perhaps the least important. In a study detecting the recurrent ZC3H7::BCOR fusion gene in endometrial stromal sarcomas, analysis of RNA sequencing data using the FusionMap algorithm revealed more than 1000 possible fusion transcripts (77). In other studies, the analysis of RNA sequencing data using specific programs “detected” a large number of fusion gene/transcript candidates but failed to find the pathogenetically essential CIC::DUX and KAT6A::CREBBP fusion genes in a small round cell tumor and an acute myeloid leukemia, respectively (78,79). The above examples illustrate a profound problem when searching for pathogenetically interesting fusion transcripts by RNA sequencing: some, often many, fusion genes or transcripts detected by the available algorithms are false positives (80), whereas on other occasions false negatives are the result, i.e., the search programs fail to detect the fusion genes of interest. Admittedly, some of the detected fusion transcripts could reflect intra-tumor heterogeneity and clonal evolution shown at both the cytogenetic and molecular level (81,82), but others are simply noise. One way to identify which of the suggested or detected fusion transcripts or rearranged genes are truly important, would involve the examination (by RNA sequencing) of many samples from the same type of cancer, especially if this can be done together with examination of matched normal tissue samples (34,35,44,83). A second method is preferred by research groups competent in cancer cytogenetics and combines the use of banding cytogenetic analysis of tumor cells with RNA sequencing (74). The rationale behind the latter, two-pronged approach is that a chromosomal rearrangement giving rise to a fusion gene is sometimes seen as the only cytogenetic aberration in which case it is assumed to represent a primary oncogenic event. Thus, karyotypic information is used to focus on fusion genes that map to chromosomal breakpoints. The reality of these gene fusions is then confirmed (or ruled out) using yet other methods such as FISH, PCR, and Sanger sequencing analyses (33,45,77,84-88). This approach also has the advantage of not overlooking “non-canonical” fusion genes or rearranged genes that lead to truncated proteins (42,51,58,89-94).
Most fusion genes known today were detected using high throughput sequencing technologies. In fact, most were found as fusion transcripts by RNA sequencing analysis whereupon they, by inference, were reported as fusion genes (72,95-97). More often than not, no chromosome banding or other cytogenetic analysis, no FISH, array comparative genome hybridization (aCGH), single nucleotide polymorphism (SNP) array, Southern blot or other methodology was used to falsify or support this conclusion. As a consequence, no genome-level confirmation is at hand that fusion gene formation actually existed in these situations, i.e., there is no fool-proof evidence of structural DNA rearrangements leading to the fusion of two different genes. Prof. Mitelman states in his database that “chromosome abnormalities giving rise to gene fusions identified through RNA sequencing are by default designated as translocations (t), unless shown to arise by other types of chromosome rearrangements (del, dup, ins, inv)” (2). Thus, for neighboring genes transcribed with the same orientation - centromere to telomere or telomere to centromere - and seemingly generating chimeras, one cannot distinguish between a true fusion resulting from an interstitial deletion and a read-through chimeric transcript resulting from cis-splicing of the two neighboring genes. The necessity of using “old-fashioned” techniques in addition to NGS-type analyses in order to avoid being swamped by false positives can be gleaned from a recap of fairly well-known examples from leukemogenesis: A 90 kbp interstitial deletion in 1p33 was shown to cause fusion of the 5′-untranslated part of the STIL centriolar assembly protein (STIL) gene with the coding part of TAL bHLH transcription factor 1, erythroid differentiation factor (TAL1) (98,99). In the STIL::TAL1 fusion gene, the expression of TAL1 is controlled by the STIL promoter causing excessive production of the TAL1 protein (98,99). Beyond any doubt, the transcriptional abnormalities really correspond to DNA-level changes, something that was demonstrated long before massively parallel sequencing processing entered the scene.
The genes solute carrier family 45 member 3 (SLC45A3) and ETS Like 4 transcription factor (ELK4) are both transcribed from telomere to centromere and map 25 kbp apart on 1q32. A chimeric SLC45A3-ELK4 transcript in prostate cancer was found to be generated by cis-splicing between the two neighboring genes although no DNA-level recombination between them existed (95,100-102). Similarly, the transcriptome-level fusion of the two neighboring genes ATP5L (also known as ATP5MG, official full name: ATP synthase, H+ transporting, mitochondrial Fo complex subunit G) with KMT2A (or MLL, official full name: lysine methyltransferase 2A), both of which map to chromosome band 11q23, could not be demonstrated to be the result of any detectable deletion between them. The ATP5L-KMT2A fusion transcript therefore most likely represents a read-through artefact (meaning that no DNA-level corollary exists to the transcriptome fusion transcript) of the two neighboring genes, which are both transcribed from centromere to telomere (103,104). In the Mitelman database, both SLC45A3-ELK4 and ATP5L-KMT2A are nevertheless listed as fusion genes caused by the inferred default translocations t(1;1)(q32;q32) and t(11;11)(q23;q23), respectively (2). In these two situations and according to present-day knowledge, the altered transcriptional features of the tumor cells do not correspond to actual genomic changes.
Are gene truncations recurrent? Yoshihara et al. (105) studied 4366 primary tumor samples from 13 tumor types and identified 8695 tumor-specific fusion transcripts, out of which 2991 were in-frame and 2852 out-of-frame. Shlien et al. (106) examined 23 breast tumors and found a) 205 gene-to-gene rearrangements in the same transcriptional orientation resulting in 133 in-frame and 368 out-of- frame fusion transcripts, b) 171 gene-to-gene fusions in opposite transcriptional orientation of which 50 generated transcripts with the sense portion of the 5′-end gene fusing to the antisense strand of the 3′-partner gene, and c) 473 genomic rearrangements that joined the 3′-part of a gene to an intergenic region. They summarized their findings as “productive, stable transcription from sense-to-antisense gene fusions and gene-to-intergenic rearrangements, suggesting that these mutation classes drive more transcriptional disruption than previously suspected”. In a large studied of 9000 primary tumors from 33 different types of cancer, Vellichirammal et al. (107) reported, in addition to many canonical fusion transcripts in which both partner genes were transcribed in the sense orientation, also non-canonical fusion transcripts in which at least one partner gene was transcribed in antisense orientation. Non-canonical fusion transcripts were found in all tumor groups except glioblastomas which carried only canonical fusion transcripts.
Johansson et al. (108) reviewed thousands of published fusion genes/transcripts extracted from the “Mitelman Database of Chromosome Aberrations and Gene Fusions in Cancer” concluding that most fusion genes detected by next generation sequencing reflected stochastic events of no pathogenetic significance. We see this as indicating that some of the database’s “fusion genes” actually represent artefacts attributable to bugs in the interpretation algorithms or even the technique itself. Johansson et al. further opined that there are fundamental differences between NGS-detected and “conventionally identified” fusion genes, i.e., chimeras identified or confirmed by cytogenetic methods (108).
As is apparent from the above summary, the actual status of newly reported fusion genes can be hard to establish. Some “fusions” may be noise, but at the same time it is also the case that what initially seemed to be a unique event based on scrutiny of the relevant literature at the time of reporting, in some cases turn out to be but the first of a series of similar reports. Indeed, when a fusion gene’s actual existence at the DNA level has been confirmed cytogenetically, experience shows that a seemingly unique chromosome aberration targeting a neoplasia-relevant gene will also soon be seen in other neoplasms. The publication interval between the first and second case, and then the third and so on, could be a measure of the frequency/rarity of the aberration, although pure chance certainly plays a role, too.
By way of example, we recently described a case of therapy-related myeloid leukemia carrying a t(8;19)(p11;q13) chromosome translocation resulting in fusion of the lysine acetyltransferase 6A gene (KAT6A) with the leucine twenty homeobox gene (LEUTX) (109). The first leukemia with t(8;19) had been reported in 1988 (110). The translocation remained a fluke event until 1995 when a second case was reported carrying the same translocation (111). The third leukemia case with a seemingly identical t(8;19) was reported in 2008, 20 years after the first one and 13 years after the second (112). In 2014, six years later, the molecular consequences of the generation of the KAT6A::LEUTX fusion gene were reported (113). At present (after a time-span of 33 years), only seven leukemias carrying t(8;19)(p11;q13)/KAT6A::LEUTX can be found in the relevant cytogenetic literature (109,114,115). Thus, the t(8;19)(p11;q13)/KAT6A::LEUTX fusion gene comes across as a very rare, but nevertheless unquestionably recurrent, genetic aberration in leukemia reported at a frequency of only 0.21-times per year (7 cases in 33 years).
In 2013, we combined cytogenetics with RNA sequencing methodology to identify a fusion of the zinc finger MYND-type containing 8 (ZMYND8) gene, mapping on chromosome subband 20q13.12, with the RELA proto-oncogene, NF-kB subunit (RELA) gene that maps to 11q13.1, in a pediatric acute erythroid leukemia (116). In 2019, Iacobucci and coworkers, examining 159 acute erythroid leukemias using next generation sequencing methodologies, found the second case of pediatric acute erythroid leukemia with the same ZMYND8::RELA chimera, indicating that ZMYND8::RELA is a rare but recurrent fusion in pediatric acute erythroid leukemias (117).
Many chromosome translocations resulting in truncated or fused genes leading to out-of-frame transcripts have been reported only once. These events might turn out to be recurrent if and when the second case, and then the third et cetera, is reported, as for the leukemias with t(8;19)/KAT6A::LEUTX described between 1988 and 1995 and t(11;20)/ZMYND8::RELA between 2013 and 2018 (33,37,38,40,53,67,89-94,118-121). In 1985, De Braekeleer and coworkers reported an acute lymphoblastic leukemia carrying a t(10;19)(q26;q13) as the sole cytogenetic change (122). Thirty-two years later, we described the second case of acute lymphoblastic leukemia with t(10;19)(q26;q13) (89). Molecular analysis demonstrated that the coding sequence of the FAM53B gene (family with sequence similarity 53 member B) from 10q26 was fused to a genomic sequence from 19q13 that mapped upstream of the solute carrier family 7 member 10 (SLC7A10). Because no molecular analyses had been performed on the case from 1985 (122), we do not know whether the two reported translocations were identical at the molecular level. However, both patients had t(10;19)(q26;q13) as the only cytogenetic change, both were diagnosed with acute lymphoblastic leukemia, and both achieved complete hematologic and cytogenetic remission under antileukemic treatment (89,122).
In 2019, our group reported the first genetically analyzed myoid hamartoma of the breast (123). The tumor had a t(5;12)(p13;q14) translocation as the sole karyotypic aberration leading to fusion of the HMGA2 gene from 12q14 with a sequence from chromosome subband 5p13.2. The data indicated that myoid hamartoma is a true neoplasm growing from a mutated mesenchymal stem cell capable of differentiating into smooth muscle cells (123). In 2022, we reported the genetic characterization of a mammary leiomyomatous tumor without signs of malignancy (124). Using cytogenetics, fluorescence in situ hybridization, RNA sequencing, reverse transcription-polymerase chain reaction, and Sanger sequencing methodologies, we showed that the tumor cells had an identical genetic profile to that found in the previously examined myoid breast hamartoma. We conclude that a chromosome translocation t(5;12)(p13;q14) resulting in fusion of HMGA2 with a sequence from chromosome subband 5p13.2 is recurrent in benign myoid neoplasms of the breast (124).
The recurrent t(2;12)(q35~37;q14~q15) chromosome translocation in lipomas results in fusion of HMGA2 with the atypical chemokine receptor 3 (ACKR3) gene (also known as RDC1) from 2q37.3. The functional impact of the fusion was assumed to be truncation of HMGA2 since ACKR3 contributes a stop codon one amino acid downstream of the fusion point (47,125). Similarly, the recurrent chromosome translocation t(9;12)(p22;q14~q15) found in lipoma, pleomorphic adenoma, and myolipoma results in fusion of HMGA2 with the nuclear factor I B (NFIB) gene from 9p23-p22.3 (lipomas and pleomorphic adenomas) or with chromosome 9 open reading frame 92 (C9orf92) from 9p22, leading to HMGA2 truncation (50,51,126,127). On the other hand, the recurrent t(4;12)(q27~28;q14~15) chromosomal rearrangement in lipoma results in fusion of HMGA2 not with a gene but with various intergenic sequences from chromosome subband 4q28.1, again leading to truncation of HMGA2 (46). The above examples illustrate how chromosome translocations resulting in truncation of a gene may be viewed as recurrent variations on the same pathogenetic theme. Indeed, the variety of chromosome aberrations, most of them translocations, targeting the HMGA2 gene locus in many types of neoplasia are the best-known examples of the fascinating genetic heterogeneity that may lie behind gene truncation as a tumorigenic mechanism (Table I) (128-130). Chromosomal rearrangements leading to truncations also target the CSF1 gene which maps to 1p13.3 in tenosynovial giant cell tumors (33-35), the collagen type II alpha 1 chain gene (COL2A1) which maps to 12q13.11 in chondrosarcomas (131,132), FOS which maps to 14q24.3 in epithelioid hemangioma of the bone and osteoblastoma (42-45,133), the trio Rho guanine nucleotide exchange factor (TRIO) gene which maps to 5p15.2 in liposarcomas and undifferentiated sarcomas (59,134,135), as well as several other loci.
Table I. Genomic aberrations giving rise to HMGA2 (on 12q14) truncation or in-frame fusions (denoted with *).
In hematologic neoplasias, chromosome rearrangements generating truncated or out-of-frame fusion transcripts have been frequently described to target RUNX1 in 21q22.12 (53,55-58,68,69,120,136-142). In most cases, RUNX1 was truncated after exon 6 or 7 (Table II). Translocations generating truncations have also been frequently reported for ETV6 in 12p13.2 (Table III) (36-41,143-145). In all examined cases, the truncation of ETV6 occurred after exon 1 or 2 of the gene (Table III).
Table II. Genomic aberrations giving rise to truncated RUNX1 (on 21q22).
MDS: Myelodysplasia; AML: acute myeloid leukemia; ALL: acute lymphoblastic leukemia; EMS: 8p11 myeloproliferative syndrome.
Table III. Genomic aberrations giving rise to truncated ETV6 (on 12p13).
ALL: Acute lymphoblastic leukemia; AEL: acute erythroid leukemia; RAEB: refractory anemia with excess blasts; AML: acute myeloid leukemia; MDS: myelodysplasia/myelodysplastic syndrome; CMML: chronic myelomonocytic leukemia; MPN-U: myeloproliferative neoplasm, unclassifiable.
Also the lysine methyltransferase 2A (KMT2A, formerly known as MLL) gene on 11q23.3 sometimes fuses with intergenic sequences of various chromosomes giving rise to out-of-frame KMT2A-fusion transcripts (146-149). Based on the MLL recombinome data, five chromosome translocations exist that result in fusion of the 5′-part of KTM2A with genomic sequences from 1p13, 6q27, 9p13, 11q23, 11q24, and 21q22 (148,149) whereas ten chromosome translocations result in out-of-frame KMT2A fusion genes (146-149).
Are gene truncations pathogenetic events (primary or secondary) or genetic noise? In cancer cytogenetics, the terms primary and secondary chromosome aberrations are frequently used to emphasize the sequential order in which neoplastic cells acquire genomic changes (150-152). Primary chromosome abnormalities are believed to be important in the transformation of susceptible somatic target cells and, hence, essential in the very establishment of a neoplasm. They may be, in fact often are, observed as sole cytogenetic changes and are not rarely phenotype-specific (150-152). Many gene truncations caused by primary chromosome aberrations, mainly translocations, are found in both solid tumors and leukemias; perhaps the most typical and yet variable among them involves the HMGA2 gene (47,128-130,153-161). Other examples of primary aberrations leading to gene truncations in solid tumor cytogenetics target CSF1 in 1p13.3 (33-35) in tenosynovial giant cell tumors and FOS in 14q24.3 in epithelioid hemangiomas of bone and osteoblastomas (42-45,133).
The genotype-phenotype or cytogenetic-diagnostic specificity is far from absolute, however, inasmuch as the same primary chromosome abnormality is sometimes found in more than one tumor type indicating shared pathogenetic pathways. For example, the t(16;21)(p11;q22) translocation leading to fusion of FUS RNA binding protein gene (FUS) from 16p11 with the ETS transcription factor ERG gene (ERG) from 21q22 to form a FUS::ERG hybrid gene on the der(21)t(16;21)(p11;q22), has been reported in acute leukemias (162-166) as well as Ewing sarcomas (167,168). The chromosome translocation t(3;12)(q28;q14) which fuses HMGA2 with the LPP gene (in 3q28; the gene’s full official name is LIM domain containing preferred translocation partner in lipoma) has been reported in lipoma, chondroid hamartoma, and chondroma (154,156,169-171). Likewise, truncation of the TYRO3 protein tyrosine kinase (TYRO3) gene on 15q15.1 through a t(10;15)(p11;q15) has been found in pediatric acute myeloid leukemia (93) as well as anaplastic ependymoma grade 3 (172). In leukemia, an exonic sequence of TYRO3 is fused to an intergenic sequence from 10p11 (93). In ependymoma, an exonic sequence of TYRO3 is fused out-of-frame with the kelch like family member 18 (KLHL18) gene from 3p21.31 (172). In both neoplasias, the putative TYRO3 truncated protein would contain the extracellular domain together with the two immunoglobulin domains, the two fibronectin type III domains, and the transmembrane domain. The protein lacks the catalytic domain of TYRO3, its autophosphorylation sites, and the carboxyl-terminal part, all of which are required to maintain TYRO3 stability (93,172).
That a cytogenetic aberration is solitary does not necessarily mean that it was the initial genetic event in neoplastic transformation since also submicroscopic genetic aberrations may be present. For example, molecular and fluorescence in situ hybridization analyses of a pediatric AML carrying trisomy 4 as the only cytogenetically visible aberration revealed the additional presence of both an FLT3-ITD mutation and a cryptic t(6;21)(q25;q22) translocation resulting in fusion of exon 7 of RUNX1 with an intergenic sequence from 6q25 (58). Which of these three changes was the initial genetic event, let alone the most pathogenetically important one, remains unknown (58). On the other hand, in a patient with myeloproliferative bone marrow disease, 90% of the cytogenetically analyzed metaphases carried a t(12;22)(q14.3;q13.2) as the sole cytogenetic aberration. The aberration generated an HMGA2::EFCAB6 fusion gene (EFCAB6 is from 22q13; the official name is EF-hand calcium binding domain 6) resulting in HMGA2 truncation (173). An additional JAK2V617F mutation was also found but only in half of the cells examined, indicating that the t(12;22)(q14.3;q13.2) leading to HMGA2::EFCAB6 was indeed the primary genetic event in the neoplastic process (173).
The secondary chromosome aberrations referred to above were acquired after the primary ones but probably were of importance in tumor progression (150-152). They, too, come across as nonrandom genetic events that may be clinically important both diagnostically and prognostically (152,174-185). By way of example, a patient with chronic myeloproliferative disorder carried a t(8;22)(p11;q11) leading to fusion of the BCR activator of RhoGEF and GTPase (BCR) and fibroblast growth factor receptor-1 (FGFR1) genes (55,186). At the time of transformation of the myeloproliferative disorder to AML, the cells had acquired an additional t(9;21)(q34;q22) [the karyotype was now 46,XY,t(8;22)(p11;q11),t(9;21)(q34;q22)] (55). The t(9;21)(q34;q22) resulted in fusion of exon 4 of RUNX1 with repetitive sequences from chromosome 9 generating a truncated RUNX1 gene (55). In an acute leukemia characterized by the pathognomonic t(9;22)(q34;q11.2)/BCR::ABL1 aberration, an additional t(15;21)(q26.1;q22) was also detected that had resulted in fusion of RUNX1 with the antisense strand of the SV2B gene, leading to formation of out-of-frame RUNX1 fusion transcripts and the production of truncated RUNX1 isoforms (57). Finally, an AML patient carrying a t(5;18)(q35;q21) as the primary cytogenetic abnormality developed a t(3;12) (p23;p13) during disease progression resulting in ETV6 rearrangement (187).
The term cytogenetic noise is used for the background level of nonconsequential aberrations, seemingly distributed in a random manner throughout the genome, that is seen in some malignancies (150,151). In contrast to primary and secondary cytogenetic aberrations that confer an evolutionary edge on the cells, and are detected in at least two cells (the minimum prerequisite for regarding them as clonal), noise aberrations are typically nonclonal. Only limited investigations on the oncogenic importance of such aberrations (if any) have been reported (188-192). Extensive cytogenetic instability has been described in elastofibromas (193,194) whereas nonclonal telomeric associations are common in giant cell tumors of bone (195-197). No studies have focused on possible associations between nonclonal structural chromosome aberrations in cancer and fusion genes that may have been formed by them. However, some of the multiple fusion transcripts detected when analyzing RNA sequencing data using fusion transcript detection programs (FusionCatcher, TopHat, JAFFAL, et cetera) might be attributable to nonclonal structural chromosome aberrations. Single cell RNA/DNA sequencing methodology might be a useful tool to assess a possible relationship between nonclonal structural chromosome aberrations and fusion genes in cancer (198-204). Because, at least in theory, sequencing may detect fusion genes at the level of single cells, this approach should be useful in studies of intratumor heterogeneity, enabling the detection of fusion genes and possible alternative fusion transcript variants in even the smallest of clones (198-204).
Are gene truncations drivers? In the era of next generation sequencing, acquired genetic abnormalities detected by this methodology are often referred to as either drivers or passengers. Driver mutations cause cancer (or at least neoplasia) inasmuch as they play a fundamental role in the transformation of a targeted somatic cell, whereas passenger mutations are randomly accumulated events that seemingly do not influence tumorigenesis one way or another (205-207). A third type of mutations called “mini drivers” have relatively weak tumor-promoting effects but are nevertheless clinically important at least to some degree (208,209). By and large, driver mutations have the same status as what we prefer to call primary genomic aberrations.
Against this background it is not surprising that evidence is mounting that some truncated genes are driver mutations in cancer. The best-known example is rearrangements leading to truncation of the HMGA2 gene, events referred to extensively above. HMGA2 consists of five exons and maps to chromosome subband 12q14.3 (chr12:65,824,483-65,966,291 in GRCh38/hg38 assembly) (153-156). The first three exons code for three domains which specifically bind to the minor groove of adenine-thymine (AT) rich DNA, hence the name AT-hooks DNA binding domains (155,210-212). Exon 5 contains a long untranslated region (3′-UTR) that has seven microRNA let-7 sites and many other regulatory sequences and controls the expression of HMGA2 (Figure 1) (213-216).
Figure 1. The high mobility group AT-hook 2 (HMGA2) gene. The chromosome band location (top) and exon 5 of the gene (bottom) are shown. On exon 5, the Let7 positions are underlined. Additional microRNA target sites predicted by TargetScanHuman 7.2 are also shown. The colors indicate different classes of target sites: Purple is for 8mer, red for 7mer-m8 and blue for 7mer-A1.
Recombination of 12q14.3 with bands from almost every other chromosome, but resulting in truncation of HMGA2, has repeatedly been reported in various neoplasms (2,129) (Table III). The molecular consequence is always the physical separation of the first three exons from the 3′-UTR of HMGA2. In vitro experiments showed that expression of truncated HMGA2 protein carrying the three AT-hook DNA-binding domains transforms mouse embryonic NIH3T3 fibroblasts (62), induces leiomyoma-like lesions in human myometrial cells (63), and promotes the growth of chondrocytes (64,217). Transgenic mice producing a truncated form of HMGA2 develop benign mesenchymal tumors (65,218).
In later years, other genes were also deemed to have driver potential when truncated through genomic rearrangements; they, too, have been at the forefront of pathogenetic interest for those who study the genomic or chromosomal abnormalities in neoplasia. In cells from tenosynovial giant cell tumors, various chromosome rearrangements, mainly translocations, often target and disrupt the CSF1 locus in 1p13.3 (chr1:109,910,849-109,930,992 in GRCh38/hg38 assembly) (219-227). CSF1 has four alternative transcripts (references sequences NM_172210.3, NM_172212.3, NM_172211.4, and NM_00757.6) which are different only in exons 6 and 9 (Figure 2) (228,229). The translation initiation codon is in exon 1 whereas the stop codon is in exon 8. CSF1 transcripts with reference sequences NM_172211.4 and NM_00757.6 have the same untranslated exon 9, while transcript NM_172212.3 differs in exon 9 compared to the other two and transcript NM_172210.3 lacks untranslated exon (Figure 2). No information is available about exon 9 of the CSF1 transcript with reference sequence NM_172212.3 whereas a lot is known about exon 9 of the two CSF1 transcripts with reference sequences NM_172211.4 and NM_00757.6. Exon 9 contains at least 14 microRNA target sites (miRNA) that regulate the expression of CSF1 and its protein, a noncanonical G-quadruplex which is involved in post-transcriptional regulation of CSF1, as well as AU-rich elements (AREs) which regulate the stability and decay of CSF1 mRNA (230-232). Woo et al. (232) showed that disruption of the miRNA target region, G-quadruplex, and AREs together dramatically increased reporter RNA levels, suggesting important roles for these cis-acting regulatory elements in the down-regulation of CSF1 mRNA.
Figure 2. The colony stimulating factor 1 (CSF1) gene. Chromosome band location (top), the four transcript variants (middle), and the last nontranslated exon 9 of the variants with accession number NM_172211.4 and NM_000757.6 (bottom) are shown. On exon 9, the G-quadruplex, the AU-rich (ARE) element, and microRNA target sites predicted by TargetScanHuman 7.2 are also shown. The colors indicate different classes of target sites: Purple is for 8mer, red for 7mer-m8 and blue for 7mer-A1.
Studying three tenosynovial giant cell tumors, Panagopoulos and coworkers (33) found that replacement of the 3′-UTR (exon 9) of CSF1 with new sequences was common to all three tumors resulting in overexpression of the protein-coding part of CSF1. That this is indeed a common pathogenetic theme in tenosynovial giant cell tumors was confirmed by two subsequent studies of larger series of tumors (34,35). Abnormal expression of CSF1 by macrophages and monocytes led to tumor development (225,226). The possible role of CSF1R tyrosine kinase inhibitor as a medical treatment for patients with tenosynovial giant cell tumors is under investigation (233-236).
Epithelioid hemangiomas of the bone and osteoblastomas are genetically characterized by rearrangements leading to fusions of the FOS gene which maps to chromosome subband 14q24.3 (chr14:75,278,828-75,282,230 in GRCh38/hg38 assembly) (42,44,45,133). FOS has four exons and codes for a leucine zipper protein which dimerizes with members of the JUN family of transcription factors to form the transcription factor complex AP-1 (237,238). Exon 4, the last exon of FOS, consists of a coding part and a 3′-UTR (Figure 3). The C-terminal region of the translated part of exon 4 (last 21 amino acids: KGSSSNEPSSDSLSSPTLLAL) is essential for the degradation of FOS protein (239,240). The 3′-UTR contains ARE and many micro-RNA target sites which regulate the expression of FOS and maintains the stability of FOS mRNA, and a 145 bp nucleotide region which enables the targeting of FOS mRNA towards the perinuclear cytoskeleton (241-253).
Figure 3. The Fos proto-oncogene, AP-1 transcription factor subunit (FOS) gene. On the translated part of exon 4, the C-terminal region involved in FOS degradation is shown. In the untranslated part of the exon, the AU-rich (ARE) element and microRNA target sites predicted by TargetScanHuman 7.2 are shown. The colors indicate different classes of target sites: Purple is for 8mer, red for 7mer-m8 and blue for 7mer-A1. In all FOS chimeras, the FOS degradation and the untranslated region are not present.
In all examined epithelioid hemangiomas of the bone and osteoblastomas with rearrangements of FOS, the breakpoint occurred in the coding part of exon 4 where the fusion event introduced a stop codon (42-45,61). As a consequence, the C-terminal part of FOS, essential for its degradation, and the 3’-UTR of FOS mRNA are removed (Figure 3). The resulting, truncated FOS protein contains the N-terminal transactivation domain which plays a crucial role in transformation, and the basic leucine zipper domain (bZIP) that makes it more stable (61,242). In vitro experiments have shown that the truncated FOS protein is resistant to degradation and has a longer half-life than wild-type FOS (1-2 hrs) (61).
The mitogen-activated protein kinase 8 (MAP3K8) gene, which maps to chromosome subband 10p11.23 (chr10:30,434,021-30,461,833 in GRCh38/hg38 assembly), has nine exons (NCBI reference sequence: NM_005204.4) (Figure 4) and codes for a cytoplasmic protein which is a member of the serine/threonine protein kinase family and activates both the MAP and JNK kinase pathways (254-256). In 1991, Miyoshi and coworkers cloned MAP3K8 (257) and named it cot (cancer Osaka thyroid) oncogene. They found two cDNA clones: one corresponding to cDNA of the normal MAP3K8, and a second, truncated clone in which part of the sequence from the eighth exon onwards was replaced by another sequence. They wrote that “…DNA sequence downstream from the 3’ rearranged point was totally different from the cot proto-oncogene sequence and the following coding sequence of the cot oncogene was abruptly terminated by a stop codon TAG. These results suggest that the cot oncogene might be activated by a C’-terminal truncation during the process of DNA transfer, although we have no idea about the C’ terminus of the normal Cot kinase because we have no information about the normal transcript of the cot proto-oncogene” (257). In 1993, Aoki and coworkers (258) showed that the truncated gene had much higher transformation activity than the normal one and suggested that the N-terminal part of MAP3K8 protein is essential for transformation whereas the C-terminal moiety of normal MAP3K8 protein (i.e., the part missing in the truncated gene product) may regulate negatively the transformation potential of the MAP3K8 protein (258).
Figure 4. The mitogen-activated protein kinase 8 (MAP3K8) gene. The chromosome band location (top) and exon 9 of the gene are shown (bottom). On the translated part of exon 9, the C-terminal region the kinase repression domain is shown. The underlined sequence is part of the PEST domain of MAP3K8 protein. The bold sequence is the degradation signal (degron).
For the purpose of the present review, we aligned the sequence reported in that article with the Human GRCh38/hg38 assembly using the BLAT algorithm (259) (CCAAGAGCCGCAGACCTACTAAAACATGAGGCCCTGAACCCGCCCAGAGAGGATCAGCCACGC-GGGCACCAAGTCATTCATGAAGGATCCTCCACCAATGACCCAAACAACTCCTGCTAGGCCCCACCT). It turned out that, in the truncated MAP3K8 cDNA clone, exon 8 of the gene (the fusion point was at nucleotide 1640 of the NCBI reference sequence: NM_005204.4) was fused with an intergenic sequence mapping to chromosome subband 5q13.3 (chr5:75,004,416-75,004,477).
In 2004, Clark and coworkers (260) reported a novel MAP3K8 truncation in lung cancer. The breakpoint occurred in exon 8 of the gene (the fusion point was at nucleotide 1711 of the NCBI reference sequence: NM_005204.4) and fused MAP3K8 with an intergenic sequence from chromosome band 9p23 (260). Truncated MAP3K8 had much higher transforming activity than wild-type MAP3K8 (260). The common theme of those two truncated MAP3K8s was that they lacked exon 9 which codes for the last 43 amino acids of the MAP3K8 protein (position 425-467 in reference sequence NP_005195.2: DSSCTGSTEESEMLKRQRSLYIDLGALAGYFNLVRGPPTLEYG) (Figure 4). This part of the protein is a kinase repression domain, has a degradation signal (degron) between positions 435 and 457 (SEMLKRQRSLYIDLGALAGYFNL), and a portion of the PEST sequence (DSSCTGSTEESEML) which acts as a signal for MAP3K8 degradation (261). Thus, the absence of the last 43 amino acids in the truncated MAP3K8 results in stabilization of the protein, higher kinase activity, and higher oncogenic capacity (261,262).
In 2019, Newman and coworkers found recurrent rearrangements of MAP3K8 in spitzoid melanomas (263). Working from an index case (an 11-year-old boy) and studying altogether 50 spitzoid melanomas, they found MAP3K8 fusion genes in 12 tumors and truncation of MAP3K8 in another five. The common theme in all 17 tumors was the absence of exon 9 of MAP3K8 (263). When truncated MAP3K8, which lacked exon 9, was expressed in NIH 3T3 cells, signs of neoplastic transformation were seen in a colony-formation assay (263). Truncations and fusions of the MAP3K8 gene have subsequently been detected in spitzoid melanomas by other researchers (237-239) and similar changes have been seen also in other melanoma subtypes (264-267). In all these settings, lack of exon 9 was the common pathogenetic theme.
In hematologic malignancies, most reported RUNX1 truncations are the consequence of chromosome translocations hitting the gene after exon 6 or 7 (Table II). These truncated genes code for a RUNX1 protein containing the Runt homology domain (RHD) but lacking the transcriptional activation domain. The truncated RUNX1 proteins may heterodimerize with CBFB and bind to DNA but do not have any transactivation ability (53,56,58,69,120,136,137,139-142,268). Thus, they are functionally similar to the isoform AML1a of the RUNX1 protein (NCBI reference sequences NM_001122607.2 and NP_001116079.1) (58,269-271). AML1a binds to the DNA sequence of target genes and inhibits the transcriptional activity of RUNX1 (271). In 32Dcl3 murine myeloid cells treated with colony stimulating factor 3 (CSF3), overexpression of AML1a suppressed granulocytic differentiation but stimulated cell proliferation (271). Overexpression of AML1a has been found in both ALL and AML-M2 patients (272). Mice have developed lymphoblastic leukemia after transplantation of murine bone marrow mononuclear cells transduced with AML1a (272).
Truncated RUNX1 proteins, resulting from chromosome translocations, were found to function as an inhibitor of normal RUNX1 protein. They increase proliferation and disrupt the differentiation program by interfering with AML1b (53,69,120,136-138). In a recent study, truncated RUNX1 protein induced expression of CSF3 receptor on 32D myeloid leukemia cells (142).
Studying RUNX1 point mutation in a mice model, Watanabe-Okochi and coworkers found that truncated RUNX1 protein (mutant S291fsX300) induced pancytopenia with erythroid dysplasia, followed by progression to MDS-RAEB or MDS/AML (273). In human CD34+ stem/progenitor cells as well as human induced pluripotent stem cells, the same mutation impaired myeloid commitment, enhanced self-renewal, and prevented granulocytic differentiation (274).
In another mouse model study, Dowdky and coworkers introduced a premature translational stop codon after amino acid 307 (Runx1Q307X) mimicking the RUNX1 mutations found in MDS/AML and CMML patients (275). RUNX1 truncated protein bound to suitable DNA sequences but failed to activate the target gene’s promoters and was unable to associate with the nuclear matrix. Furthermore, Runx1Q307X homozygous mice exhibited embryonic lethality at E12.5 due to central nervous system hemorrhage and a complete lack of hematopoietic stem cell function (275). Taking this and the other above-mentioned data into consideration, it seems safe to conclude that truncated RUNX1 protein is at least a contributing factor in leukemogenesis.
Recently, Kogure et al. reported that 13% of adult T-cell leukemias/lymphomas and 12% of diffuse large B-cell lymphomas carry a truncation of the 3′-end of the REL proto-oncogene, NF-kB subunit (REL) gene which maps to chromosome subband 2p16.1 (276). The consequence is upregulation of REL expression and generation of a REL protein with gain-of-function (276).
Also, functional studies of out-of-frame fusion transcripts suggest that they have the properties of drivers in neoplasia. In ovarian cancer SKOV3 cells, a fusion transcript between the genes LAMC2 from 1q25.3 (official full name: laminin subunit gamma 2) and NR6A1 (nuclear receptor subfamily 6 group A member 1) from 9q33.3 resulted in a truncated LAMC2 protein that lacked the essential functional domains I and II, but instead had tumor promoting properties (277). Overexpression and suppression of LAMC2::NR6A1 fusion transcripts significantly increased or decreased the tumorigenic growth of neoplastic cells in mouse models, respectively (277).
In a Ewing sarcoma carrying the pathognomonic EWSR1::FLI1 fusion gene, Dupain and coworkers (172,278) detected two variants of an LMO3::BORCS5 out-of-frame fusion transcript leading to loss of the functional domain LIM2 from the LMO3 gene (official full name: LIM domain only 3) and disruption of BORCS5 (official full name: BLOC-1 related complex subunit 5) (172,278). In vitro studies showed that LMO3::BORCS5 fusion transcripts increased proliferation, decreased expression of apoptosis-related genes, and deregulated genes involved in differentiation. Expression of LMO3::BORCS5 in NIH-3T3 cells and subcutaneous injection into the flank of mice induced tumors in the animals (172,278).
In a patient with myelodysplasia, the cytogenetically detected t(12;17)(p13;q11) chromosome translocation was shown to result in fusion of exon 2 of ETV6 with an antisense sequence from intron 19 of TAOK1 (from 17q11). The chimeric ETV6::TAOK1 transcript had an RNA-interfering effect on the expression of normal TAOK1 resulting in less production of TAOK1 protein. In addition, it encoded a C-terminal truncated ETV6 protein (145). In the zebrafish model, truncated forms of ETV6 had a dominant-negative effect on the function of normal ETV6 protein disrupting both primitive and definitive hematopoiesis (60).
Many of the reciprocal KMT2A fusion genes such as IKZF1::KMT2A, PBX1::KMT2A, and JAK1::KMT2A are out-of-frame (149). However, these out-of-frame fusions can express the 3-terminal part of KMT2A through an internal KMT2A promoter (149,279). The N-terminal truncated version of KMT2A protein may acetylate H4K16 through the association with MOF protein or methylate H3K4 by the SET domain complex (149,279).
Conclusion
Although gene truncations or out-of-frame fusion transcripts brought about by chromosome translocations and other cytogenetic rearrangements have received relatively little emphasis in studies of neoplastic transformation, data are accumulating that such changes may play important roles in tumorigenesis and leukemogenesis. It is worthy of emphasis that next generation sequencing technologies used alone - without cytogenetic backing and/or the use of other molecular genetic techniques for falsification/verification purposes - are not sufficient when it comes to detecting such genetic aberrations; many false positives are detected while at the same time pathogenetically important changes may be overlooked. Gene truncations/out-of-frame fusion transcripts may remove transcriptional and translational regulatory elements affecting gene expression and/or the stability and properties of the encoded protein. On other occasions, they may introduce a premature stop codon whose effect may be similar to that of a nonsense pathogenetic polymorphism. Apart from the highly prevalent truncation of HMGA2 in mostly benign connective tissue tumors, whole genome or whole transcriptome sequencing of large series of a variety of other tumors has detected gene truncations/out-of-frame fusion transcripts that at present come across as rare, but that later may be shown to be recurrent, perhaps even common. Examples are truncations of CSF1 in tenosynovial giant cell tumors, of FOS in epithelioid hemangiomas of the bone and osteoblastomas, of MTRK8 in spitzoid melanomas, and of REL in T-cell leukemias/lymphomas and diffuse large B-cell lymphoma. The search for other neoplasia-inducing acquired genetic alterations of this type should be intensified, both by cytogenetic and molecular genetic means.
Conflicts of Interest
The Authors declare that they have no potential conflicts of interest with regards to this study.
Authorsʼ Contributions
Both Authors (IP and SH) wrote the manuscript.
Acknowledgements
The research work of the authors was supported by grants from Radiumhospitalets Legater.
References
- 1.Heim S, Mitelman F. Wiley-Blackwell. 2015. Cancer cytogenetics: Chromosomal and molecular genetic abberations of tumor cells. Fourth Edition edn. [Google Scholar]
- 2.Mitelman F, Johansson B, Mertens F. Mitelman database of chromosome aberrations and gene fusions in cancer. 2022. Available at: https://mitelmandatabase.isb-cgc.org/ [Last accessed on June 26, 2022]
- 3.Mitelman F, Johansson B, Mertens F. The impact of translocations and gene fusions on cancer causation. Nat Rev Cancer. 2007;7(4):233–245. doi: 10.1038/nrc2091. [DOI] [PubMed] [Google Scholar]
- 4.Rabbitts TH. Chromosomal translocations in human cancer. Nature. 1994;372(6502):143–149. doi: 10.1038/372143a0. [DOI] [PubMed] [Google Scholar]
- 5.Rowley JD. The critical role of chromosome translocations in human leukemias. Annu Rev Genet. 1998;32:495–519. doi: 10.1146/annurev.genet.32.1.495. [DOI] [PubMed] [Google Scholar]
- 6.Xiao X, Garbutt CC, Hornicek F, Guo Z, Duan Z. Advances in chromosomal translocations and fusion genes in sarcomas and potential therapeutic applications. Cancer Treat Rev. 2018;63:61–70. doi: 10.1016/j.ctrv.2017.12.001. [DOI] [PubMed] [Google Scholar]
- 7.Kas K, Voz ML, Röijer E, Aström AK, Meyen E, Stenman G, Van de Ven WJ. Promoter swapping between the genes for a novel zinc finger protein and beta-catenin in pleiomorphic adenomas with t(3;8)(p21;q12) translocations. Nat Genet. 1997;15(2):170–174. doi: 10.1038/ng0297-170. [DOI] [PubMed] [Google Scholar]
- 8.Voz ML, Aström AK, Kas K, Mark J, Stenman G, Van de Ven WJ. The recurrent translocation t(5;8)(p13;q12) in pleomorphic adenomas results in upregulation of PLAG1 gene expression under control of the LIFR promoter. Oncogene. 1998;16(11):1409–1416. doi: 10.1038/sj.onc.1201660. [DOI] [PubMed] [Google Scholar]
- 9.Van Dyck F, Declercq J, Braem CV, Van de Ven WJ. PLAG1, the prototype of the PLAG gene family: versatility in tumour development (review) Int J Oncol. 2007;30(4):765–774. [PubMed] [Google Scholar]
- 10.Oliveira AM, Hsi BL, Weremowicz S, Rosenberg AE, Dal Cin P, Joseph N, Bridge JA, Perez-Atayde AR, Fletcher JA. USP6 (Tre2) fusion oncogenes in aneurysmal bone cyst. Cancer Res. 2004;64(6):1920–1923. doi: 10.1158/0008-5472.can-03-2827. [DOI] [PubMed] [Google Scholar]
- 11.Oliveira AM, Perez-Atayde AR, Dal Cin P, Gebhardt MC, Chen CJ, Neff JR, Demetri GD, Rosenberg AE, Bridge JA, Fletcher JA. Aneurysmal bone cyst variant translocations upregulate USP6 transcription by promoter swapping with the ZNF9, COL1A1, TRAP150, and OMD genes. Oncogene. 2005;24(21):3419–3426. doi: 10.1038/sj.onc.1208506. [DOI] [PubMed] [Google Scholar]
- 12.Patel NR, Chrisinger JSA, Demicco EG, Sarabia SF, Reuther J, Kumar E, Oliveira AM, Billings SD, Bovée JVMG, Roy A, Lazar AJ, Lopez-Terrada DH, Wang WL. USP6 activation in nodular fasciitis by promoter-swapping gene fusions. Mod Pathol. 2017;30(11):1577–1588. doi: 10.1038/modpathol.2017.78. [DOI] [PubMed] [Google Scholar]
- 13.Hiemcke-Jiwa LS, van Gorp JM, Fisher C, Creytens D, van Diest PJ, Flucke U. USP6-associated neoplasms: a rapidly expanding family of lesions. Int J Surg Pathol. 2020;28(8):816–825. doi: 10.1177/1066896920938878. [DOI] [PubMed] [Google Scholar]
- 14.Panagopoulos I, Gorunova L, Andersen K, Lobmaier I, Lund-Iversen M, Micci F, Heim S. Fusion of the Lumican (LUM) gene with the ubiquitin specific peptidase 6 (USP6) gene in an aneurysmal bone cyst carrying a t(12;17)(q21;p13) chromosome translocation. Cancer Genomics Proteomics. 2020;17(5):555–561. doi: 10.21873/cgp.20211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Nakayama S, Nishio J, Aoki M, Koga K, Nabeshima K, Yamamoto T. Ubiquitin-specific peptidase 6 (USP6)-associated fibroblastic/myofibroblastic tumors: evolving concepts. Cancer Genomics Proteomics. 2021;18(2):93–101. doi: 10.21873/cgp.20244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Nowell PC, Croce CM. Chromosomal approaches to oncogenes and oncogenesis. FASEB J. 1988;2(15):3054–3060. doi: 10.1096/fasebj.2.15.3056765. [DOI] [PubMed] [Google Scholar]
- 17.Gauwerky CE, Croce CM. Chromosomal translocations in leukaemia. Semin Cancer Biol. 1993;4(6):333–340. [PubMed] [Google Scholar]
- 18.Hunger SP, Cleary ML. Chimaeric oncoproteins resulting from chromosomal translocations in acute lymphoblastic leukaemia. Semin Cancer Biol. 1993;4(6):387–399. [PubMed] [Google Scholar]
- 19.Ladanyi M. The emerging molecular genetics of sarcoma translocations. Diagn Mol Pathol. 1995;4(3):162–173. doi: 10.1097/00019606-199509000-00003. [DOI] [PubMed] [Google Scholar]
- 20.Brenner JC, Chinnaiyan AM. Translocations in epithelial cancers. Biochim Biophys Acta. 2009;1796(2):201–215. doi: 10.1016/j.bbcan.2009.04.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Edwards PA. Fusion genes and chromosome translocations in the common epithelial cancers. J Pathol. 2010;220(2):244–254. doi: 10.1002/path.2632. [DOI] [PubMed] [Google Scholar]
- 22.Zheng J. Oncogenic chromosomal translocations and human cancer (review) Oncol Rep. 2013;30(5):2011–2019. doi: 10.3892/or.2013.2677. [DOI] [PubMed] [Google Scholar]
- 23.Apperley JF. Chronic myeloid leukaemia. Lancet. 2015;385(9976):1447–1459. doi: 10.1016/S0140-6736(13)62120-0. [DOI] [PubMed] [Google Scholar]
- 24.Minciacchi VR, Kumar R, Krause DS. Chronic myeloid leukemia: a model disease of the past, present and future. Cells. 2021;10(1):117. doi: 10.3390/cells10010117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Panagopoulos I, Heim S. Interstitial deletions generating fusion genes. Cancer Genomics Proteomics. 2021;18(3):167–196. doi: 10.21873/cgp.20251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Kundu M, Liu PP. Function of the inv(16) fusion gene CBFB-MYH11. Curr Opin Hematol. 2001;8(4):201–205. doi: 10.1097/00062752-200107000-00004. [DOI] [PubMed] [Google Scholar]
- 27.Reilly JT. Pathogenesis of acute myeloid leukaemia and inv(16)(p13;q22): a paradigm for understanding leukaemogenesis. Br J Haematol. 2005;128(1):18–34. doi: 10.1111/j.1365-2141.2004.05236.x. [DOI] [PubMed] [Google Scholar]
- 28.von Bergh A, Gargallo P, De Prijck B, Vranckx H, Marschalek R, Larripa I, Kluin P, Schuuring E, Hagemeijer A. Cryptic t(4;11) encoding MLL-AF4 due to insertion of 5’ MLL sequences in chromosome 4. Leukemia. 2001;15(4):595–600. doi: 10.1038/sj.leu.2402050. [DOI] [PubMed] [Google Scholar]
- 29.Morel F, Le Bris MJ, Douet-Guilbert N, Duchemin J, Herry A, Le Calvez G, Marion V, Berthou C, De Braekeleer M. Insertion of chromosome 11 in chromosome 4 resulting in a 5’MLL-3’AF4 fusion gene in a case of adult acute lymphoblastic leukemia. Cancer Genet Cytogenet. 2003;145(1):74–77. doi: 10.1016/s0165-4608(03)00029-3. [DOI] [PubMed] [Google Scholar]
- 30.Soler G, Radford I, Meyer C, Marschalek R, Brouzes C, Ghez D, Romana S, Berger R. MLL insertion with MLL-MLLT3 gene fusion in acute leukemia: case report and review of the literature. Cancer Genet Cytogenet. 2008;183(1):53–59. doi: 10.1016/j.cancergencyto.2008.01.016. [DOI] [PubMed] [Google Scholar]
- 31.Matveeva E, Kazakova A, Olshanskaya Y, Tsaur G, Shelikhova L, Meyer C, Marschalek R, Novichkova G, Maschan M, Maschan A. A new variant of KMT2A(MLL)-FLNA fusion transcript in acute myeloid leukemia with ins(X;11)(q28;q23q23) Cancer Genet. 2015;208(4):148–151. doi: 10.1016/j.cancergen.2015.03.001. [DOI] [PubMed] [Google Scholar]
- 32.Jones DT, Kocialkowski S, Liu L, Pearson DM, Bäcklund LM, Ichimura K, Collins VP. Tandem duplication producing a novel oncogenic BRAF fusion gene defines the majority of pilocytic astrocytomas. Cancer Res. 2008;68(21):8673–8677. doi: 10.1158/0008-5472.CAN-08-2097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Panagopoulos I, Brandal P, Gorunova L, Bjerkehagen B, Heim S. Novel CSF1-S100A10 fusion gene and CSF1 transcript identified by RNA sequencing in tenosynovial giant cell tumors. Int J Oncol. 2014;44(5):1425–1432. doi: 10.3892/ijo.2014.2326. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Ho J, Peters T, Dickson BC, Swanson D, Fernandez A, Frova-Seguin A, Valentin MA, Schramm U, Sultan M, Nielsen TO, Demicco EG. Detection of CSF1 rearrangements deleting the 3’ UTR in tenosynovial giant cell tumors. Genes Chromosomes Cancer. 2020;59(2):96–105. doi: 10.1002/gcc.22807. [DOI] [PubMed] [Google Scholar]
- 35.Tsuda Y, Hirata M, Katayama K, Motoi T, Matsubara D, Oda Y, Fujita M, Kobayashi H, Kawano H, Nishida Y, Sakai T, Okuma T, Goto T, Ogura K, Kawai A, Ae K, Anazawa U, Suehara Y, Iwata S, Miyano S, Imoto S, Shibata T, Nakagawa H, Yamaguchi R, Tanaka S, Matsuda K. Massively parallel sequencing of tenosynovial giant cell tumors reveals novel CSF1 fusion transcripts and novel somatic CBL mutations. Int J Cancer. 2019;145(12):3276–3284. doi: 10.1002/ijc.32421. [DOI] [PubMed] [Google Scholar]
- 36.Yagasaki F, Jinnai I, Yoshida S, Yokoyama Y, Matsuda A, Kusumoto S, Kobayashi H, Terasaki H, Ohyashiki K, Asou N, Murohashi I, Bessho M, Hirashima K. Fusion of TEL/ETV6 to a novel ACS2 in myelodysplastic syndrome and acute myelogenous leukemia with t(5;12)(q31;p13) Genes Chromosomes Cancer. 1999;26(3):192–202. doi: 10.1002/(sici)1098-2264(199911)26:3<192::aid-gcc2>3.0.co;2-e. [DOI] [PubMed] [Google Scholar]
- 37.Odero MD, Vizmanos JL, Román JP, Lahortiga I, Panizo C, Calasanz MJ, Zeleznik-Le NJ, Rowley JD, Novo FJ. A novel gene, MDS2, is fused to ETV6/TEL in a t(1;12)(p36.1;p13) in a patient with myelodysplastic syndrome. Genes Chromosomes Cancer. 2002;35(1):11–19. doi: 10.1002/gcc.10090. [DOI] [PubMed] [Google Scholar]
- 38.Murga Penas EM, Cools J, Algenstaedt P, Hinz K, Seeger D, Schafhausen P, Schilling G, Marynen P, Hossfeld DK, Dierlamm J. A novel cryptic translocation t(12;17)(p13;p12-p13) in a secondary acute myeloid leukemia results in a fusion of the ETV6 gene and the antisense strand of the PER1 gene. Genes Chromosomes Cancer. 2003;37(1):79–83. doi: 10.1002/gcc.10175. [DOI] [PubMed] [Google Scholar]
- 39.Belloni E, Trubia M, Mancini M, Derme V, Nanni M, Lahortiga I, Riccioni R, Confalonieri S, Lo-Coco F, Di Fiore PP, Pelicci PG. A new complex rearrangement involving the ETV6, LOC115548, and MN1 genes in a case of acute myeloid leukemia. Genes Chromosomes Cancer. 2004;41(3):272–277. doi: 10.1002/gcc.20081. [DOI] [PubMed] [Google Scholar]
- 40.Panagopoulos I, Strömbeck B, Isaksson M, Heldrup J, Olofsson T, Johansson B. Fusion of ETV6 with an intronic sequence of the BAZ2A gene in a paediatric pre-B acute lymphoblastic leukaemia with a cryptic chromosome 12 rearrangement. Br J Haematol. 2006;133(3):270–275. doi: 10.1111/j.1365-2141.2006.06020.x. [DOI] [PubMed] [Google Scholar]
- 41.Zhang L, Wang M, Wang Z, Zeng Z, Wen L, Xu Y, Yao L, Cen J, Li H, Pan J, Sun A, Wu D, Chen S, Ma L, Yang X. Identification of a novel ETV6 truncated fusion gene in myeloproliferative neoplasm, unclassifiable with t(4;12)(q12;p13) Ann Hematol. 2020;99(10):2445–2447. doi: 10.1007/s00277-020-04207-y. [DOI] [PubMed] [Google Scholar]
- 42.van IJzendoorn DG, de Jong D, Romagosa C, Picci P, Benassi MS, Gambarotti M, Daugaard S, van de Sande M, Szuhai K, Bovée JV. Fusion events lead to truncation of FOS in epithelioid hemangioma of bone. Genes Chromosomes Cancer. 2015;54(9):565–574. doi: 10.1002/gcc.22269. [DOI] [PubMed] [Google Scholar]
- 43.Huang SC, Zhang L, Sung YS, Chen CL, Krausz T, Dickson BC, Kao YC, Agaram NP, Fletcher CD, Antonescu CR. Frequent FOS gene rearrangements in epithelioid hemangioma: a molecular study of 58 cases with morphologic reappraisal. Am J Surg Pathol. 2015;39(10):1313–1321. doi: 10.1097/PAS.0000000000000469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Fittall MW, Mifsud W, Pillay N, Ye H, Strobl AC, Verfaillie A, Demeulemeester J, Zhang L, Berisha F, Tarabichi M, Young MD, Miranda E, Tarpey PS, Tirabosco R, Amary F, Grigoriadis AE, Stratton MR, Van Loo P, Antonescu CR, Campbell PJ, Flanagan AM, Behjati S. Recurrent rearrangements of FOS and FOSB define osteoblastoma. Nat Commun. 2018;9(1):2150. doi: 10.1038/s41467-018-04530-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Panagopoulos I, Gorunova L, Lobmaier I, Andersen K, Kostolomov I, Lund-Iversen M, Bjerkehagen B, Heim S. FOS-ANKH and FOS-RUNX2 fusion genes in osteoblastoma. Cancer Genomics Proteomics. 2020;17(2):161–168. doi: 10.21873/cgp.20176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Agostini A, Gorunova L, Bjerkehagen B, Lobmaier I, Heim S, Panagopoulos I. Molecular characterization of the t(4;12)(q27~28;q14~15) chromosomal rearrangement in lipoma. Oncol Lett. 2016;12(3):1701–1704. doi: 10.3892/ol.2016.4834. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Broberg K, Zhang M, Strömbeck B, Isaksson M, Nilsson M, Mertens F, Mandahl N, Panagopoulos I. Fusion of RDC1 with HMGA2 in lipomas as the result of chromosome aberrations involving 2q35-37 and 12q13-15. Int J Oncol. 2002;21(2):321–326. [PubMed] [Google Scholar]
- 48.Lacaria M, El Demellawy D, McGowan-Jordan J. A rare case of pediatric lipoma with t(9;12)(p22;q14) and evidence of HMGA2-NFIB gene fusion. Cancer Genet. 2017;216-217:100–104. doi: 10.1016/j.cancergen.2017.07.011. [DOI] [PubMed] [Google Scholar]
- 49.Nilsson M, Mertens F, Höglund M, Mandahl N, Panagopoulos I. Truncation and fusion of HMGA2 in lipomas with rearrangements of 5q32—>q33 and 12q14—>q15. Cytogenet Genome Res. 2006;112(1-2):60–66. doi: 10.1159/000087514. [DOI] [PubMed] [Google Scholar]
- 50.Nilsson M, Panagopoulos I, Mertens F, Mandahl N. Fusion of the HMGA2 and NFIB genes in lipoma. Virchows Arch. 2005;447(5):855–858. doi: 10.1007/s00428-005-0037-9. [DOI] [PubMed] [Google Scholar]
- 51.Panagopoulos I, Gorunova L, Agostini A, Lobmaier I, Bjerkehagen B, Heim S. Fusion of the HMGA2 and C9orf92 genes in myolipoma with t(9;12)(p22;q14) Diagn Pathol. 2016;11:22. doi: 10.1186/s13000-016-0472-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Pierron A, Fernandez C, Saada E, Keslair F, Hery G, Zattara H, Pedeutour F. HMGA2-NFIB fusion in a pediatric intramuscular lipoma: a novel case of NFIB alteration in a large deep-seated adipocytic tumor. Cancer Genet Cytogenet. 2009;195(1):66–70. doi: 10.1016/j.cancergencyto.2009.06.009. [DOI] [PubMed] [Google Scholar]
- 53.Ramsey H, Zhang DE, Richkind K, Burcoglu-O’Ral A, Hromas R. Fusion of AML1/Runx1 to copine VIII, a novel member of the copine family, in an aggressive acute myelogenous leukemia with t(12;21) translocation. Leukemia. 2003;17(8):1665–1666. doi: 10.1038/sj.leu.2403048. [DOI] [PubMed] [Google Scholar]
- 54.Nguyen TT, Ma LN, Slovak ML, Bangs CD, Cherry AM, Arber DA. Identification of novel Runx1 (AML1) translocation partner genes SH3D19, YTHDf2, and ZNF687 in acute myeloid leukemia. Genes Chromosomes Cancer. 2006;45(10):918–932. doi: 10.1002/gcc.20355. [DOI] [PubMed] [Google Scholar]
- 55.Agerstam H, Lilljebjörn H, Lassen C, Swedin A, Richter J, Vandenberghe P, Johansson B, Fioretos T. Fusion gene-mediated truncation of RUNX1 as a potential mechanism underlying disease progression in the 8p11 myeloproliferative syndrome. Genes Chromosomes Cancer. 2007;46(7):635–643. doi: 10.1002/gcc.20442. [DOI] [PubMed] [Google Scholar]
- 56.Giguère A, Hébert J. CLCA2, a novel RUNX1 partner gene in a therapy-related leukemia with t(1;21)(p22;q22) Cancer Genet Cytogenet. 2010;202(2):94–100. doi: 10.1016/j.cancergencyto.2010.07.116. [DOI] [PubMed] [Google Scholar]
- 57.Giguère A, Hébert J. Identification of a novel fusion gene involving RUNX1 and the antisense strand of SV2B in a BCR-ABL1-positive acute leukemia. Genes Chromosomes Cancer. 2013;52(12):1114–1122. doi: 10.1002/gcc.22105. [DOI] [PubMed] [Google Scholar]
- 58.Panagopoulos I, Torkildsen S, Gorunova L, Ulvmoen A, Tierens A, Zeller B, Heim S. RUNX1 truncation resulting from a cryptic and novel t(6;21)(q25;q22) chromosome translocation in acute myeloid leukemia: A case report. Oncol Rep. 2016;36(5):2481–2488. doi: 10.3892/or.2016.5119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Delespaul L, Lesluyes T, Pérot G, Brulard C, Lartigue L, Baud J, Lagarde P, Le Guellec S, Neuville A, Terrier P, Vince-Ranchère D, Schmidt S, Debant A, Coindre JM, Chibon F. Recurrent TRIO fusion in nontranslocation-related sarcomas. Clin Cancer Res. 2017;23(3):857–867. doi: 10.1158/1078-0432.CCR-16-0290. [DOI] [PubMed] [Google Scholar]
- 60.Rasighaemi P, Liongue C, Onnebo SM, Ward AC. Functional analysis of truncated forms of ETV6. Br J Haematol. 2015;171(4):658–662. doi: 10.1111/bjh.13428. [DOI] [PubMed] [Google Scholar]
- 61.van IJzendoorn DGP, Forghany Z, Liebelt F, Vertegaal AC, Jochemsen AG, Bovée JVMG, Szuhai K, Baker DA. Functional analyses of a human vascular tumor FOS variant identify a novel degradation mechanism and a link to tumorigenesis. J Biol Chem. 2017;292(52):21282–21290. doi: 10.1074/jbc.C117.815845. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Fedele M, Berlingieri MT, Scala S, Chiariotti L, Viglietto G, Rippel V, Bullerdiek J, Santoro M, Fusco A. Truncated and chimeric HMGI-C genes induce neoplastic transformation of NIH3T3 murine fibroblasts. Oncogene. 1998;17(4):413–418. doi: 10.1038/sj.onc.1201952. [DOI] [PubMed] [Google Scholar]
- 63.Mas A, Cervelló I, Fernández-Álvarez A, Faus A, Díaz A, Burgués O, Casado M, Simón C. Overexpression of the truncated form of High Mobility Group A proteins (HMGA2) in human myometrial cells induces leiomyoma-like tissue formation. Mol Hum Reprod. 2015;21(4):330–338. doi: 10.1093/molehr/gau114. [DOI] [PubMed] [Google Scholar]
- 64.Richter A, Lübbing M, Frank HG, Nolte I, Bullerdiek JC, von Ahsen I. High-mobility group protein HMGA2-derived fragments stimulate the proliferation of chondrocytes and adipose tissue-derived stem cells. Eur Cell Mater. 2011;21:355–363. doi: 10.22203/ecm.v021a26. [DOI] [PubMed] [Google Scholar]
- 65.Arlotta P, Tai AK, Manfioletti G, Clifford C, Jay G, Ono SJ. Transgenic mice expressing a truncated form of the high mobility group I-C protein develop adiposity and an abnormally high prevalence of lipomas. J Biol Chem. 2000;275(19):14394–14400. doi: 10.1074/jbc.m000564200. [DOI] [PubMed] [Google Scholar]
- 66.Ikeda K, Mason PJ, Bessler M. 3’UTR-truncated Hmga2 cDNA causes MPN-like hematopoiesis by conferring a clonal growth advantage at the level of HSC in mice. Blood. 2011;117(22):5860–5869. doi: 10.1182/blood-2011-02-334425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Micci F, Panagopoulos I, Haugom L, Andersen HK, Tjønnfjord GE, Beiske K, Heim S. t(3;21)(q22;q22) leading to truncation of the RYK gene in atypical chronic myeloid leukemia. Cancer Lett. 2009;277(2):205–211. doi: 10.1016/j.canlet.2008.12.016. [DOI] [PubMed] [Google Scholar]
- 68.Hromas R, Shopnick R, Jumean HG, Bowers C, Varella-Garcia M, Richkind K. A novel syndrome of radiation-associated acute myeloid leukemia involving AML1 gene translocations. Blood. 2000;95(12):4011–4013. [PubMed] [Google Scholar]
- 69.Hromas R, Busse T, Carroll A, Mack D, Shopnick R, Zhang DE, Nakshatri H, Richkind K. Fusion AML1 transcript in a radiation-associated leukemia results in a truncated inhibitory AML1 protein. Blood. 2001;97(7):2168–2170. doi: 10.1182/blood.v97.7.2168. [DOI] [PubMed] [Google Scholar]
- 70.Mardis ER. Next-generation DNA sequencing methods. Annu Rev Genomics Hum Genet. 2008;9:387–402. doi: 10.1146/annurev.genom.9.081307.164359. [DOI] [PubMed] [Google Scholar]
- 71.Meyerson M, Gabriel S, Getz G. Advances in understanding cancer genomes through second-generation sequencing. Nat Rev Genet. 2010;11(10):685–696. doi: 10.1038/nrg2841. [DOI] [PubMed] [Google Scholar]
- 72.Kumar S, Razzaq SK, Vo AD, Gautam M, Li H. Identifying fusion transcripts using next generation sequencing. Wiley Interdiscip Rev RNA. 2016;7(6):811–823. doi: 10.1002/wrna.1382. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Wang L. Identification of cancer gene fusions based on advanced analysis of the human genome or transcriptome. Front Med. 2013;7(3):280–289. doi: 10.1007/s11684-013-0265-3. [DOI] [PubMed] [Google Scholar]
- 74.Panagopoulos I, Thorsen J, Gorunova L, Micci F, Heim S. Sequential combination of karyotyping and RNA-sequencing in the search for cancer-specific fusion genes. Int J Biochem Cell Biol. 2014;53:462–465. doi: 10.1016/j.biocel.2014.05.018. [DOI] [PubMed] [Google Scholar]
- 75.Wang Q, Xia J, Jia P, Pao W, Zhao Z. Application of next generation sequencing to human gene fusion detection: computational tools, features and perspectives. Brief Bioinform. 2013;14(4):506–519. doi: 10.1093/bib/bbs044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Jilani M, Haspel N. Computational methods for detecting large-scale structural rearrangements in chromosomes. In: Bioinformatics. In: Helder IN, editor. Brisbane, Australia. 2021. [PubMed] [Google Scholar]
- 77.Panagopoulos I, Thorsen J, Gorunova L, Haugom L, Bjerkehagen B, Davidson B, Heim S, Micci F. Fusion of the ZC3H7B and BCOR genes in endometrial stromal sarcomas carrying an X;22-translocation. Genes Chromosomes Cancer. 2013;52(7):610–618. doi: 10.1002/gcc.22057. [DOI] [PubMed] [Google Scholar]
- 78.Panagopoulos I, Gorunova L, Bjerkehagen B, Heim S. The “grep” command but not FusionMap, FusionFinder or ChimeraScan captures the CIC-DUX4 fusion gene from whole transcriptome sequencing data on a small round cell tumor with t(4;19)(q35;q13) PLoS One. 2014;9(6):e99439. doi: 10.1371/journal.pone.0099439. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Panagopoulos I, Torkildsen S, Gorunova L, Tierens A, Tjønnfjord GE, Heim S. Comparison between karyotyping-FISH-reverse transcription PCR and RNA-sequencing-fusion gene identification programs in the detection of KAT6A-CREBBP in acute myeloid leukemia. PLoS One. 2014;9(5):e96570. doi: 10.1371/journal.pone.0096570. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Nicorici D, Satalan M, Edgren H, Kangaspeska S, Murumagi A, Kallioniemi O, Virtanen S, Kilkku O. FusionCatcher - a tool for finding somatic fusion genes in paired-end RNAsequencing data. biorxiv. 2020 doi: 10.1101/011650. [DOI] [Google Scholar]
- 81.Teixeira MR, Heim S. Cytogenetic analysis of tumor clonality. Adv Cancer Res. 2011;112:127–149. doi: 10.1016/B978-0-12-387688-1.00005-3. [DOI] [PubMed] [Google Scholar]
- 82.Gerlinger M, Rowan AJ, Horswell S, Math M, Larkin J, Endesfelder D, Gronroos E, Martinez P, Matthews N, Stewart A, Tarpey P, Varela I, Phillimore B, Begum S, McDonald NQ, Butler A, Jones D, Raine K, Latimer C, Santos CR, Nohadani M, Eklund AC, Spencer-Dene B, Clark G, Pickering L, Stamp G, Gore M, Szallasi Z, Downward J, Futreal PA, Swanton C. Intratumor heterogeneity and branched evolution revealed by multiregion sequencing. N Engl J Med. 2012;366(10):883–892. doi: 10.1056/NEJMoa1113205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Honeyman JN, Simon EP, Robine N, Chiaroni-Clarke R, Darcy DG, Lim II, Gleason CE, Murphy JM, Rosenberg BR, Teegan L, Takacs CN, Botero S, Belote R, Germer S, Emde AK, Vacic V, Bhanot U, LaQuaglia MP, Simon SM. Detection of a recurrent DNAJB1-PRKACA chimeric transcript in fibrolamellar hepatocellular carcinoma. Science. 2014;343(6174):1010–1014. doi: 10.1126/science.1249484. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Tanas MR, Sboner A, Oliveira AM, Erickson-Johnson MR, Hespelt J, Hanwright PJ, Flanagan J, Luo Y, Fenwick K, Natrajan R, Mitsopoulos C, Zvelebil M, Hoch BL, Weiss SW, Debiec-Rychter M, Sciot R, West RB, Lazar AJ, Ashworth A, Reis-Filho JS, Lord CJ, Gerstein MB, Rubin MA, Rubin BP. Identification of a disease-defining gene fusion in epithelioid hemangioendothelioma. Sci Transl Med. 2011;3(98):98ra82. doi: 10.1126/scitranslmed.3002409. [DOI] [PubMed] [Google Scholar]
- 85.Lee CH, Ou WB, Mariño-Enriquez A, Zhu M, Mayeda M, Wang Y, Guo X, Brunner AL, Amant F, French CA, West RB, McAlpine JN, Gilks CB, Yaffe MB, Prentice LM, McPherson A, Jones SJ, Marra MA, Shah SP, van de Rijn M, Huntsman DG, Dal Cin P, Debiec-Rychter M, Nucci MR, Fletcher JA. 14-3-3 fusion oncogenes in high-grade endometrial stromal sarcoma. Proc Natl Acad Sci U.S.A. 2012;109(3):929–934. doi: 10.1073/pnas.1115528109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Nyquist KB, Panagopoulos I, Thorsen J, Haugom L, Gorunova L, Bjerkehagen B, Fosså A, Guriby M, Nome T, Lothe RA, Skotheim RI, Heim S, Micci F. Whole-transcriptome sequencing identifies novel IRF2BP2-CDX1 fusion gene brought about by translocation t(1;5)(q42;q32) in mesenchymal chondrosarcoma. PLoS One. 2012;7(11):e49705. doi: 10.1371/journal.pone.0049705. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Micci F, Thorsen J, Panagopoulos I, Nyquist KB, Zeller B, Tierens A, Heim S. High-throughput sequencing identifies an NFIA/CBFA2T3 fusion gene in acute erythroid leukemia with t(1;16)(p31;q24) Leukemia. 2013;27(4):980–982. doi: 10.1038/leu.2012.266. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Panagopoulos I, Thorsen J, Gorunova L, Micci F, Haugom L, Davidson B, Heim S. RNA sequencing identifies fusion of the EWSR1 and YY1 genes in mesothelioma with t(14;22)(q32;q12) Genes Chromosomes Cancer. 2013;52(8):733–740. doi: 10.1002/gcc.22068. [DOI] [PubMed] [Google Scholar]
- 89.Panagopoulos I, Gorunova L, Torkildsen S, Tierens A, Heim S, Micci F. FAM53B truncation caused by t(10;19)(q26;q13) chromosome translocation in acute lymphoblastic leukemia. Oncol Lett. 2017;13(4):2216–2220. doi: 10.3892/ol.2017.5705. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Panagopoulos I, Gorunova L, Lobmaier I, Bjerkehagen B, Heim S. Identification of SETD2-NF1 fusion gene in a pediatric spindle cell tumor with the chromosomal translocation t(3;17)(p21;q12) Oncol Rep. 2017;37(6):3181–3188. doi: 10.3892/or.2017.5628. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Torkildsen S, Brunetti M, Gorunova L, Spetalen S, Beiske K, Heim S, Panagopoulos I. Rearrangement of the chromatin organizer special AT-rich binding protein 1 gene, SATB1, resulting from a t(3;5)(p24;q14) chromosomal translocation in acute myeloid leukemia. Anticancer Res. 2017;37(2):693–698. doi: 10.21873/anticanres.11365. [DOI] [PubMed] [Google Scholar]
- 92.Panagopoulos I, Gorunova L, Andersen HK, Bergrem A, Dahm A, Andersen K, Micci F, Heim S. PAN3-PSMA2 fusion resulting from a novel t(7;13)(p14;q12) chromosome translocation in a myelodysplastic syndrome that evolved into acute myeloid leukemia. Exp Hematol Oncol. 2018;7:7. doi: 10.1186/s40164-018-0099-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Brunetti M, Zeller B, Tierens A, Heim S, Micci F, Panagopoulos I. TYRO3 truncation resulting from a t(10;15)(p11;q15) chromosomal translocation in pediatric acute myeloid leukemia. Anticancer Res. 2020;40(11):6115–6121. doi: 10.21873/anticanres.14632. [DOI] [PubMed] [Google Scholar]
- 94.Panagopoulos I, Gorunova L, Lobmaier I, Andersen K, Lund-Iversen M, Micci F, Heim S. Fusion of the COL4A5 gene with NR2F2-AS1 in a hemangioma carrying a t(X;15)(q22;q26) chromosomal translocation. Cancer Genomics Proteomics. 2020;17(4):383–390. doi: 10.21873/cgp.20197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Maher CA, Kumar-Sinha C, Cao X, Kalyana-Sundaram S, Han B, Jing X, Sam L, Barrette T, Palanisamy N, Chinnaiyan AM. Transcriptome sequencing to detect gene fusions in cancer. Nature. 2009;458(7234):97–101. doi: 10.1038/nature07638. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Heyer EE, Deveson IW, Wooi D, Selinger CI, Lyons RJ, Hayes VM, O’Toole SA, Ballinger ML, Gill D, Thomas DM, Mercer TR, Blackburn J. Diagnosis of fusion genes using targeted RNA sequencing. Nat Commun. 2019;10(1):1388. doi: 10.1038/s41467-019-09374-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Avenarius MR, Miller CR, Arnold MA, Koo S, Roberts R, Hobby M, Grossman T, Moyer Y, Wilson RK, Mardis ER, Gastier-Foster JM, Pfau RB. Genetic characterization of pediatric sarcomas by targeted RNA sequencing. J Mol Diagn. 2020;22(10):1238–1245. doi: 10.1016/j.jmoldx.2020.07.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Aplan PD, Lombardi DP, Ginsberg AM, Cossman J, Bertness VL, Kirsch IR. Disruption of the human SCL locus by “illegitimate” V-(D)-J recombinase activity. Science. 1990;250(4986):1426–1429. doi: 10.1126/science.2255914. [DOI] [PubMed] [Google Scholar]
- 99.Brown L, Cheng JT, Chen Q, Siciliano MJ, Crist W, Buchanan G, Baer R. Site-specific recombination of the tal-1 gene is a common occurrence in human T cell leukemia. EMBO J. 1990;9(10):3343–3351. doi: 10.1002/j.1460-2075.1990.tb07535.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Rickman DS, Pflueger D, Moss B, VanDoren VE, Chen CX, de la Taille A, Kuefer R, Tewari AK, Setlur SR, Demichelis F, Rubin MA. SLC45A3-ELK4 is a novel and frequent erythroblast transformation-specific fusion transcript in prostate cancer. Cancer Res. 2009;69(7):2734–2738. doi: 10.1158/0008-5472.CAN-08-4926. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Kumar-Sinha C, Kalyana-Sundaram S, Chinnaiyan AM. SLC45A3-ELK4 chimera in prostate cancer: spotlight on cis-splicing. Cancer Discov. 2012;2(7):582–585. doi: 10.1158/2159-8290.CD-12-0212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Zhang Y, Gong M, Yuan H, Park HG, Frierson HF, Li H. Chimeric transcript generated by cis-splicing of adjacent genes regulates prostate cancer cell proliferation. Cancer Discov. 2012;2(7):598–607. doi: 10.1158/2159-8290.CD-12-0042. [DOI] [PubMed] [Google Scholar]
- 103.Ferrari A, Ghelli Luserna Di Rora A, Domizio C, Papayannidis C, Simonetti G, Maria Hernández-Rivas J, Rondoni M, Giglio F, Abruzzese E, Imovilli A, Iacobucci I, Calistri D, Martinelli G. Rearrangements of ATP5L-KMT2A in acute lymphoblastic leukaemia. Br J Haematol. 2021;192(6):e139–e144. doi: 10.1111/bjh.17265. [DOI] [PubMed] [Google Scholar]
- 104.Gestrich CK, Sadri N, Sinno MG, Pateva I, Meyerson HJ. Reciprocal ATP5L-KMT2A gene fusion in a paediatric B lymphoblastic leukaemia/lymphoma (B-ALL) patient. Br J Haematol. 2020;191(2):e61–e64. doi: 10.1111/bjh.17000. [DOI] [PubMed] [Google Scholar]
- 105.Yoshihara K, Wang Q, Torres-Garcia W, Zheng S, Vegesna R, Kim H, Verhaak RG. The landscape and therapeutic relevance of cancer-associated transcript fusions. Oncogene. 2015;34(37):4845–4854. doi: 10.1038/onc.2014.406. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Shlien A, Raine K, Fuligni F, Arnold R, Nik-Zainal S, Dronov S, Mamanova L, Rosic A, Ju YS, Cooke SL, Ramakrishna M, Papaemmanuil E, Davies HR, Tarpey PS, Van Loo P, Wedge DC, Jones DR, Martin S, Marshall J, Anderson E, Hardy C, ICGC Breast Cancer Working Group , Oslo Breast Cancer Research Consortium , Barbashina V, Aparicio SA, Sauer T, Garred Ø, Vincent-Salomon A, Mariani O, Boyault S, Fatima A, Langerød A, Borg Å, Thomas G, Richardson AL, Børresen-Dale AL, Polyak K, Stratton MR, Campbell PJ. Direct transcriptional consequences of somatic mutation in breast cancer. Cell Rep. 2016;16(7):2032–2046. doi: 10.1016/j.celrep.2016.07.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Vellichirammal NN, Albahrani A, Banwait JK, Mishra NK, Li Y, Roychoudhury S, Kling MJ, Mirza S, Bhakat KK, Band V, Joshi SS, Guda C. Pan-cancer analysis reveals the diverse landscape of novel sense and antisense fusion transcripts. Mol Ther Nucleic Acids. 2020;19:1379–1398. doi: 10.1016/j.omtn.2020.01.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Johansson B, Mertens F, Schyman T, Björk J, Mandahl N, Mitelman F. Most gene fusions in cancer are stochastic events. Genes Chromosomes Cancer. 2019;58(9):607–611. doi: 10.1002/gcc.22745. [DOI] [PubMed] [Google Scholar]
- 109.Panagopoulos I, Andersen K, Ramslien LF, Ikonomou IM, Micci F, Heim S. Therapy-related myeloid leukemia with the translocation t(8;19)(p11;q13) leading to a KAT6A-LEUTX fusion gene. Anticancer Res. 2021;41(4):1753–1760. doi: 10.21873/anticanres.14940. [DOI] [PubMed] [Google Scholar]
- 110.Brizard A, Guilhot F, Huret JL, Benz-Lemoine E, Tanzer J. The 8p11 anomaly in “monoblastic” leukaemia. Leuk Res. 1988;12(8):693–697. doi: 10.1016/0145-2126(88)90105-1. [DOI] [PubMed] [Google Scholar]
- 111.Stark B, Resnitzky P, Jeison M, Luria D, Blau O, Avigad S, Shaft D, Kodman Y, Gobuzov R, Ash S. A distinct subtype of M4/M5 acute myeloblastic leukemia (AML) associated with t(8:16)(p11:p13), in a patient with the variant t(8:19)(p11:q13)—case report and review of the literature. Leuk Res. 1995;19(6):367–379. doi: 10.1016/0145-2126(94)00150-9. [DOI] [PubMed] [Google Scholar]
- 112.Gervais C, Murati A, Helias C, Struski S, Eischen A, Lippert E, Tigaud I, Penther D, Bastard C, Mugneret F, Poppe B, Speleman F, Talmant P, VanDen Akker J, Baranger L, Barin C, Luquet I, Nadal N, Nguyen-Khac F, Maarek O, Herens C, Sainty D, Flandrin G, Birnbaum D, Mozziconacci MJ, Lessard M, Groupe Francophone de Cytogénétique Hématologique Acute myeloid leukaemia with 8p11 (MYST3) rearrangement: an integrated cytologic, cytogenetic and molecular study by the groupe francophone de cytogénétique hématologique. Leukemia. 2008;22(8):1567–1575. doi: 10.1038/leu.2008.128. [DOI] [PubMed] [Google Scholar]
- 113.Chinen Y, Taki T, Tsutsumi Y, Kobayashi S, Matsumoto Y, Sakamoto N, Kuroda J, Horiike S, Nishida K, Ohno H, Uike N, Taniwaki M. The leucine twenty homeobox (LEUTX) gene, which lacks a histone acetyltransferase domain, is fused to KAT6A in therapy-related acute myeloid leukemia with t(8;19)(p11;q13) Genes Chromosomes Cancer. 2014;53(4):299–308. doi: 10.1002/gcc.22140. [DOI] [PubMed] [Google Scholar]
- 114.Eason AC, Bunting ST, Peterson JF, Saxe D, Sabnis HS. Acute myeloid leukemia in an infant with t(8;19)(p11.2;q13) translocation: case report and a review of the literature. Case Rep Hematol. 2019;2019:4198415. doi: 10.1155/2019/4198415. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Sramkova L, Cermakova J, Kutkova K, Zemanova Z, Pavlicek P, Zuna J, Stary J, Zaliova M. Rapidly progressing acute myeloid leukemia with KAT6A-LEUTX fusion in a newborn. Pediatr Blood Cancer. 2020;67(10):e28663. doi: 10.1002/pbc.28663. [DOI] [PubMed] [Google Scholar]
- 116.Panagopoulos I, Micci F, Thorsen J, Haugom L, Buechner J, Kerndrup G, Tierens A, Zeller B, Heim S. Fusion of ZMYND8 and RELA genes in acute erythroid leukemia. PLoS One. 2013;8(5):e63663. doi: 10.1371/journal.pone.0063663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Iacobucci I, Wen J, Meggendorfer M, Choi JK, Shi L, Pounds SB, Carmichael CL, Masih KE, Morris SM, Lindsley RC, Janke LJ, Alexander TB, Song G, Qu C, Li Y, Payne-Turner D, Tomizawa D, Kiyokawa N, Valentine M, Valentine V, Basso G, Locatelli F, Enemark EJ, Kham SKY, Yeoh AEJ, Ma X, Zhou X, Sioson E, Rusch M, Ries RE, Stieglitz E, Hunger SP, Wei AH, To LB, Lewis ID, D’Andrea RJ, Kile BT, Brown AL, Scott HS, Hahn CN, Marlton P, Pei D, Cheng C, Loh ML, Ebert BL, Meshinchi S, Haferlach T, Mullighan CG. Genomic subtyping and therapeutic targeting of acute erythroleukemia. Nat Genet. 2019;51(4):694–704. doi: 10.1038/s41588-019-0375-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Panagopoulos I, Gorunova L, Bjerkehagen B, Lobmaier I, Heim S. Fusion of the TBL1XR1 and HMGA1 genes in splenic hemangioma with t(3;6)(q26;p21) Int J Oncol. 2016;48(3):1242–1250. doi: 10.3892/ijo.2015.3310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Paulien S, Maarek O, Daniel MT, Berger R. A novel translocation, t(9;21)(q13;q22) rearranging the RUNX1 gene in acute myelomonocytic leukemia. Ann Genet. 2002;45(2):67–69. doi: 10.1016/s0003-3995(02)01113-9. [DOI] [PubMed] [Google Scholar]
- 120.Rodriguez-Perales S, Torres-Ruiz R, Suela J, Acquadro F, Martin MC, Yebra E, Ramirez JC, Alvarez S, Cigudosa JC. Truncated RUNX1 protein generated by a novel t(1;21)(p32;q22) chromosomal translocation impairs the proliferation and differentiation of human hematopoietic progenitors. Oncogene. 2016;35(1):125–134. doi: 10.1038/onc.2015.70. [DOI] [PubMed] [Google Scholar]
- 121.Stevens-Kroef MJ, Schoenmakers EF, van Kraaij M, Huys E, Vermeulen S, van der Reijden B, van Kessel AG. Identification of truncated RUNX1 and RUNX1-PRDM16 fusion transcripts in a case of t(1;21)(p36;q22)-positive therapy-related AML. Leukemia. 2006;20(6):1187–1189. doi: 10.1038/sj.leu.2404210. [DOI] [PubMed] [Google Scholar]
- 122.De Braekeleer M, Poon MC, Russell J, Lin CC. A case of acute lymphoblastic leukemia with t(10;19)(q26;q13) Cancer Genet Cytogenet. 1985;16(4):369–372. doi: 10.1016/0165-4608(85)90247-x. [DOI] [PubMed] [Google Scholar]
- 123.Panagopoulos I, Gorunova L, Andersen HK, Pedersen TD, Lømo J, Lund-Iversen M, Micci F, Heim S. Genetic characterization of myoid hamartoma of the breast. Cancer Genomics Proteomics. 2019;16(6):563–568. doi: 10.21873/cgp.20158. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Panagopoulos I, Gorunova L, Andersen K, Lund-Iversen M, Hognestad HR, Lobmaier I, Micci F, Heim S. Chromosomal translocation t(5;12)(p13;q14) leading to fusion of high-mobility group AT-hook 2 gene with intergenic sequences from chromosome sub-band 5p13.2 in benign myoid neoplasms of the breast: a second case. Cancer Genomics Proteomics. 2022;19(4):445–455. doi: 10.21873/cgp.20331. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Hatano H, Morita T, Ogose A, Hotta T, Kobayashi H, Segawa H, Uchiyama T, Takenouchi T, Sato T. Clinicopathological features of lipomas with gene fusions involving HMGA2. Anticancer Res. 2008;28(1B):535–538. [PubMed] [Google Scholar]
- 126.Geurts JM, Schoenmakers EF, Röijer E, Aström AK, Stenman G, van de Ven WJ. Identification of NFIB as recurrent translocation partner gene of HMGIC in pleomorphic adenomas. Oncogene. 1998;16(7):865–872. doi: 10.1038/sj.onc.1201609. [DOI] [PubMed] [Google Scholar]
- 127.Italiano A, Ebran N, Attias R, Chevallier A, Monticelli I, Mainguené C, Benchimol D, Pedeutour F. NFIB rearrangement in superficial, retroperitoneal, and colonic lipomas with aberrations involving chromosome band 9p22. Genes Chromosomes Cancer. 2008;47(11):971–977. doi: 10.1002/gcc.20602. [DOI] [PubMed] [Google Scholar]
- 128.Zhang S, Mo Q, Wang X. Oncological role of HMGA2 (Review) Int J Oncol. 2019;55(4):775–788. doi: 10.3892/ijo.2019.4856. [DOI] [PubMed] [Google Scholar]
- 129.Unachukwu U, Chada K, D’Armiento J. High mobility group AT-Hook 2 (HMGA2) oncogenicity in mesenchymal and epithelial neoplasia. Int J Mol Sci. 2020;21(9):3151. doi: 10.3390/ijms21093151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Mansoori B, Mohammadi A, Ditzel HJ, Duijf PHG, Khaze V, Gjerstorff MF, Baradaran B. HMGA2 as a critical regulator in cancer development. Genes (Basel) 2021;12(2):269. doi: 10.3390/genes12020269. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Tarpey PS, Behjati S, Cooke SL, Van Loo P, Wedge DC, Pillay N, Marshall J, O’Meara S, Davies H, Nik-Zainal S, Beare D, Butler A, Gamble J, Hardy C, Hinton J, Jia MM, Jayakumar A, Jones D, Latimer C, Maddison M, Martin S, McLaren S, Menzies A, Mudie L, Raine K, Teague JW, Tubio JM, Halai D, Tirabosco R, Amary F, Campbell PJ, Stratton MR, Flanagan AM, Futreal PA. Frequent mutation of the major cartilage collagen gene COL2A1 in chondrosarcoma. Nat Genet. 2013;45(8):923–926. doi: 10.1038/ng.2668. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Totoki Y, Yoshida A, Hosoda F, Nakamura H, Hama N, Ogura K, Yoshida A, Fujiwara T, Arai Y, Toguchida J, Tsuda H, Miyano S, Kawai A, Shibata T. Unique mutation portraits and frequent COL2A1 gene alteration in chondrosarcoma. Genome Res. 2014;24(9):1411–1420. doi: 10.1101/gr.160598.113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Tsuda Y, Suurmeijer AJH, Sung YS, Zhang L, Healey JH, Antonescu CR. Epithelioid hemangioma of bone harboring FOS and FOSB gene rearrangements: A clinicopathologic and molecular study. Genes Chromosomes Cancer. 2021;60(1):17–25. doi: 10.1002/gcc.22898. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Stransky N, Cerami E, Schalm S, Kim JL, Lengauer C. The landscape of kinase fusions in cancer. Nat Commun. 2014;5:4846. doi: 10.1038/ncomms5846. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Suster DI, Deshpande V, Chebib I, Taylor MS, Mullen J, Bredella MA, Nielsen GP. Spindle cell liposarcoma with a TRIO-TERT fusion transcript. Virchows Arch. 2019;475(3):391–394. doi: 10.1007/s00428-019-02545-5. [DOI] [PubMed] [Google Scholar]
- 136.Nucifora G, Begy CR, Erickson P, Drabkin HA, Rowley JD. The 3;21 translocation in myelodysplasia results in a fusion transcript between the AML1 gene and the gene for EAP, a highly conserved protein associated with the Epstein-Barr virus small RNA EBER 1. Proc Natl Acad Sci U.S.A. 1993;90(16):7784–7788. doi: 10.1073/pnas.90.16.7784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Nucifora G, Birn DJ, Espinosa R 3rd, Erickson P, LeBeau MM, Roulston D, McKeithan TW, Drabkin H, Rowley JD. Involvement of the AML1 gene in the t(3;21) in therapy-related leukemia and in chronic myeloid leukemia in blast crisis. Blood. 1993;81(10):2728–2734. [PubMed] [Google Scholar]
- 138.Zent CS, Mathieu C, Claxton DF, Zhang DE, Tenen DG, Rowley JD, Nucifora G. The chimeric genes AML1/MDS1 and AML1/EAP inhibit AML1B activation at the CSF1R promoter, but only AML1/MDS1 has tumor-promoter properties. Proc Natl Acad Sci U.S.A. 1996;93(3):1044–1048. doi: 10.1073/pnas.93.3.1044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Mikhail FM, Coignet L, Hatem N, Mourad ZI, Farawela HM, El Kaffash DM, Farahat N, Nucifora G. A novel gene, FGA7, is fused to RUNX1/AML1 in a t(4;21)(q28;q22) in a patient with T-cell acute lymphoblastic leukemia. Genes Chromosomes Cancer. 2004;39(2):110–118. doi: 10.1002/gcc.10302. [DOI] [PubMed] [Google Scholar]
- 140.Maki K, Sasaki K, Sugita F, Nakamura Y, Mitani K. Acute myeloid leukemia with t(7;21)(q11.2;q22) expresses a novel, reversed-sequence RUNX1-DTX2 chimera. Int J Hematol. 2012;96(2):268–273. doi: 10.1007/s12185-012-1112-z. [DOI] [PubMed] [Google Scholar]
- 141.Wang H, Chen X, Xu Z, Tan Y, Qi X, Zhang L, Xu A, Ren F. Identification of a novel fusion gene, RUNX1-PRPF38A, in acute myeloid leukemia. Int J Lab Hematol. 2017;39(4):e90–e93. doi: 10.1111/ijlh.12642. [DOI] [PubMed] [Google Scholar]
- 142.Abe A, Yamamoto Y, Katsumi A, Yamamoto H, Okamoto A, Inaguma Y, Iriyama C, Tokuda M, Okamoto M, Emi N, Tomita A. Truncated RUNX1 generated by the fusion of RUNX1 to antisense GRIK2 via a cryptic chromosome translocation enhances sensitivity to granulocyte colony-stimulating factor. Cytogenet Genome Res. 2020;160(5):255–263. doi: 10.1159/000508012. [DOI] [PubMed] [Google Scholar]
- 143.Suto Y, Sato Y, Smith SD, Rowley JD, Bohlander SK. A t(6;12)(q23;p13) results in the fusion of ETV6 to a novel gene, STL, in a B-cell ALL cell line. Genes Chromosomes Cancer. 1997;18(4):254–268. doi: 10.1002/(sici)1098-2264(199704)18:4<254::aid-gcc3>3.0.co;2-#. [DOI] [PubMed] [Google Scholar]
- 144.Cools J, Mentens N, Odero MD, Peeters P, Wlodarska I, Delforge M, Hagemeijer A, Marynen P. Evidence for position effects as a variant ETV6-mediated leukemogenic mechanism in myeloid leukemias with a t(4;12)(q11-q12;p13) or t(5;12)(q31;p13) Blood. 2002;99(5):1776–1784. doi: 10.1182/blood.v99.5.1776. [DOI] [PubMed] [Google Scholar]
- 145.Sugimoto T, Tomita A, Abe A, Iriyama C, Kiyoi H, Naoe T. Chimeric antisense RNA derived from chromosomal translocation modulates target gene expression. Haematologica. 2012;97(8):1278–1280. doi: 10.3324/haematol.2011.057869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Mori T, Nishimura N, Hasegawa D, Kawasaki K, Kosaka Y, Uchide K, Yanai T, Hayakawa A, Takeshima Y, Nishio H, Matsuo M. Persistent detection of a novel MLL-SACM1L rearrangement in the absence of leukemia. Leuk Res. 2010;34(10):1398–1401. doi: 10.1016/j.leukres.2010.05.001. [DOI] [PubMed] [Google Scholar]
- 147.Meyer C, Kowarz E, Yip SF, Wan TS, Chan TK, Dingermann T, Chan LC, Marschalek R. A complex MLL rearrangement identified five years after initial MDS diagnosis results in out-of-frame fusions without progression to acute leukemia. Cancer Genet. 2011;204(10):557–562. doi: 10.1016/j.cancergen.2011.10.001. [DOI] [PubMed] [Google Scholar]
- 148.Meyer C, Hofmann J, Burmeister T, Gröger D, Park TS, Emerenciano M, Pombo de Oliveira M, Renneville A, Villarese P, Macintyre E, Cavé H, Clappier E, Mass-Malo K, Zuna J, Trka J, De Braekeleer E, De Braekeleer M, Oh SH, Tsaur G, Fechina L, van der Velden VH, van Dongen JJ, Delabesse E, Binato R, Silva ML, Kustanovich A, Aleinikova O, Harris MH, Lund-Aho T, Juvonen V, Heidenreich O, Vormoor J, Choi WW, Jarosova M, Kolenova A, Bueno C, Menendez P, Wehner S, Eckert C, Talmant P, Tondeur S, Lippert E, Launay E, Henry C, Ballerini P, Lapillone H, Callanan MB, Cayuela JM, Herbaux C, Cazzaniga G, Kakadiya PM, Bohlander S, Ahlmann M, Choi JR, Gameiro P, Lee DS, Krauter J, Cornillet-Lefebvre P, Te Kronnie G, Schäfer BW, Kubetzko S, Alonso CN, zur Stadt U, Sutton R, Venn NC, Izraeli S, Trakhtenbrot L, Madsen HO, Archer P, Hancock J, Cerveira N, Teixeira MR, Lo Nigro L, Möricke A, Stanulla M, Schrappe M, Sedék L, Szczepański T, Zwaan CM, Coenen EA, van den Heuvel-Eibrink MM, Strehl S, Dworzak M, Panzer-Grümayer R, Dingermann T, Klingebiel T, Marschalek R. The MLL recombinome of acute leukemias in 2013. Leukemia. 2013;27(11):2165–2176. doi: 10.1038/leu.2013.135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Meyer C, Burmeister T, Gröger D, Tsaur G, Fechina L, Renneville A, Sutton R, Venn NC, Emerenciano M, Pombo-de-Oliveira MS, Barbieri Blunck C, Almeida Lopes B, Zuna J, Trka J, Ballerini P, Lapillonne H, De Braekeleer M, Cazzaniga G, Corral Abascal L, van der Velden VHJ, Delabesse E, Park TS, Oh SH, Silva MLM, Lund-Aho T, Juvonen V, Moore AS, Heidenreich O, Vormoor J, Zerkalenkova E, Olshanskaya Y, Bueno C, Menendez P, Teigler-Schlegel A, Zur Stadt U, Lentes J, Göhring G, Kustanovich A, Aleinikova O, Schäfer BW, Kubetzko S, Madsen HO, Gruhn B, Duarte X, Gameiro P, Lippert E, Bidet A, Cayuela JM, Clappier E, Alonso CN, Zwaan CM, van den Heuvel-Eibrink MM, Izraeli S, Trakhtenbrot L, Archer P, Hancock J, Möricke A, Alten J, Schrappe M, Stanulla M, Strehl S, Attarbaschi A, Dworzak M, Haas OA, Panzer-Grümayer R, Sedék L, Szczepański T, Caye A, Suarez L, Cavé H, Marschalek R. The MLL recombinome of acute leukemias in 2017. Leukemia. 2018;32(2):273–284. doi: 10.1038/leu.2017.213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Heim S, Mitelman F. Primary chromosome abnormalities in human neoplasia. Adv Cancer Res. 1989;52:1–43. doi: 10.1016/s0065-230x(08)60209-2. [DOI] [PubMed] [Google Scholar]
- 151.Mitelman F, Heim S. Chromosome abnormalities in cancer. Cancer Detect Prev. 1990;14(5):527–537. [PubMed] [Google Scholar]
- 152.Johansson B, Mertens F, Mitelman F. Primary vs. secondary neoplasia-associated chromosomal abnormalities—balanced rearrangements vs. genomic imbalances. Genes Chromosomes Cancer. 1996;16(3):155–163. doi: 10.1002/(SICI)1098-2264(199607)16:3<155::AID-GCC1>3.0.CO;2-Y. [DOI] [PubMed] [Google Scholar]
- 153.Patel UA, Bandiera A, Manfioletti G, Giancotti V, Chau KY, Crane-Robinson C. Expression and cDNA cloning of human HMGI-C phosphoprotein. Biochem Biophys Res Commun. 1994;201(1):63–70. doi: 10.1006/bbrc.1994.1669. [DOI] [PubMed] [Google Scholar]
- 154.Ashar HR, Fejzo MS, Tkachenko A, Zhou X, Fletcher JA, Weremowicz S, Morton CC, Chada K. Disruption of the architectural factor HMGI-C: DNA-binding AT hook motifs fused in lipomas to distinct transcriptional regulatory domains. Cell. 1995;82(1):57–65. doi: 10.1016/0092-8674(95)90052-7. [DOI] [PubMed] [Google Scholar]
- 155.Chau KY, Patel UA, Lee KL, Lam HY, Crane-Robinson C. The gene for the human architectural transcription factor HMGI-C consists of five exons each coding for a distinct functional element. Nucleic Acids Res. 1995;23(21):4262–4266. doi: 10.1093/nar/23.21.4262. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Schoenmakers EF, Wanschura S, Mols R, Bullerdiek J, Van den Berghe H, Van de Ven WJ. Recurrent rearrangements in the high mobility group protein gene, HMGI-C, in benign mesenchymal tumours. Nat Genet. 1995;10(4):436–444. doi: 10.1038/ng0895-436. [DOI] [PubMed] [Google Scholar]
- 157.Kurose K, Mine N, Doi D, Ota Y, Yoneyama K, Konishi H, Araki T, Emi M. Novel gene fusion of COX6C at 8q22-23 to HMGIC at 12q15 in a uterine leiomyoma. Genes Chromosomes Cancer. 2000;27(3):303–307. doi: 10.1002/(sici)1098-2264(200003)27:3<303::aid-gcc11>3.0.co;2-3. [DOI] [PubMed] [Google Scholar]
- 158.Mine N, Kurose K, Konishi H, Araki T, Nagai H, Emi M. Fusion of a sequence from HEI10 (14q11) to the HMGIC gene at 12q15 in a uterine leiomyoma. Jpn J Cancer Res. 2001;92(2):135–139. doi: 10.1111/j.1349-7006.2001.tb01075.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Pierantoni GM, Santulli B, Caliendo I, Pentimalli F, Chiappetta G, Zanesi N, Santoro M, Bulrich F, Fusco A. HMGA2 locus rearrangement in a case of acute lymphoblastic leukemia. Int J Oncol. 2003;23(2):363–367. [PubMed] [Google Scholar]
- 160.Odero MD, Grand FH, Iqbal S, Ross F, Roman JP, Vizmanos JL, Andrieux J, Laï JL, Calasanz MJ, Cross NC. Disruption and aberrant expression of HMGA2 as a consequence of diverse chromosomal translocations in myeloid malignancies. Leukemia. 2005;19(2):245–252. doi: 10.1038/sj.leu.2403605. [DOI] [PubMed] [Google Scholar]
- 161.Nyquist KB, Panagopoulos I, Thorsen J, Roberto R, Wik HS, Tierens A, Heim S, Micci F. t(12;13)(q14;q31) leading to HMGA2 upregulation in acute myeloid leukaemia. Br J Haematol. 2012;157(6):769–771. doi: 10.1111/j.1365-2141.2012.09081.x. [DOI] [PubMed] [Google Scholar]
- 162.Shimizu K, Ichikawa H, Tojo A, Kaneko Y, Maseki N, Hayashi Y, Ohira M, Asano S, Ohki M. An ets-related gene, ERG, is rearranged in human myeloid leukemia with t(16;21) chromosomal translocation. Proc Natl Acad Sci U.S.A. 1993;90(21):10280–10284. doi: 10.1073/pnas.90.21.10280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Ichikawa H, Shimizu K, Hayashi Y, Ohki M. An RNA-binding protein gene, TLS/FUS, is fused to ERG in human myeloid leukemia with t(16;21) chromosomal translocation. Cancer Res. 1994;54(11):2865–2868. [PubMed] [Google Scholar]
- 164.Panagopoulos I, Aman P, Fioretos T, Höglund M, Johansson B, Mandahl N, Heim S, Behrendtz M, Mitelman F. Fusion of the FUS gene with ERG in acute myeloid leukemia with t(16;21)(p11;q22) Genes Chromosomes Cancer. 1994;11(4):256–262. doi: 10.1002/gcc.2870110408. [DOI] [PubMed] [Google Scholar]
- 165.Kanazawa T, Ogawa C, Taketani T, Taki T, Hayashi Y, Morikawa A. TLS/FUS-ERG fusion gene in acute lymphoblastic leukemia with t(16;21)(p11;q22) and monitoring of minimal residual disease. Leuk Lymphoma. 2005;46(12):1833–1835. doi: 10.1080/10428190500162203. [DOI] [PubMed] [Google Scholar]
- 166.Panagopoulos I, Gorunova L, Zeller B, Tierens A, Heim S. Cryptic FUS-ERG fusion identified by RNA-sequencing in childhood acute myeloid leukemia. Oncol Rep. 2013;30(6):2587–2592. doi: 10.3892/or.2013.2751. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Shing DC, McMullan DJ, Roberts P, Smith K, Chin SF, Nicholson J, Tillman RM, Ramani P, Cullinane C, Coleman N. FUS/ERG gene fusions in Ewing’s tumors. Cancer Res. 2003;63(15):4568–4576. [PubMed] [Google Scholar]
- 168.Berg T, Kalsaas AH, Buechner J, Busund LT. Ewing sarcoma-peripheral neuroectodermal tumor of the kidney with a FUS-ERG fusion transcript. Cancer Genet Cytogenet. 2009;194(1):53–57. doi: 10.1016/j.cancergencyto.2009.06.002. [DOI] [PubMed] [Google Scholar]
- 169.Petit MM, Mols R, Schoenmakers EF, Mandahl N, Van de Ven WJ. LPP, the preferred fusion partner gene of HMGIC in lipomas, is a novel member of the LIM protein gene family. Genomics. 1996;36(1):118–129. doi: 10.1006/geno.1996.0432. [DOI] [PubMed] [Google Scholar]
- 170.Rogalla P, Kazmierczak B, Meyer-Bolte K, Tran KH, Bullerdiek J. The t(3;12)(q27;q14-q15) with underlying HMGIC-LPP fusion is not determining an adipocytic phenotype. Genes Chromosomes Cancer. 1998;22(2):100–104. doi: 10.1002/(sici)1098-2264(199806)22:2<100::aid-gcc3>3.0.co;2-0. [DOI] [PubMed] [Google Scholar]
- 171.Dahlén A, Mertens F, Rydholm A, Brosjö O, Wejde J, Mandahl N, Panagopoulos I. Fusion, disruption, and expression of HMGA2 in bone and soft tissue chondromas. Mod Pathol. 2003;16(11):1132–1140. doi: 10.1097/01.MP.0000092954.42656.94. [DOI] [PubMed] [Google Scholar]
- 172.Dupain C, Harttrampf AC, Boursin Y, Lebeurrier M, Rondof W, Robert-Siegwald G, Khoueiry P, Geoerger B, Massaad-Massade L. Discovery of new fusion transcripts in a cohort of pediatric solid cancers at relapse and relevance for personalized medicine. Mol Ther. 2019;27(1):200–218. doi: 10.1016/j.ymthe.2018.10.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Martin SE, Sausen M, Joseph A, Kingham BF, Martin ES. Identification of a HMGA2-EFCAB6 gene rearrangement following next-generation sequencing in a patient with a t(12;22)(q14.3;q13.2) and JAK2V617F-positive myeloproliferative neoplasm. Cancer Genet. 2012;205(6):295–303. doi: 10.1016/j.cancergen.2012.03.006. [DOI] [PubMed] [Google Scholar]
- 174.Heim S, Mitelman F. Secondary chromosome aberrations in the acute leukemias. Cancer Genet Cytogenet. 1986;22(4):331–338. doi: 10.1016/0165-4608(86)90025-7. [DOI] [PubMed] [Google Scholar]
- 175.Berger R, Le Coniat M, Derré J, Vecchione D. Secondary nonrandom chromosomal abnormalities of band 13q34 in Burkitt lymphoma-leukemia. Genes Chromosomes Cancer. 1989;1(2):115–118. doi: 10.1002/gcc.2870010202. [DOI] [PubMed] [Google Scholar]
- 176.Johansson B, Mertens F, Mitelman F. Secondary chromosomal abnormalities in acute leukemias. Leukemia. 1994;8(6):953–962. [PubMed] [Google Scholar]
- 177.Mandahl N, Mertens F, Aman P, Rydholm A, Brosjo O, Willen H, Mitelman F. Nonrandom secondary chromosome-aberrations in liposarcomas with t(12, 16) Int J Oncol. 1994;4(2):307–310. doi: 10.3892/ijo.4.2.307. [DOI] [PubMed] [Google Scholar]
- 178.Mandahl N, Limon J, Mertens F, Nedoszytko B, Gibas Z, Denis A, Willen H, Rydholm A, Szadowska A, Kreicbergs A, Rys J, Debiecrychter M, Mitelman F. Nonrandom secondary chromosome-aberrations in synovial sarcomas with t(x-18) Int J Oncol. 1995;7(3):495–499. doi: 10.3892/ijo.7.3.495. [DOI] [PubMed] [Google Scholar]
- 179.Kullendorff CM, Mertens F, Donnér M, Wiebe T, Akerman M, Mandahl N. Cytogenetic aberrations in Ewing sarcoma: are secondary changes associated with clinical outcome. Med Pediatr Oncol. 1999;32(2):79–83. doi: 10.1002/(sici)1096-911x(199902)32:2<79::aid-mpo1>3.0.co;2-r. [DOI] [PubMed] [Google Scholar]
- 180.Kovacs BZ, Niggli FK, Betts DR. Aberrations involving 13q12 approximately q14 are frequent secondary events in childhood acute lymphoblastic leukemia. Cancer Genet Cytogenet. 2004;151(2):157–161. doi: 10.1016/j.cancergencyto.2003.09.022. [DOI] [PubMed] [Google Scholar]
- 181.Attarbaschi A, Mann G, Strehl S, König M, Steiner M, Jeitler V, Lion T, Dworzak MN, Gadner H, Haas OA. Deletion of 11q23 is a highly specific nonrandom secondary genetic abnormality of ETV6/RUNX1-rearranged childhood acute lymphoblastic leukemia. Leukemia. 2007;21(3):584–586. doi: 10.1038/sj.leu.2404507. [DOI] [PubMed] [Google Scholar]
- 182.Espinet B, Salaverria I, Beà S, Ruiz-Xivillé N, Balagué O, Salido M, Costa D, Carreras J, Rodríguez-Vicente AE, Luís García J, Hernández-Rivas JM, Calasanz MJ, Siebert R, Ferrer A, Salar A, Carrió A, Polo N, García-Marco JA, Domingo A, González-Barca E, Romagosa V, Marugán I, López-Guillermo A, Millá F, Luís Mate J, Luño E, Sanzo C, Collado R, Oliver I, Monzó S, Palacín A, González T, Sant F, Salinas R, Ardanaz MT, Font L, Escoda L, Florensa L, Serrano S, Campo E, Solé F. Incidence and prognostic impact of secondary cytogenetic aberrations in a series of 145 patients with mantle cell lymphoma. Genes Chromosomes Cancer. 2010;49(5):439–451. doi: 10.1002/gcc.20754. [DOI] [PubMed] [Google Scholar]
- 183.Kjeldsen E. Oligo-based aCGH analysis reveals cryptic unbalanced der(6)t(X;6) in pediatric t(12;21)-positive acute lymphoblastic leukemia. Exp Mol Pathol. 2016;101(1):38–43. doi: 10.1016/j.yexmp.2016.05.010. [DOI] [PubMed] [Google Scholar]
- 184.Morales-Chacón K, Bourlon C, Acosta-Medina AA, Bourlon MT, Aguayo A, Tuna-Aguilar E. Impact of additional cytogenetic abnormalities on the clinical behavior of patients with chronic myeloid leukemia: report on a Latin American population. Clin Lymphoma Myeloma Leuk. 2019;19(6):e299–e306. doi: 10.1016/j.clml.2019.02.007. [DOI] [PubMed] [Google Scholar]
- 185.Han SY, Mrózek K, Voutsinas J, Wu Q, Morgan EA, Vestergaard H, Ohgami R, Kluin PM, Kristensen TK, Pullarkat S, Møller MB, Schiefer AI, Baughn LB, Kim Y, Czuchlewski D, Hilberink JR, Horny HP, George TI, Dolan M, Ku NK, Arana Yi C, Pullarkat V, Kohlschmidt J, Salhotra A, Soma L, Bloomfield CD, Chen D, Sperr WR, Marcucci G, Cho C, Akin C, Gotlib J, Broesby-Olsen S, Larson M, Linden MA, Deeg HJ, Hoermann G, Perales MA, Hornick JL, Litzow MR, Nakamura R, Weisdorf D, Borthakur G, Huls G, Valent P, Ustun C, Yeung CCS. Secondary cytogenetic abnormalities in core-binding factor AML harboring inv(16) vs. t(8;21) Blood Adv. 2021;5(10):2481–2489. doi: 10.1182/bloodadvances.2020003605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Fioretos T, Panagopoulos I, Lassen C, Swedin A, Billström R, Isaksson M, Strömbeck B, Olofsson T, Mitelman F, Johansson B. Fusion of the BCR and the fibroblast growth factor receptor-1 (FGFR1) genes as a result of t(8;22)(p11;q11) in a myeloproliferative disorder: the first fusion gene involving BCR but not ABL. Genes Chromosomes Cancer. 2001;32(4):302–310. doi: 10.1002/gcc.1195. [DOI] [PubMed] [Google Scholar]
- 187.Daraki A, Bourantas LK, Manola KN. Translocation t(5;18)(q35;q21) as a rare nonrandom abnormality in acute myeloid leukemia. Cytogenet Genome Res. 2013;139(4):289–294. doi: 10.1159/000348786. [DOI] [PubMed] [Google Scholar]
- 188.Heng HH, Bremer SW, Stevens J, Ye KJ, Miller F, Liu G, Ye CJ. Cancer progression by non-clonal chromosome aberrations. J Cell Biochem. 2006;98(6):1424–1435. doi: 10.1002/jcb.20964. [DOI] [PubMed] [Google Scholar]
- 189.Heng HH, Liu G, Bremer S, Ye KJ, Stevens J, Ye CJ. Clonal and non-clonal chromosome aberrations and genome variation and aberration. Genome. 2006;49(3):195–204. doi: 10.1139/g06-023. [DOI] [PubMed] [Google Scholar]
- 190.Heng HH, Stevens JB, Liu G, Bremer SW, Ye KJ, Reddy PV, Wu GS, Wang YA, Tainsky MA, Ye CJ. Stochastic cancer progression driven by non-clonal chromosome aberrations. J Cell Physiol. 2006;208(2):461–472. doi: 10.1002/jcp.20685. [DOI] [PubMed] [Google Scholar]
- 191.Heng HH, Regan SM, Liu G, Ye CJ. Why it is crucial to analyze non clonal chromosome aberrations or NCCAs. Mol Cytogenet. 2016;9:15. doi: 10.1186/s13039-016-0223-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Imataki O, Kubo H, Takeuchi A, Uemura M, Kadowaki N. Nonclonal chromosomal alterations and poor survival in cytopenic patients without hematological malignancies. Mol Cytogenet. 2019;12:46. doi: 10.1186/s13039-019-0458-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Vanni R, Marras S, Faa G, Uccheddu A, Dal Cin P, Sciot R, Samson I, Van den Berghe H. Chromosome instability in elastofibroma. Cancer Genet Cytogenet. 1999;111(2):182–183. doi: 10.1016/s0165-4608(98)00243-x. [DOI] [PubMed] [Google Scholar]
- 194.McComb EN, Feely MG, Neff JR, Johansson SL, Nelson M, Bridge JA. Cytogenetic instability, predominantly involving chromosome 1, is characteristic of elastofibroma. Cancer Genet Cytogenet. 2001;126(1):68–72. doi: 10.1016/s0165-4608(00)00395-2. [DOI] [PubMed] [Google Scholar]
- 195.Bridge JA, Neff JR, Mouron BJ. Giant cell tumor of bone. Chromosomal analysis of 48 specimens and review of the literature. Cancer Genet Cytogenet. 1992;58(1):2–13. doi: 10.1016/0165-4608(92)90125-r. [DOI] [PubMed] [Google Scholar]
- 196.McComb EN, Johansson SL, Neff JR, Nelson M, Bridge JA. Chromosomal anomalies exclusive of telomeric associations in giant cell tumor of bone. Cancer Genet Cytogenet. 1996;88(2):163–166. doi: 10.1016/0165-4608(95)00362-2. [DOI] [PubMed] [Google Scholar]
- 197.Gorunova L, Vult von Steyern F, Storlazzi CT, Bjerkehagen B, Follerås G, Heim S, Mandahl N, Mertens F. Cytogenetic analysis of 101 giant cell tumors of bone: nonrandom patterns of telomeric associations and other structural aberrations. Genes Chromosomes Cancer. 2009;48(7):583–602. doi: 10.1002/gcc.20667. [DOI] [PubMed] [Google Scholar]
- 198.Zhao Y, Carter R, Natarajan S, Varn FS, Compton DA, Gawad C, Cheng C, Godek KM. Single-cell RNA sequencing reveals the impact of chromosomal instability on glioblastoma cancer stem cells. BMC Med Genomics. 2019;12(1):79. doi: 10.1186/s12920-019-0532-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Andor N, Lau BT, Catalanotti C, Sathe A, Kubit M, Chen J, Blaj C, Cherry A, Bangs CD, Grimes SM, Suarez CJ, Ji HP. Joint single cell DNA-seq and RNA-seq of gastric cancer cell lines reveals rules of in vitro evolution. NAR Genom Bioinform. 2020;2(2):lqaa016. doi: 10.1093/nargab/lqaa016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Jin Z, Huang W, Shen N, Li J, Wang X, Dong J, Park PJ, Xi R. Single-cell gene fusion detection by scFusion. Nat Commun. 2022;13(1):1084. doi: 10.1038/s41467-022-28661-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Sun G, Li Z, Rong D, Zhang H, Shi X, Yang W, Zheng W, Sun G, Wu F, Cao H, Tang W, Sun Y. Single-cell RNA sequencing in cancer: Applications, advances, and emerging challenges. Mol Ther Oncolytics. 2021;21:183–206. doi: 10.1016/j.omto.2021.04.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Tian L, Jabbari JS, Thijssen R, Gouil Q, Amarasinghe SL, Voogd O, Kariyawasam H, Du MRM, Schuster J, Wang C, Su S, Dong X, Law CW, Lucattini A, Prawer YDJ, Collar-Fernández C, Chung JD, Naim T, Chan A, Ly CH, Lynch GS, Ryall JG, Anttila CJA, Peng H, Anderson MA, Flensburg C, Majewski I, Roberts AW, Huang DCS, Clark MB, Ritchie ME. Comprehensive characterization of single-cell full-length isoforms in human and mouse with long-read sequencing. Genome Biol. 2021;22(1):310. doi: 10.1186/s13059-021-02525-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Zhang Y, Wang D, Peng M, Tang L, Ouyang J, Xiong F, Guo C, Tang Y, Zhou Y, Liao Q, Wu X, Wang H, Yu J, Li Y, Li X, Li G, Zeng Z, Tan Y, Xiong W. Single-cell RNA sequencing in cancer research. J Exp Clin Cancer Res. 2021;40(1):81. doi: 10.1186/s13046-021-01874-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Davidson NM, Chen Y, Sadras T, Ryland GL, Blombery P, Ekert PG, Göke J, Oshlack A. JAFFAL: detecting fusion genes with long-read transcriptome sequencing. Genome Biol. 2022;23(1):10. doi: 10.1186/s13059-021-02588-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Pon JR, Marra MA. Driver and passenger mutations in cancer. Annu Rev Pathol. 2015;10:25–50. doi: 10.1146/annurev-pathol-012414-040312. [DOI] [PubMed] [Google Scholar]
- 206.Buisson R, Langenbucher A, Bowen D, Kwan EE, Benes CH, Zou L, Lawrence MS. Passenger hotspot mutations in cancer driven by APOBEC3A and mesoscale genomic features. Science. 2019;364(6447):eaaw2872. doi: 10.1126/science.aaw2872. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Kumar S, Warrell J, Li S, McGillivray PD, Meyerson W, Salichos L, Harmanci A, Martinez-Fundichely A, Chan CWY, Nielsen MM, Lochovsky L, Zhang Y, Li X, Lou S, Pedersen JS, Herrmann C, Getz G, Khurana E, Gerstein MB. Passenger mutations in more than 2,500 cancer genomes: overall molecular functional impact and consequences. Cell. 2020;180(5):915–927.e16. doi: 10.1016/j.cell.2020.01.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Castro-Giner F, Ratcliffe P, Tomlinson I. The mini-driver model of polygenic cancer evolution. Nat Rev Cancer. 2015;15(11):680–685. doi: 10.1038/nrc3999. [DOI] [PubMed] [Google Scholar]
- 209.Bennett L, Howell M, Memon D, Smowton C, Zhou C, Miller CJ. Mutation pattern analysis reveals polygenic mini-drivers associated with relapse after surgery in lung adenocarcinoma. Sci Rep. 2018;8(1):14830. doi: 10.1038/s41598-018-33276-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Cui T, Leng F. Specific recognition of AT-rich DNA sequences by the mammalian high mobility group protein AT-hook 2: a SELEX study. Biochemistry. 2007;46(45):13059–13066. doi: 10.1021/bi701269s. [DOI] [PubMed] [Google Scholar]
- 211.Krahn N, Meier M, To V, Booy EP, McEleney K, O’Neil JD, McKenna SA, Patel TR, Stetefeld J. Nanoscale assembly of high-mobility group AT-Hook 2 protein with DNA replication fork. Biophys J. 2017;113(12):2609–2620. doi: 10.1016/j.bpj.2017.10.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212.Su L, Deng Z, Leng F. The Mammalian High Mobility Group protein AT-Hook 2 (HMGA2): Biochemical and biophysical properties, and its association with adipogenesis. Int J Mol Sci. 2020;21(10):3710. doi: 10.3390/ijms21103710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213.Borrmann L, Wilkening S, Bullerdiek J. The expression of HMGA genes is regulated by their 3’UTR. Oncogene. 2001;20(33):4537–4541. doi: 10.1038/sj.onc.1204577. [DOI] [PubMed] [Google Scholar]
- 214.Lee YS, Dutta A. The tumor suppressor microRNA let-7 represses the HMGA2 oncogene. Genes Dev. 2007;21(9):1025–1030. doi: 10.1101/gad.1540407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215.Mayr C, Hemann MT, Bartel DP. Disrupting the pairing between let-7 and Hmga2 enhances oncogenic transformation. Science. 2007;315(5818):1576–1579. doi: 10.1126/science.1137999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216.Kristjánsdóttir K, Fogarty EA, Grimson A. Systematic analysis of the Hmga2 3’ UTR identifies many independent regulatory sequences and a novel interaction between distal sites. RNA. 2015;21(7):1346–1360. doi: 10.1261/rna.051177.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217.Richter A, Hauschild G, Murua Escobar H, Nolte I, Bullerdiek J. Application of high-mobility-group-A proteins increases the proliferative activity of chondrocytes in vitro. Tissue Eng Part A. 2009;15(3):473–477. doi: 10.1089/ten.tea.2007.0308. [DOI] [PubMed] [Google Scholar]
- 218.Zaidi MR, Okada Y, Chada KK. Misexpression of full-length HMGA2 induces benign mesenchymal tumors in mice. Cancer Res. 2006;66(15):7453–7459. doi: 10.1158/0008-5472.CAN-06-0931. [DOI] [PubMed] [Google Scholar]
- 219.Morris SW, Valentine MB, Shapiro DN, Sublett JE, Deaven LL, Foust JT, Roberts WM, Cerretti DP, Look AT. Reassignment of the human CSF1 gene to chromosome 1p13-p21. Blood. 1991;78(8):2013–2020. [PubMed] [Google Scholar]
- 220.Saltman DL, Dolganov GM, Hinton LM, Lovett M. Reassignment of the human macrophage colony stimulating factor gene to chromosome 1p13-21. Biochem Biophys Res Commun. 1992;182(3):1139–1143. doi: 10.1016/0006-291x(92)91850-p. [DOI] [PubMed] [Google Scholar]
- 221.Dal Cin P, Sciot R, Samson I, De Smet L, De Wever I, Van Damme B, Van den Berghe H. Cytogenetic characterization of tenosynovial giant cell tumors (nodular tenosynovitis) Cancer Res. 1994;54(15):3986–3987. [PubMed] [Google Scholar]
- 222.Sciot R, Rosai J, Dal Cin P, de Wever I, Fletcher CD, Mandahl N, Mertens F, Mitelman F, Rydholm A, Tallini G, van den Berghe H, Vanni R, Willén H. Analysis of 35 cases of localized and diffuse tenosynovial giant cell tumor: a report from the Chromosomes and Morphology (CHAMP) study group. Mod Pathol. 1999;12(6):576–579. [PubMed] [Google Scholar]
- 223.Nilsson M, Höglund M, Panagopoulos I, Sciot R, Dal Cin P, Debiec-Rychter M, Mertens F, Mandahl N. Molecular cytogenetic mapping of recurrent chromosomal breakpoints in tenosynovial giant cell tumors. Virchows Arch. 2002;441(5):475–480. doi: 10.1007/s00428-002-0640-y. [DOI] [PubMed] [Google Scholar]
- 224.Brandal P, Bjerkehagen B, Heim S. Molecular cytogenetic characterization of tenosynovial giant cell tumors. Neoplasia. 2004;6(5):578–583. doi: 10.1593/neo.04202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 225.West RB, Rubin BP, Miller MA, Subramanian S, Kaygusuz G, Montgomery K, Zhu S, Marinelli RJ, De Luca A, Downs-Kelly E, Goldblum JR, Corless CL, Brown PO, Gilks CB, Nielsen TO, Huntsman D, van de Rijn M. A landscape effect in tenosynovial giant-cell tumor from activation of CSF1 expression by a translocation in a minority of tumor cells. Proc Natl Acad Sci U.S.A. 2006;103(3):690–695. doi: 10.1073/pnas.0507321103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226.Cupp JS, Miller MA, Montgomery KD, Nielsen TO, O’Connell JX, Huntsman D, van de Rijn M, Gilks CB, West RB. Translocation and expression of CSF1 in pigmented villonodular synovitis, tenosynovial giant cell tumor, rheumatoid arthritis and other reactive synovitides. Am J Surg Pathol. 2007;31(6):970–976. doi: 10.1097/PAS.0b013e31802b86f8. [DOI] [PubMed] [Google Scholar]
- 227.Nakayama S, Nishio J, Nakatani K, Nabeshima K, Yamamoto T. Giant cell tumor of tendon sheath with a t(1;1)(p13;p34) chromosomal translocation. Anticancer Res. 2020;40(8):4373–4377. doi: 10.21873/anticanres.14440. [DOI] [PubMed] [Google Scholar]
- 228.Kawasaki ES, Ladner MB, Wang AM, Van Arsdell J, Warren MK, Coyne MY, Schweickart VL, Lee MT, Wilson KJ, Boosman A. Molecular cloning of a complementary DNA encoding human macrophage-specific colony-stimulating factor (CSF-1) Science. 1985;230(4723):291–296. doi: 10.1126/science.2996129. [DOI] [PubMed] [Google Scholar]
- 229.Ladner MB, Martin GA, Noble JA, Nikoloff DM, Tal R, Kawasaki ES, White TJ. Human CSF-1: gene structure and alternative splicing of mRNA precursors. EMBO J. 1987;6(9):2693–2698. doi: 10.1002/j.1460-2075.1987.tb02561.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Bonafé N, Gilmore-Hebert M, Folk NL, Azodi M, Zhou Y, Chambers SK. Glyceraldehyde-3-phosphate dehydrogenase binds to the AU-Rich 3’ untranslated region of colony-stimulating factor-1 (CSF-1) messenger RNA in human ovarian cancer cells: possible role in CSF-1 posttranscriptional regulation and tumor phenotype. Cancer Res. 2005;65(9):3762–3771. doi: 10.1158/0008-5472.CAN-04-3954. [DOI] [PubMed] [Google Scholar]
- 231.Woo HH, László CF, Greco S, Chambers SK. Regulation of colony stimulating factor-1 expression and ovarian cancer cell behavior in vitro by miR-128 and miR-152. Mol Cancer. 2012;11:58. doi: 10.1186/1476-4598-11-58. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232.Woo HH, Baker T, Laszlo C, Chambers SK. Nucleolin mediates microRNA-directed CSF-1 mRNA deadenylation but increases translation of CSF-1 mRNA. Mol Cell Proteomics. 2013;12(6):1661–1677. doi: 10.1074/mcp.M112.025288. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233.Benner B, Good L, Quiroga D, Schultz TE, Kassem M, Carson WE, Cherian MA, Sardesai S, Wesolowski R. Pexidartinib, a novel small molecule CSF-1R inhibitor in use for tenosynovial giant cell tumor: a systematic review of pre-clinical and clinical development. Drug Des Devel Ther. 2020;14:1693–1704. doi: 10.2147/DDDT.S253232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234.Monestime S, Lazaridis D. Pexidartinib (TURALIO™): The first FDA-indicated systemic treatment for tenosynovial giant cell tumor. Drugs R D. 2020;20(3):189–195. doi: 10.1007/s40268-020-00314-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235.Fujiwara T, Yakoub MA, Chandler A, Christ AB, Yang G, Ouerfelli O, Rajasekhar VK, Yoshida A, Kondo H, Hata T, Tazawa H, Dogan Y, Moore MAS, Fujiwara T, Ozaki T, Purdue E, Healey JH. CSF1/CSF1R signaling inhibitor pexidartinib (PLX3397) reprograms tumor-associated macrophages and stimulates T-cell infiltration in the sarcoma microenvironment. Mol Cancer Ther. 2021;20(8):1388–1399. doi: 10.1158/1535-7163.MCT-20-0591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236.Smith BD, Kaufman MD, Wise SC, Ahn YM, Caldwell TM, Leary CB, Lu WP, Tan G, Vogeti L, Vogeti S, Wilky BA, Davis LE, Sharma M, Ruiz-Soto R, Flynn DL. Vimseltinib: A precision CSF1R therapy for tenosynovial giant cell tumors and diseases promoted by macrophages. Mol Cancer Ther. 2021;20(11):2098–2109. doi: 10.1158/1535-7163.MCT-21-0361. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237.Hess J, Angel P, Schorpp-Kistner M. AP-1 subunits: quarrel and harmony among siblings. J Cell Sci. 2004;117(Pt 25):5965–5973. doi: 10.1242/jcs.01589. [DOI] [PubMed] [Google Scholar]
- 238.Durchdewald M, Angel P, Hess J. The transcription factor Fos: a Janus-type regulator in health and disease. Histol Histopathol. 2009;24(11):1451–1461. doi: 10.14670/HH-24.1451. [DOI] [PubMed] [Google Scholar]
- 239.Acquaviva C, Brockly F, Ferrara P, Bossis G, Salvat C, Jariel-Encontre I, Piechaczyk M. Identification of a C-terminal tripeptide motif involved in the control of rapid proteasomal degradation of c-Fos proto-oncoprotein during the G(0)-to-S phase transition. Oncogene. 2001;20(51):7563–7572. doi: 10.1038/sj.onc.1204880. [DOI] [PubMed] [Google Scholar]
- 240.Acquaviva C, Bossis G, Ferrara P, Brockly F, Jariel-Encontre I, Piechaczyk M. Multiple degradation pathways for Fos family proteins. Ann N Y Acad Sci. 2002;973:426–434. doi: 10.1111/j.1749-6632.2002.tb04677.x. [DOI] [PubMed] [Google Scholar]
- 241.Shyu AB, Greenberg ME, Belasco JG. The c-fos transcript is targeted for rapid decay by two distinct mRNA degradation pathways. Genes Dev. 1989;3(1):60–72. doi: 10.1101/gad.3.1.60. [DOI] [PubMed] [Google Scholar]
- 242.Jooss KU, Funk M, Müller R. An autonomous N-terminal transactivation domain in Fos protein plays a crucial role in transformation. EMBO J. 1994;13(6):1467–1475. doi: 10.1002/j.1460-2075.1994.tb06401.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243.Schiavi SC, Wellington CL, Shyu AB, Chen CY, Greenberg ME, Belasco JG. Multiple elements in the c-fos protein-coding region facilitate mRNA deadenylation and decay by a mechanism coupled to translation. J Biol Chem. 1994;269(5):3441–3448. [PubMed] [Google Scholar]
- 244.Veyrune JL, Carillo S, Vié A, Blanchard JM. c-fos mRNA instability determinants present within both the coding and the 3’ non coding region link the degradation of this mRNA to its translation. Oncogene. 1995;11(10):2127–2134. [PubMed] [Google Scholar]
- 245.Dalgleish G, Veyrune JL, Blanchard JM, Hesketh J. mRNA localization by a 145-nucleotide region of the c-fos 3’—untranslated region. Links to translation but not stability. J Biol Chem. 2001;276(17):13593–13599. doi: 10.1074/jbc.M001141200. [DOI] [PubMed] [Google Scholar]
- 246.Bossis G, Ferrara P, Acquaviva C, Jariel-Encontre I, Piechaczyk M. c-Fos proto-oncoprotein is degraded by the proteasome independently of its own ubiquitinylation in vivo. Mol Cell Biol. 2003;23(20):7425–7436. doi: 10.1128/MCB.23.20.7425-7436.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247.Ferrara P, Andermarcher E, Bossis G, Acquaviva C, Brockly F, Jariel-Encontre I, Piechaczyk M. The structural determinants responsible for c-Fos protein proteasomal degradation differ according to the conditions of expression. Oncogene. 2003;22(10):1461–1474. doi: 10.1038/sj.onc.1206266. [DOI] [PubMed] [Google Scholar]
- 248.Lee HJ, Palkovits M, Young WS 3rd. miR-7b, a microRNA up-regulated in the hypothalamus after chronic hyperosmolar stimulation, inhibits Fos translation. Proc Natl Acad Sci U.S.A. 2006;103(42):15669–15674. doi: 10.1073/pnas.0605781103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249.Adler J, Reuven N, Kahana C, Shaul Y. c-Fos proteasomal degradation is activated by a default mechanism, and its regulation by NAD(P)H:quinone oxidoreductase 1 determines c-Fos serum response kinetics. Mol Cell Biol. 2010;30(15):3767–3778. doi: 10.1128/MCB.00899-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 250.Matoulkova E, Michalova E, Vojtesek B, Hrstka R. The role of the 3’ untranslated region in post-transcriptional regulation of protein expression in mammalian cells. RNA Biol. 2012;9(5):563–576. doi: 10.4161/rna.20231. [DOI] [PubMed] [Google Scholar]
- 251.Errico MC, Felicetti F, Bottero L, Mattia G, Boe A, Felli N, Petrini M, Bellenghi M, Pandha HS, Calvaruso M, Tripodo C, Colombo MP, Morgan R, Carè A. The abrogation of the HOXB7/PBX2 complex induces apoptosis in melanoma through the miR-221&222-c-FOS pathway. Int J Cancer. 2013;133(4):879–892. doi: 10.1002/ijc.28097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Konno Y, Dong P, Xiong Y, Suzuki F, Lu J, Cai M, Watari H, Mitamura T, Hosaka M, Hanley SJ, Kudo M, Sakuragi N. MicroRNA-101 targets EZH2, MCL-1 and FOS to suppress proliferation, invasion and stem cell-like phenotype of aggressive endometrial cancer cells. Oncotarget. 2014;5(15):6049–6062. doi: 10.18632/oncotarget.2157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253.Pellegrini KL, Han T, Bijol V, Saikumar J, Craciun FL, Chen WW, Fuscoe JC, Vaidya VS. MicroRNA-155 deficient mice experience heightened kidney toxicity when dosed with cisplatin. Toxicol Sci. 2014;141(2):484–492. doi: 10.1093/toxsci/kfu143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254.Vougioukalaki M, Kanellis DC, Gkouskou K, Eliopoulos AG. Tpl2 kinase signal transduction in inflammation and cancer. Cancer Lett. 2011;304(2):80–89. doi: 10.1016/j.canlet.2011.02.004. [DOI] [PubMed] [Google Scholar]
- 255.Xu D, Matsumoto ML, McKenzie BS, Zarrin AA. TPL2 kinase action and control of inflammation. Pharmacol Res. 2018;129:188–193. doi: 10.1016/j.phrs.2017.11.031. [DOI] [PubMed] [Google Scholar]
- 256.Njunge LW, Estania AP, Guo Y, Liu W, Yang L. Tumor progression locus 2 (TPL2) in tumor-promoting inflammation, tumorigenesis and tumor immunity. Theranostics. 2020;10(18):8343–8364. doi: 10.7150/thno.45848. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 257.Miyoshi J, Higashi T, Mukai H, Ohuchi T, Kakunaga T. Structure and transforming potential of the human cot oncogene encoding a putative protein kinase. Mol Cell Biol. 1991;11(8):4088–4096. doi: 10.1128/mcb.11.8.4088-4096.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 258.Aoki M, Hamada F, Sugimoto T, Sumida S, Akiyama T, Toyoshima K. The human cot proto-oncogene encodes two protein serine/threonine kinases with different transforming activities by alternative initiation of translation. J Biol Chem. 1993;268(30):22723–22732. [PubMed] [Google Scholar]
- 259.Kent WJ. BLAT—the BLAST-like alignment tool. Genome Res. 2002;12(4):656–664. doi: 10.1101/gr.229202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 260.Clark AM, Reynolds SH, Anderson M, Wiest JS. Mutational activation of the MAP3K8 protooncogene in lung cancer. Genes Chromosomes Cancer. 2004;41(2):99–108. doi: 10.1002/gcc.20069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 261.Gándara ML, López P, Hernando R, Castaño JG, Alemany S. The COOH-terminal domain of wild-type Cot regulates its stability and kinase specific activity. Mol Cell Biol. 2003;23(20):7377–7390. doi: 10.1128/MCB.23.20.7377-7390.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262.Ceci JD, Patriotis CP, Tsatsanis C, Makris AM, Kovatch R, Swing DA, Jenkins NA, Tsichlis PN, Copeland NG. Tpl-2 is an oncogenic kinase that is activated by carboxy-terminal truncation. Genes Dev. 1997;11(6):688–700. doi: 10.1101/gad.11.6.688. [DOI] [PubMed] [Google Scholar]
- 263.Newman S, Fan L, Pribnow A, Silkov A, Rice SV, Lee S, Shao Y, Shaner B, Mulder H, Nakitandwe J, Shurtleff S, Azzato EM, Wu G, Zhou X, Barnhill R, Easton J, Nichols KE, Ellison DW, Downing JR, Pappo A, Potter PM, Zhang J, Bahrami A. Clinical genome sequencing uncovers potentially targetable truncations and fusions of MAP3K8 in spitzoid and other melanomas. Nat Med. 2019;25(4):597–602. doi: 10.1038/s41591-019-0373-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 264.Lehmann BD, Shaver TM, Johnson DB, Li Z, Gonzalez-Ericsson PI, Sánchez V, Shyr Y, Sanders ME, Pietenpol JA. Identification of targetable recurrent MAP3K8 rearrangements in melanomas lacking known driver mutations. Mol Cancer Res. 2019;17(9):1842–1853. doi: 10.1158/1541-7786.MCR-19-0257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 265.Quan VL, Zhang B, Mohan LS, Shi K, Isales MC, Panah E, Taxter TJ, Beaubier N, White K, Gerami P. Activating structural alterations in MAPK genes are distinct genetic drivers in a unique subgroup of spitzoid neoplasms. Am J Surg Pathol. 2019;43(4):538–548. doi: 10.1097/PAS.0000000000001213. [DOI] [PubMed] [Google Scholar]
- 266.Houlier A, Pissaloux D, Masse I, Tirode F, Karanian M, Pincus LB, McCalmont TH, LeBoit PE, Bastian BC, Yeh I, de la Fouchardière A. Melanocytic tumors with MAP3K8 fusions: report of 33 cases with morphological-genetic correlations. Mod Pathol. 2020;33(5):846–857. doi: 10.1038/s41379-019-0384-8. [DOI] [PubMed] [Google Scholar]
- 267.Quan VL, Zhang B, Zhang Y, Mohan LS, Shi K, Wagner A, Kruse L, Taxter T, Beaubier N, White K, Zou L, Gerami P. Integrating next-generation sequencing with morphology improves prognostic and biologic classification of spitz neoplasms. J Invest Dermatol. 2020;140(8):1599–1608. doi: 10.1016/j.jid.2019.12.031. [DOI] [PubMed] [Google Scholar]
- 268.Ogawa E, Maruyama M, Kagoshima H, Inuzuka M, Lu J, Satake M, Shigesada K, Ito Y. PEBP2/PEA2 represents a family of transcription factors homologous to the products of the Drosophila runt gene and the human AML1 gene. Proc Natl Acad Sci U.S.A. 1993;90(14):6859–6863. doi: 10.1073/pnas.90.14.6859. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 269.Miyoshi H, Shimizu K, Kozu T, Maseki N, Kaneko Y, Ohki M. t(8;21) breakpoints on chromosome 21 in acute myeloid leukemia are clustered within a limited region of a single gene, AML1. Proc Natl Acad Sci U.S.A. 1991;88(23):10431–10434. doi: 10.1073/pnas.88.23.10431. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270.Miyoshi H, Ohira M, Shimizu K, Mitani K, Hirai H, Imai T, Yokoyama K, Soeda E, Ohki M. Alternative splicing and genomic structure of the AML1 gene involved in acute myeloid leukemia. Nucleic Acids Res. 1995;23(14):2762–2769. doi: 10.1093/nar/23.14.2762. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 271.Tanaka T, Tanaka K, Ogawa S, Kurokawa M, Mitani K, Nishida J, Shibata Y, Yazaki Y, Hirai H. An acute myeloid leukemia gene, AML1, regulates hemopoietic myeloid cell differentiation and transcriptional activation antagonistically by two alternative spliced forms. EMBO J. 1995;14(2):341–350. doi: 10.1002/j.1460-2075.1995.tb07008.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272.Liu X, Zhang Q, Zhang DE, Zhou C, Xing H, Tian Z, Rao Q, Wang M, Wang J. Overexpression of an isoform of AML1 in acute leukemia and its potential role in leukemogenesis. Leukemia. 2009;23(4):739–745. doi: 10.1038/leu.2008.350. [DOI] [PubMed] [Google Scholar]
- 273.Watanabe-Okochi N, Kitaura J, Ono R, Harada H, Harada Y, Komeno Y, Nakajima H, Nosaka T, Inaba T, Kitamura T. AML1 mutations induced MDS and MDS/AML in a mouse BMT model. Blood. 2008;111(8):4297–4308. doi: 10.1182/blood-2007-01-068346. [DOI] [PubMed] [Google Scholar]
- 274.Gerritsen M, Yi G, Tijchon E, Kuster J, Schuringa JJ, Martens JHA, Vellenga E. RUNX1 mutations enhance self-renewal and block granulocytic differentiation in human in vitro models and primary AMLs. Blood Adv. 2019;3(3):320–332. doi: 10.1182/bloodadvances.2018024422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 275.Dowdy CR, Xie R, Frederick D, Hussain S, Zaidi SK, Vradii D, Javed A, Li X, Jones SN, Lian JB, van Wijnen AJ, Stein JL, Stein GS. Definitive hematopoiesis requires Runx1 C-terminal-mediated subnuclear targeting and transactivation. Hum Mol Genet. 2010;19(6):1048–1057. doi: 10.1093/hmg/ddp568. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 276.Kogure Y, Kameda T, Koya J, Yoshimitsu M, Nosaka K, Yasunaga JI, Imaizumi Y, Watanabe M, Saito Y, Ito Y, McClure MB, Tabata M, Shingaki S, Yoshifuji K, Chiba K, Okada A, Kakiuchi N, Nannya Y, Kamiunten A, Tahira Y, Akizuki K, Sekine M, Shide K, Hidaka T, Kubuki Y, Kitanaka A, Hidaka M, Nakano N, Utsunomiya A, Sica RA, Acuna-Villaorduna A, Janakiram M, Shah U, Ramos JC, Shibata T, Takeuchi K, Takaori-Kondo A, Miyazaki Y, Matsuoka M, Ishitsuka K, Shiraishi Y, Miyano S, Ogawa S, Ye BH, Shimoda K, Kataoka K. Whole-genome landscape of adult T-cell leukemia/lymphoma. Blood. 2022;139(7):967–982. doi: 10.1182/blood.2021013568. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 277.Daisuke H, Kato H, Fukumura K, Mayeda A, Miyagi Y, Seiki M, Koshikawa N. Novel LAMC2 fusion protein has tumor-promoting properties in ovarian carcinoma. Cancer Sci. 2021;112(12):4957–4967. doi: 10.1111/cas.15149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 278.Dupain C, Gracia C, Harttrampf AC, Rivière J, Geoerger B, Massaad-Massade L. Newly identified LMO3-BORCS5 fusion oncogene in Ewing sarcoma at relapse is a driver of tumor progression. Oncogene. 2019;38(47):7200–7215. doi: 10.1038/s41388-019-0914-3. [DOI] [PubMed] [Google Scholar]
- 279.Scharf S, Zech J, Bursen A, Schraets D, Oliver PL, Kliem S, Pfitzner E, Gillert E, Dingermann T, Marschalek R. Transcription linked to recombination: a gene-internal promoter coincides with the recombination hot spot II of the human MLL gene. Oncogene. 2007;26(10):1361–1371. doi: 10.1038/sj.onc.1209948. [DOI] [PubMed] [Google Scholar]
- 280.Bianchini L, Birtwisle L, Saâda E, Bazin A, Long E, Roussel JF, Michiels JF, Forest F, Dani C, Myklebost O, Birtwisle-Peyrottes I, Pedeutour F. Identification of PPAP2B as a novel recurrent translocation partner gene of HMGA2 in lipomas. Genes Chromosomes Cancer. 2013;52(6):580–590. doi: 10.1002/gcc.22055. [DOI] [PubMed] [Google Scholar]
- 281.Geurts JM, Schoenmakers EF, Röijer E, Stenman G, Van de Ven WJ. Expression of reciprocal hybrid transcripts of HMGIC and FHIT in a pleomorphic adenoma of the parotid gland. Cancer Res. 1997;57(1):13–17. [PubMed] [Google Scholar]
- 282.Storlazzi CT, Albano F, Locunsolo C, Lonoce A, Funes S, Guastadisegni MC, Cimarosto L, Impera L, D’Addabbo P, Panagopoulos I, Specchia G, Rocchi M. t(3;12)(q26;q14) in polycythemia vera is associated with upregulation of the HMGA2 gene. Leukemia. 2006;20(12):2190–2192. doi: 10.1038/sj.leu.2404418. [DOI] [PubMed] [Google Scholar]
- 283.Petit MM, Swarts S, Bridge JA, Van de Ven WJ. Expression of reciprocal fusion transcripts of the HMGIC and LPP genes in parosteal lipoma. Cancer Genet Cytogenet. 1998;106(1):18–23. doi: 10.1016/s0165-4608(98)00038-7. [DOI] [PubMed] [Google Scholar]
- 284.Hodgson A, Swanson D, Tang S, Dickson BC, Turashvili G. Gene fusions characterize a subset of uterine cellular leiomyomas. Genes Chromosomes Cancer. 2020 doi: 10.1002/gcc.22888. [DOI] [PubMed] [Google Scholar]
- 285.Komuro A, Raja E, Iwata C, Soda M, Isogaya K, Yuki K, Ino Y, Morikawa M, Todo T, Aburatani H, Suzuki H, Ranjit M, Natsume A, Mukasa A, Saito N, Okada H, Mano H, Miyazono K, Koinuma D. Identification of a novel fusion gene HMGA2-EGFR in glioblastoma. Int J Cancer. 2018;142(8):1627–1639. doi: 10.1002/ijc.31179. [DOI] [PubMed] [Google Scholar]
- 286.Velagaleti GV, Tonk VS, Hakim NM, Wang X, Zhang H, Erickson-Johnson MR, Medeiros F, Oliveira AM. Fusion of HMGA2 to COG5 in uterine leiomyoma. Cancer Genet Cytogenet. 2010;202(1):11–16. doi: 10.1016/j.cancergencyto.2010.06.002. [DOI] [PubMed] [Google Scholar]
- 287.Mine N, Kurose K, Nagai H, Doi D, Ota Y, Yoneyama K, Konishi H, Araki T, Emi M. Gene fusion involving HMGIC is a frequent aberration in uterine leiomyomas. J Hum Genet. 2001;46(7):408–412. doi: 10.1007/s100380170059. [DOI] [PubMed] [Google Scholar]
- 288.Afshari MK, Fehr A, Nevado PT, Andersson MK, Stenman G. Activation of PLAG1 and HMGA2 by gene fusions involving the transcriptional regulator gene NFIB. Genes Chromosomes Cancer. 2020;59(11):652–660. doi: 10.1002/gcc.22885. [DOI] [PubMed] [Google Scholar]
- 289.Lee MY, da Silva B, Ramirez DC, Maki RG. Novel HMGA2-YAP1 fusion gene in aggressive angiomyxoma. BMJ Case Rep. 2019;12(5):e227475. doi: 10.1136/bcr-2018-227475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 290.Panagopoulos I, Bjerkehagen B, Gorunova L, Taksdal I, Heim S. Rearrangement of chromosome bands 12q14~15 causing HMGA2-SOX5 gene fusion and HMGA2 expression in extraskeletal osteochondroma. Oncol Rep. 2015;34(2):577–584. doi: 10.3892/or.2015.4035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 291.Panagopoulos I, Gorunova L, Brunetti M, Agostini A, Andersen HK, Lobmaier I, Bjerkehagen B, Heim S. Genetic heterogeneity in leiomyomas of deep soft tissue. Oncotarget. 2017;8(30):48769–48781. doi: 10.18632/oncotarget.17953. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 292.Panagopoulos I, Gorunova L, Bjerkehagen B, Lobmaier I, Heim S. The recurrent chromosomal translocation t(12;18)(q14~15;q12~21) causes the fusion gene HMGA2-SETBP1 and HMGA2 expression in lipoma and osteochondrolipoma. Int J Oncol. 2015;47(3):884–890. doi: 10.3892/ijo.2015.3099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 293.Persson F, Andrén Y, Winnes M, Wedell B, Nordkvist A, Gudnadottir G, Dahlenfors R, Sjögren H, Mark J, Stenman G. High-resolution genomic profiling of adenomas and carcinomas of the salivary glands reveals amplification, rearrangement, and fusion of HMGA2. Genes Chromosomes Cancer. 2009;48(1):69–82. doi: 10.1002/gcc.20619. [DOI] [PubMed] [Google Scholar]
- 294.Panagopoulos I, Bjerkehagen B, Gorunova L, Berner JM, Boye K, Heim S. Several fusion genes identified by whole transcriptome sequencing in a spindle cell sarcoma with rearrangements of chromosome arm 12q and MDM2 amplification. Int J Oncol. 2014;45(5):1829–1836. doi: 10.3892/ijo.2014.2605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 295.Brennan CW, Verhaak RG, McKenna A, Campos B, Noushmehr H, Salama SR, Zheng S, Chakravarty D, Sanborn JZ, Berman SH, Beroukhim R, Bernard B, Wu CJ, Genovese G, Shmulevich I, Barnholtz-Sloan J, Zou L, Vegesna R, Shukla SA, Ciriello G, Yung WK, Zhang W, Sougnez C, Mikkelsen T, Aldape K, Bigner DD, Van Meir EG, Prados M, Sloan A, Black KL, Eschbacher J, Finocchiaro G, Friedman W, Andrews DW, Guha A, Iacocca M, O’Neill BP, Foltz G, Myers J, Weisenberger DJ, Penny R, Kucherlapati R, Perou CM, Hayes DN, Gibbs R, Marra M, Mills GB, Lander E, Spellman P, Wilson R, Sander C, Weinstein J, Meyerson M, Gabriel S, Laird PW, Haussler D, Getz G, Chin L, TCGA Research Network The somatic genomic landscape of glioblastoma. Cell. 2013;155(2):462–477. doi: 10.1016/j.cell.2013.09.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 296.Panagopoulos I, Gorunova L, Andersen K, Lobmaier I, Micci F, Heim S. Fusion of high mobility group AT-hook 2 gene (HMGA2) with the chromosome 12 open reading frame 42 gene (C12orf42) in an aggressive angiomyxoma with del(12)(q14q23) as the sole cytogenetic anomaly. Cancer Genomics Proteomics. 2022;19(5):576–583. doi: 10.21873/cgp.20342. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 297.Kazmierczak B, Hennig Y, Wanschura S, Rogalla P, Bartnitzke S, Van de Ven W, Bullerdiek J. Description of a novel fusion transcript between HMGI-C, a gene encoding for a member of the high mobility group proteins, and the mitochondrial aldehyde dehydrogenase gene. Cancer Res. 1995;55(24):6038–6039. [PubMed] [Google Scholar]
- 298.Brahmi M, Alberti L, Tirode F, Karanian M, Eberst L, Pissaloux D, Cassier P, Blay JY. Complete response to CSF1R inhibitor in a translocation variant of teno-synovial giant cell tumor without genomic alteration of the CSF1 gene. Ann Oncol. 2018;29(6):1488–1489. doi: 10.1093/annonc/mdy129. [DOI] [PubMed] [Google Scholar]
- 299.Agaimy A, Michal M, Stoehr R, Ferrazzi F, Fabian P, Michal M, Franchi A, Haller F, Folpe AL, Kösemehmetoğlu K. Recurrent novel HMGA2-NCOR2 fusions characterize a subset of keratin-positive giant cell-rich soft tissue tumors. Mod Pathol. 2021;34(8):1507–1520. doi: 10.1038/s41379-021-00789-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 300.Panagopoulos I, Andersen K, Gorunova L, Lund-Iversen M, Lobmaier I, Heim S. Recurrent fusion of the genes for high-mobility group AT-hook 2 (HMGA2) and nuclear receptor co-repressor 2 (NCOR2) in osteoclastic giant cell-rich tumors of bone. Cancer Genomics Proteomics. 2022;19(2):163–177. doi: 10.21873/cgp.20312. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 301.Petit MM, Schoenmakers EF, Huysmans C, Geurts JM, Mandahl N, Van de Ven WJ. LHFP, a novel translocation partner gene of HMGIC in a lipoma, is a member of a new family of LHFP-like genes. Genomics. 1999;57(3):438–441. doi: 10.1006/geno.1999.5778. [DOI] [PubMed] [Google Scholar]
- 302.Takahashi T, Nagai N, Oda H, Ohama K, Kamada N, Miyagawa K. Evidence for RAD51L1/HMGIC fusion in the pathogenesis of uterine leiomyoma. Genes Chromosomes Cancer. 2001;30(2):196–201. doi: 10.1002/1098-2264(2000)9999:9999<::aid-gcc1078>3.0.co;2-8. [DOI] [PubMed] [Google Scholar]
- 303.Quade BJ, Weremowicz S, Neskey DM, Vanni R, Ladd C, Dal Cin P, Morton CC. Fusion transcripts involving HMGA2 are not a common molecular mechanism in uterine leiomyomata with rearrangements in 12q15. Cancer Res. 2003;63(6):1351–1358. [PubMed] [Google Scholar]
- 304.Panagopoulos I, Andersen K, Gorunova L, Eilert-Olsen M, Lund-Iversen M, Wessel-Aas T, Lloret I, Micci F, Heim S. Presence of a t(12;18)(q14;q21) chromosome translocation and fusion of the genes for high-mobility group AT-Hook 2 (HMGA2) and WNT inhibitory factor 1 (WIF1) in infrapatellar fat pad cells from a patient with Hoffa’s disease. Cancer Genomics Proteomics. 2022;19(5):584–590. doi: 10.21873/cgp.20343. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 305.Kazmierczak B, Pohnke Y, Bullerdiek J. Fusion transcripts between the HMGIC gene and RTVL-H-related sequences in mesenchymal tumors without cytogenetic aberrations. Genomics. 1996;38(2):223–226. doi: 10.1006/geno.1996.0619. [DOI] [PubMed] [Google Scholar]







