Next Article in Journal
A Perspective on Natural and Nature-Inspired Small Molecules Targeting Phosphodiesterase 9 (PDE9): Chances and Challenges against Neurodegeneration
Next Article in Special Issue
Targeting the Copper Transport System to Improve Treatment Efficacies of Platinum-Containing Drugs in Cancer Chemotherapy
Previous Article in Journal
Contact System Activation in Plasma from Dengue Patients Might Harness Endothelial Virus Replication through the Signaling of Bradykinin Receptors
Previous Article in Special Issue
NLRP1 Functions Downstream of the MAPK/ERK Signaling via ATF4 and Contributes to Acquired Targeted Therapy Resistance in Human Metastatic Melanoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting Loss of Heterozygosity: A Novel Paradigm for Cancer Therapy

Science for Life Laboratory, Department of Immunology, Genetics and Pathology, Uppsala University, SE-75185 Uppsala, Sweden
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2021, 14(1), 57; https://0-doi-org.brum.beds.ac.uk/10.3390/ph14010057
Submission received: 14 December 2020 / Revised: 8 January 2021 / Accepted: 9 January 2021 / Published: 13 January 2021
(This article belongs to the Special Issue Novel Therapeutic Targets in Cancer)

Abstract

:
Loss of heterozygosity (LOH) is a common genetic event in the development of cancer. In certain tumor types, LOH can affect more than 20% of the genome, entailing loss of allelic variation in thousands of genes. This reduction of heterozygosity creates genetic differences between tumor and normal cells, providing opportunities for development of novel cancer therapies. Here, we review and summarize (1) mutations associated with LOH on chromosomes which have been shown to be promising biomarkers of cancer risk or the prediction of clinical outcomes in certain types of tumors; (2) loci undergoing LOH that can be targeted for development of novel anticancer drugs as well as (3) LOH in tumors provides up-and-coming possibilities to understand the underlying mechanisms of cancer evolution and to discover novel cancer vulnerabilities which are worth a further investigation in the near future.

1. Introduction

Several different somatic genetic and epigenetic processes contribute to the development of cancer, including copy number alterations, deletions, rearrangements or translocations of certain genes, somatic point mutations, and hypermethylation of promoters [1]. Loss of heterozygosity (LOH) was originally discovered using polymorphic markers which were heterozygous in germline DNA but homozygous in the tumor, and is common in the evolution of many cancer types [2]. In general, there are two types of LOH, (1) LOH with copy number losses (CNL-LOH), with a typical example of being losing the wildtype allele of a tumor suppressor, and (2) copy number neutral LOH (CNN-LOH), exemplified by the presence of two mutant alleles of WT1 (11p), FLT3 (13q), CEBPA (19q) and RUNX1 (21q) which resulted in a growth advantage in tumors, such as in leukemia [3]. A complete or partial deletion of a chromosome leads to CNL-LOH, while CNN-LOH is mainly caused by acquired uniparental disomy (UPD) and gene conversion, and occurs without net change in the copy number [4,5] (Figure 1). In principle, the presence of two mutant alleles generated by CNN-LOH could lead to an alteration at the gene expression level, however, it has been recently shown that there are allelic differences in gene expression [6], indicating that the exact expression level could be associated with specific gene expression patterns and regulated by other mechanisms. This review will primarily focus on LOH with copy number loss (CNL-LOH), as it is a common phenomenon in cancer and more thoroughly investigated in comparison to copy number neutral LOH. The process of LOH is highly associated with a reduction of copy number of wild-type allele in individuals thus increases the impact of a genotype which becomes dominant when the function of wild-type allele is lost under LOH [4,7]. In certain types of cancers, such as colorectal carcinomas, LOH can affect more than 20% of the genome [8], causing loss of one allele of thousands of genes. Thus, LOH decreases the level of heterozygosity in cancer cells and creates significantly distinct genetic characters between tumor and normal cells.

2. Loss of Tumor Suppressor Genes by LOH

The function loss of tumor suppressor gene frequently involves the process of LOH, whereby a wild-type allele is lost, leaving only an inactivated allele in the cancer genome. A good example is the loss or inactivation of both alleles of the retinoblastoma gene (Rb) in retinoblastomas, which suggested that the Rb gene belongs to a class of human cancer genes which function as tumor suppressors [2,9]. Following this finding, numerous candidate tumor suppressors were discovered by characterizing sites of prevalent LOH in human cancers [10,11]. For example, as a consequence of LOH, the tumor suppressor gene TP53 is inactivated, contributing to the development of many cancer types such as breast [12], lung and stomach [13] as well as chronic lymphocytic leukemia [14]. The LOH at 10q23, where the PTEN gene resides, has been linked to the development of breast cancer [15]. Further, there are inherited cancer predisposition syndromes which have germline mutations in tumor suppressor genes under LOH, such as BRCA1 at 17q21. Women carrying germline LOH at BRCA1 have an 85% lifetime risk of developing breast cancer and a greatly elevated risk of ovarian cancer [16,17]. DPC4 (for deleted in pancreatic cancer, locus 4, SMAD4) on chromosome 18q was identified as a tumor suppressor gene in pancreatic, colon, bladder, biliary tumors as well as head and neck tumors [18,19,20]. LOH at the DPC4 locus was detected in 51% of sporadic colon cancers [21]. Several more recently discovered tumor suppressor genes have also been reported to undergo LOH. For example, tripartite motif containing 3 (TRIM3) at 11p15.5 is lost in ~20% of glioblastomas (GBM) [22]. The Cut homeobox 1 (CUX1) gene at 7q22.1, suggested as both a potential tumor suppressor and an oncogene [23], is a target of LOH in many cancers. Taken together, loss of function of tumor suppressor genes frequently involves LOH and is linked to the development of cancer.

3. LOH Leads to Loss of Non-Driver Genes and Endows Cancer Cells with Unique Vulnerabilities

Beyond the direct consequences of LOH on bona fide cancer genes, particularly tumor suppressor genes, numerous non-driver genes located nearby or distally on the same chromosome arm may also undergo LOH. The establishment of LOH maps by cytogenetics [24] or, more recently, by genome-wide copy number analyses and genome sequencing, have enabled identification of chromosomal regions lost in different tumor types [25,26,27,28]. A study of ~105 LOH events in 363 glioblastoma and 513 ovarian cancer samples revealed that LOH selectively occurs in early replicating regions, especially near RNA pol II-bound transcription start sites [29]. Together, these and other studies demonstrate that LOH in cancer genomes is not only affecting tumor suppressor genes but also numerous non-driver genes.
When one allele of an essential non-driver gene undergoes LOH in cancer cells, cancer cells should not be able to survive if the remaining allele is further lost or inhibited, whereas normal cells will be able to survive relying solely on the remaining allele [25], leading to an unique vulnerability in cancer cells. For example, several studies have together indicated that the process of LOH can affect the extent of ROS tolerance [30]. An LOH event relevant to ROS balance affects the Oct1/Pou2f1 gene. The transcription factor Oct1/Pou2f1 is involved in a wide variety of functions in organism development, particularly relating to the development of neuroendocrine system [31]. Oct1 promotes glycolytic metabolism and mitotic stability and plays important roles in stress responses [32,33], and loss of one or both Oct1 alleles has been associated with an upregulation of oxidative metabolism and increased levels of reactive oxygen species (ROS), thus inducing a coordinate metabolic shift and hypersensitive to oxidative stress [32,33]. Glutathione peroxidase (GPx-1) is a selenium-containing antioxidant protein mediating the reaction of hydrogen peroxide to water, using reducing equivalents from glutathione. A higher frequency of loss of heterozygosity at the GPx-1 locus at 3p2 was first observed in lung tumors and lung tumor-derived cell lines [34]. Further studies revealed that GPx-1 LOH is prevalent also in breast, colon, kidney and head and neck cancers [35]. Clinical investigations have revealed that LOH of non-driver genes could engender unique vulnerabilities in rare cancer types. Pheochromocytomas and paragangliomas are rare neuroendocrine tumors [36], and ~40% of tumors arise in patients with germline LOH of the succinate dehydrogenase (SDHx) genes. The SDHx protein is located on the inner membrane of mitochondria and plays an important role in cellular energy metabolism by linking the Krebs cycle to mitochondrial oxidative phosphorylation [37,38]. Taken together, loss of non-driver genes by LOH can cause unique vulnerabilities in cancer cells and thereby provide a novel class of therapeutic targets for cancer drug discovery and personalized medicine.

4. Loss of Heterozygosity Provides Novel Therapeutic Targets for Cancer Treatment

Targeted anti-cancer treatments typically rely on genetic differences between cancer and normal cells to achieve a specific inhibitory effect on cancer cells. Targets include oncogenes which are activated by mutation, tumor suppressor genes which are inactivated by mutation, or perturbed cell signaling in the maintenance of genome integrity or regular cellular metabolism [39,40]. However, the era of cancer genomics has revealed few novel targetable oncogenes but many tumor suppressor genes that, unfortunately, are difficult to target therapeutically [1,41]. For example, the TP53 tumor suppressor gene is mutated in >50% of human tumors [42]. Among the 393 amino acids of P53, sequences at positions 5–28 have p53 transcription activity and pathogenic mutations are enriched at R248Q, R273H, and R282W which affect DNA binding, or in R175H, Y220C, G245S, and R249S which are called conformational mutations [43,44]. The mechanism of p53 in cancer development has been thoroughly investigated, with more than 70K papers published since 1979. Nevertheless, there is still no drug targeting p53 available in the clinic, mainly due to the lack of a good binding site to serve as a direct target in the mutant p53 structure. Another example is MYC, which is amplified in nearly 14% of cases of at least 12 cancer types [45,46] and approximately 50% of patients with high-risk disease [47,48]. Currently, several MYC inhibitors have been identified from phenotypic screens, including 10058-F4, atorvastatin and the recently discovered Omomyc [49,50,51], however, all of them still have a long way to go from bench to clinic [52,53].
Beyond the successful use of immunotherapy, it now appears necessary to go beyond the oncogenes and suppressor genes in the hunt for novel drug targets based on genomic aberrations. Thus, non-driver genes represent an alternative target class that merits further exploration (Figure 2). In 1999, the possibility of targeting LOH for anticancer therapy was first explored. The RPA70 gene at 17p13.3, which encodes the 70 kDa subunit of human replication protein A which plays a vital role in DNA replication, homologous recombination, and nucleotide excision repair in vitro [54] and any disturbance to the role of RPA70 is lethal to the cell [55]. RPA70 undergoes LOH in 44% of colon cancers, 58% of ovarian cancers, 20% of breast cancers, and 27% of non-small cell lung carcinomas. The oligonucleotide ISIS 1278 targeting segments of RPA70 mRNA containing variants effectively inhibited survival of cells expressing only the RPA70 mRNA with the exact complementary sequence, but was less effective in cells expressing the mismatched target, suggesting that developing anticancer agents based on normal genetic variation under LOH in essential genes is a feasible strategy for anticancer therapy [56].
Recently, the Beroukhim group have identified 5664 variants in 1278 essential genes that undergo LOH in cancer and pointed out that allele-specific inactivation of either of two essential genes (PRIM1 and EXOSC8), which have been rigorously validated as genetic dependencies in cancer, is lethal in cancer cells [25,57,58]. PRIM1 encodes the catalytic subunit of DNA primase and contains two common SNPs, of which one (rs2277339) undergoes frequent LOH across 33 different cancer types, occurring in 21% of lung adenocarcinomas, 18% of ovarian cancers, and 17% of pancreatic cancers. The other important gene reported in this study is EXOSC8 encoding Rrp43, a component of the RNA exosome which is an essential multi-protein complex regulating RNA processing and degradation [58]. The position of rs117135638 in EXOSC8 undergoes LOH in 29% of cancers, including 72% of lung squamous cell carcinomas, 62% of ovarian cancers, 46% of lung adenocarcinomas, and 40% of breast cancers. Allele-specific disruption using sgRNA targeting rs2277339 in PRIM1 and rs117135638 in EXOSC8 strongly reduces the growth in cells containing targeted LOH allele of rs2277339 or rs117135638, while cells harboring the non-targeted allele remain intact [25]. Although there were no followed drug screens by targeting the products of the two essential genes, the authors discussed the possibility of developing allele-specific small molecule inhibitors using the canSAR protein annotation tool [59,60] which provides a prioritization of targets based on general drugability, suggesting that cancer vulnerabilities generating from LOH represent viable targets for novel anticancer drug development.
We recently demonstrated that cancer cells will develop unique vulnerabilities when LOH events affect non-essential genes. We analyzed variants obtained from 1092 clinical samples in the 1000 Genomes project and selected 23,532 non-synonymous small nucleotide variants (SNVs) in functional protein domains with LOH frequency ≥0.5% for a further confirmation and 45 common non-synonymous small nucleotide variants (nsSNVs) near the catalytic sites of 17 enzymes that frequently undergo LOH were identified. After a series of strict filtration and selection, the gastrointestinal drug metabolic enzyme N-acetyltransferase 2 (NAT2) at 8p22, which is frequently lost in colorectal cancers, was confirmed as a top target. It is worth noting that, the following proof of concept experiments suggested that NAT2 harboring mutant allele (NAT2*6A, rs1799930) has a 10-fold reduced activity comparing with the wild type allele (NAT2*13A). From a total of 189,018 compounds, we identified 6-(4-aminophenyl)-N-(3,4,5-trimethoxyphenyl) pyrazin-2-amine (APA) which preferentially kills cells expressing slow NAT2 (NAT2*6A, rs1799930) [26]. Comparing to the previously described efforts, this represented the first compound targeting cancer vulnerabilities stemming from LOH of non-cancer genes which supports that LOH targeting is a practically exploitable paradigm in the discovery of novel drugs for cancer treatments.
It is noteworthy that while only one or two target genes were selected for further validation in these studies [25,26], several other loci were reported within each respective study. These are no doubt rich resources for pursuing novel drug targets among the cancer vulnerabilities associated with LOH.

5. Loci Undergoing Loss of Heterozygosity as Clinical Biomarkers

Loss of heterozygosity is a common genetic event in the development of many cancer types and occurs in every step of tumorigenesis [2]. At the same time, several studies have demonstrated that LOH at a specific gene locus could be used as biomarkers of cancer risk or the prediction of clinical outcomes in certain types of tumors. Here we summarized a subset of reported LOH related biomarkers (Table 1). Some have been extensively studied, such as the tumor suppressor TP53 on 17p13, while others are relevant for certain tumor types, such as HLA-A02 in patients with synovial sarcoma.

5.1. Biomarkers on Chromosome 1q Associated with LOH

In primary neuroblastoma, loss of heterozygosity on 1p36 has been reported in 23–35% of patients, and was shown to be significantly associated with prognostic markers of aggressive neuroblastoma when patients are diagnosed. Moreover, loss of 1p36 has been reported to predict both poor event-free and overall survival [61,62] and the genomic status of 1p has been implemented as a risk stratifying marker in the German trial NB2004 [83]. The underlying driver gene mutation is unknown, but may involve one or more neuroblastoma tumor suppressor genes and loss of 1p36 could promote tumor growth and favor the effects from the MYCN oncogene [84,85].

5.2. Biomarkers on Chromosome 3q Associated with LOH

FRA3B locating on chromosome 3p14.2, is one of the most active common fragile sites in the human genome and has been reported to incur deletions or translocation breakpoints in many types of cancers including lung cancers [86]. The FHIT gene is one of the most well-known genes at this common fragile site and has been reported to undergo LOH in >76% of lung cancer [87,88]. Interestingly, several studies have revealed that a significantly higher level of FHIT LOH occurs in lung cancers from smokers compared with that from nonsmokers, indicating that this could be a predictive marker of an early event in the genesis of smoking-related cancers [65,66]. Another example is CACNA2D3 encoding a protein which is an auxiliary member of the alpha-2-delta subunit family of the voltage-dependent calcium channel complex. There are two commonly SNP sites undergoing LOH (rs589281, rs6797113) at 3p21 which have been reported to associate with poor clinical outcome in esophageal cancer [67,68]. The von Hippel-Lindau (VHL) tumor suppressor gene at 3p25 plays a vital role in the regulation of other genes and control of cell division. LOH at two loci (rs1642742 and rs1642743) of VHL gene results in a loss of VHL protein function, has been proposed as a candidate predictive biomarker for clinical outcome in clear-cell renal-cell carcinoma (ccRCC) patients [69].

5.3. Biomarkers on Chromosome 6q Associated with LOH

Loss of heterozygosity of the human leukocyte antigen (HLA) locus locating on Chromosome 6p21 was associated with low peptide diversity in colorectal cancers (CRCs), entailing a poor response to immune checkpoint inhibitors such as inhibitors targeting mitogen-activated protein kinase kinase (MEK) [70]. Similarly, LOH at HLA-A02 has been proposed as a predictive biomarker for synovial sarcoma and is prognostic of poor outcome [71]. Fatty acid-binding protein-7 (FABP7) is involved in the transportation of intracellular long-chain fatty acid [89] and has been shown to regulate several pathways leading to the inhibition of cell proliferation and tissue differentiation [90]. The expression of FABP7 correlated with survival rate in patients with glioblastoma [72] and the frequency of LOH at 6q22.31 was 50% in metastatic melanomas, compared with 0 of 14 in primary melanomas (p = 0.0017) indicating that FABP7 is a potential diagnostic biomarker of early-stage melanoma dissemination in blood [73].

5.4. Biomarkers on Chromosome 16q Associated with LOH

The B-lymphocyte antigen CD19 at 16p11.2 is expressed in all B lineage cells [91]. The chimeric antigen receptor (CAR) T cell therapy CTL019 (Tisagenlecleucel) is designed to use T cells of patients expressing a CD19-specific, 4-1BB/CD3 ζ-signaling CAR [92]. Patients with relapsed or refractory B cell acute lymphoblastic leukemia (B-ALL) receiving CTL019 have achieved 70% to 94% complete remission in several clinical trials. However, more than 30% of tumors that responded to CTL019 therapy eventually recurred [93,94] which was found to be caused by frameshift mutations leading to a truncated protein with a nonfunctional or absent transmembrane domain and consequently to a loss of surface antigen [77]. Notably, there was no CD19 LOH in samples from 7 untreated patients), however, 8/9 relapsed patients had acquired LOH at CD19 [95], indicating that irreversible loss of heterozygosity in CD19 could be used as a biomarker for an outcome prediction after the CAR T cells therapy CTL019 [77].

5.5. Biomarkers on Chromosome 17q Associated with LOH

The tumor suppressor TP53, located on 17p13.1, contains 11 exons of which the genomic integrity of exons 5-8 plays a particularly important role for its activity. These TP53 gene exons are also mutational hotspots in cancer [96,97] and >80% of TP53 mutations occur within this region. So far, more than 27,000 somatic mutation in TP53 have been catalogued in the International agency for research on cancer (IARC) TP53 database [98]. Further, LOH at 17p13.1 is one of the most frequent genetic alterations leading to human cancers [78,79]. Analysis of TP53 mutations from 1,420 tumor samples revealed that loss of one TP53 allele may be a sufficient driver in breast cancers of the luminal B subtype as a higher fraction of wild-type tumors with LOH is noticed [99]. Codon 72 of the TP53 gene contains a well-known polymorphic site with the two alleles arginine and proline (Arg72Pro) [100]. Analysis of TP53 mutations from 204 Danish women and revealed that heterozygous patients losing either the Arg72 or Pro72 variant in the TP53 gene because of LOH had a reduction in disease-free survival compared with patients retaining the polymorphism (Arg72Pro) [101]. Another analysis revealed that patients with homozygous deletion under LOH or both del17p and TP53 mutation had a significantly worse outcome, comparing with patients who had only del17p or TP53 mutation (3-year overall survival 84% versus 29% (p = 0.02) and 3-year progression-free survival 73% versus 29% (p = 0.04), respectively) [102].

5.6. Biomarkers on Other Chromosomes Associated with LOH

Loss of heterozygosity at 9p was observed to be significantly associated with poorer prognosis of glioma patients [74]. Another study reported that LOH at 9p13 was associated with poor survival in squamous cell carcinoma/adenocarcinoma patients who had surgical resection. The nel-like1 (NELL1) gene at 11p15 is reported to frequently undergo loss of heterozygosity in esophageal adenocarcinoma (EAC) [76]. Another study from Bepler et al. reported an association between NELL1 LOH at position of 11p15.5 and poor survival outcome in 180 patients who had lung cancer [103]. A number of studies have revealed that colorectal cancer patients at stage III with LOH at SMAD4 at 18q showed a poorer overall 5-year survival rate comparing with patients harboring non-18q LOH [81]. A meta-analysis of data from 27 studies and 2189 patients [82] extended this finding, showing that LOH on chromosome 18q could assist in predicting the clinical outcome after therapies in stage III colorectal cancer (CRC).

6. Loss of Heterozygosity in Tumors Provides Up-and-Coming Possibilities to Understand the Underlying Mechanisms of Cancer Evolution

As described above loci undergoing loss of heterozygosity could be used as clinical biomarkers or the prediction of clinical outcomes in certain types of tumors and provide novel therapeutic targets for cancer treatment which have been proved applicable in current different studies. Here, last but not the least, studies of loss of heterozygosity in tumors also provides up-and-coming possibilities to understand the underlying mechanism of cancer evolution which surpasses the scheme of targeting point mutations in tumorigenesis and focuses on the contribution of losses and gains of genetic material through whole chromosomes or chromosome arms [1].
A very recent study [104] relating to non-small cell lung cancer (NSCLC) has observed a significant enrichment for whole-genome doubling (WGD) which is coordinated with an extensive loss of heterozygosity, indicating that loss of heterozygosity and whole-genome doubling are common events [105]. Then they pursue to understand the effects of natural selection on LOH and WGD during cancer evolution by comparing the early (pre-WGD) and late (post-WGD) mutations within segments of LOH based on their modified dN/dS ratio. The dN/dS ratio has a long history in the study of selection in species evolution [106,107], by comparing synonymous (silent; dS) and nonsynonymous (amino acid-changing; dN) substitution rates in protein-coding DNA sequences [107]. Based on their modified dN/dS ratio test [41] and a well-accepted criteria: ratio of dN/dS > 1 indicates a positive selection and an enrichment of non-silent mutations in cancer genes [41,108,109], López et al. [104]. focused on identifying new cancer genes based on the dN/dS ratio which plays an important role in cancer evolution. Firstly, using their method for calculating dN/dS ratio and a series of strict filtration, the two well-established tumor suppressor genes TP53 and PTEN were discovered with a dN/dS ratio > 50 in lung squamous cell carcinoma (LUSC), indicating the reliability of their analysis method. ZNF750 (zinc finger protein 750) which is described as a tumor suppressor gene in squamous cell carcinoma [110,111], is as well subject to high positive selection (dN/dS ratio > 50) in regions of LOH in lung squamous cell carcinoma (LUSC) and head and neck squamous cell carcinoma (HNSC) [104]. Notably, they further expanded the analysis to the remaining cancer subtypes in TCGA and 33 potential tumor suppressor genes were identified by limiting the analyses to sections of LOH and 14 additional novel genes (WWC1, FPF12, NT5DC3,NCLN, KRTAP195, GRIK2, GLRA1, FAXDC2, FAM19A3, CRYGC, CLEC4E, BC02, ARPP21 and AC061992.1) were discovered which were assumed tumor suppressor genes but not included in the catalogue of somatic mutations in cancer (COSMIC) [112]. This approach provides a unique angle to discover novel cancer genes which play roles in cancer evolution. Moreover, many following studies could be carried out based on this discovery, such as cancer prediction and targeted cancer therapies.
Recently, Zaccaria et al. [113] introduced another algorithm named copy-number haplotype inference in single cells using evolutionary links (CHISEL) to study cancer genes which are involved in the early stage of tumor evolution. CHISEL algorithm is able to deduce allele- and haplotype-specific copy numbers in single cells [113] and overcome the problems in already existing single-cell barcoding technologies which are usually caused by the extremely low sequencing coverage (<0.05× per cell). Based on CHISEL algorithm, they reanalyzed 10× genomics chromium single-cell DNA sequencing data from two patients at early stage of breast cancer (patient 1: a triple negative ductal carcinoma in situ and patient 2: at stage 1 infiltrating ductal carcinoma) and identified large-scale allele-specific copy-number aberrations (CNAs) which, on average, cover ~25% of the genome in at least 100 cells however were uncharacterized by previous used total copy-number analysis. Further investigation into the ~25% CNAs of the genome revealed that the most common type of allele-specific CNAs identified by CHISEL are copy-neutral LOHs and contains several genes which have been reported in the previous 560 breast cancer database [114], such as ARID1B and ESR1 on chromosome 6q, PTEN on chromosome 10q, BRCA2 and RB1 on chromosome 13, It is worth noting that all the genes undergoing copy-neutral LOHs have been reported in almost all type of tumor cells [115,116,117,118,119,120,121,122,123,124,125,126,127], indicating that these copy-neutral LOHs mutations may take an important role at an early stage during the tumor evolution.
In 2018, Palin et al. [128] have performed a high-resolution allelic imbalance (AI) landscape in 1699 colorectal tumor/normal DNA pairs and among them, 256 were later whole-genome sequenced. The results revealed that three loci commonly undergoing LOH are TP53 (63%, 1072 of 1699 cancers), SMAD4 (61% or 1040 of 1699 cancer) and 8p21.3 (47% or 806 of 1699 cancers) [128]. It is worth noting that 8p21.3 shows no obvious candidate target genes. However, previous studies have revealed that a frequent loss of chromosome 8p is firmly linked to colorectal tumorigenesis [129,130] and 8p loss and MYC gain (also in 8q) significantly co-occurred in the same samples, indicating that a frequent loss of 8p may at least to some extent be an accompanying event which is driven by the strong positive selective of MYC gain.
Moreover, the study from Palin et al. is as well important to understand the mechanism underlying cancer evolution. When discussing tumorigenesis, we always like to refer to the hallmarks of cancer [131], believing that cancer is the process of accumulating mutations of genes. However, one common question based on this theory is still hard to answer: why so few genes are commonly mutated in most cancers [1,132] and one obstacle occurs in cancer therapies which is still hard to solve: commonly mutated genes such as TP53 [1,41] and MYC [52,53] are hard to target. This study of developing genetic aberrations provides a fresh perspective that it is not point mutations which are frequently selected for tumorigenesis. Instead, gross chromosomal changes which could affect the expression levels of cancer genes as well as many other genes through a single event during the process of tumorigenesis. Although these studies mainly focused on colorectal cancer [128,133,134,135], the conclusion encourages researchers to carry out similar studies in other types of cancer. It is believed that studies of loss of heterozygosity in tumors provide up-and-coming possibilities to understand the progress of tumorigenesis which is a totally different but complementary aspect to studies of targeting point mutations.
Moreover, emerging studies have shown that findings on loss of heterozygosity shed a light on understanding the mechanism underlying tumor immune evasion during tumor evolution. An important step in the process of neoantigen presentation and cytolytic T cell response is regulated by class I human leukocyte antigen (HLA), which enables recognition by T cell receptors through expressing intra-cellular peptides on the surface of cell [136,137]. Thus, the downregulation of HLA expression at transcriptional or translational level should be able to reduce the ability of antigen presentation, leading to the consequence of immune evasion. Indeed, the downregulation of HLA presentation has been discovered in several individual studies and has as well been shown to link to poor outcome in clinical therapies across different cancer types [138,139,140,141,142]. The first report trying to explain this phenomenon from the perspective of loci undergoing loss of heterozygosity was published in 2016 [143] and described that after the treatment of seven patients with lung metastases by infusing ~1.11 × 1011 HLA-C*08:02—restricted tumor-infiltrating lymphocytes, six out of seven patients observed an objective regression on evaluation after 9-month therapy. However, one patient had progressed and the sequencing result revealed that this patient has lost the chromosome 6 haplotype encoding the HLA-C*08:02 class I major histocompatibility complex (MHC) molecule, providing a mechanism for tumor immune evasion [143]. A similar result revealing that loss of heterozygosity of the human leukocyte antigen is associated with low peptide diversity and leads to a poor response to immune checkpoint inhibitors has been reported in colorectal cancers [70]. Another study relating to lung cancer accelerated the study about the relationship of loss of heterozygosity at HLA and immune escape in evolution of lung cancer [144]. After analyzing the HLA locus in 288 TRACERx non-small-cell lung cancer (NSCLC) exomes from 96 patients [145,146], researchers observed that LOH occurs in 40% of early-stage non-small-cell lung cancers, indicating that LOH HLA could be an advantage during the evolution of non-small-cell lung cancers. To clarify the timing of HLA LOH in NSCLC evolution, sequencing data from 37 NSCLC primary tumors with matched brain metastases were analyzed [147], revealing that 46% had HLA LOH with a trend toward enrichment of HLA LOH in brain metastases from the same patient, comparing with the primary tumor originally from lung. Thus, LOH HLA more likely appears at a later stage in cancer development and may provide an advantage under natural selection in order to escape immune predation. These findings further illustrated the importance of LOH in cancer evolution and tumor immune escape. As more pre- and post-therapy data emerges from different cohorts, there is great potential to investigate the extent of HLA LOH in different cancer types at different stages of cancer evolution, to improve understanding of the mechanisms underlying tumor immune evasion. Following the idea in this study, the other immune related LOH loci would be discovered in other research.

7. Future Perspective

Loss of heterozygosity (LOH) is a common genetic event in the development of cancer and is known to play an essential role in the somatic loss of wild-type alleles in cancers. Moreover, a vast number of non-driver genes concomitantly undergo LOH, corresponding to more than 20% of the total genome, such as in colorectal carcinomas [8,148]. This creates patient specific differences between tumor and normal cells, which may in turn generate vulnerabilities in cancer cells that could potentially be targeted. Here, we have reviewed and summarized mutations associated with LOH on chromosomes which have been shown to be promising biomarkers of cancer risk or predictive of clinical outcomes in certain types of tumors. Moreover, with the development of genome-wide analysis of clinically relevant data, targeting cancer cell vulnerabilities associated with LOH could provide a novel paradigm for cancer therapies in populations harboring the same LOH allele. As loss of heterozygosity is a common genetic event in cancer genomes, analysis of LOH could in addition provide a therapeutic option for precision medicine. Importantly, LOH specific inhibitory probes or cytotoxic drugs from small molecule screening have been identified, proving the concept that LOH targeting in cancer cells is a fertile avenue for novel anticancer drug discovery and development [26]. To progress this field, additional genomic analysis based on data from large clinical sample cohorts should be performed in order to find exploitable nucleotide level and structural variants at LOH loci in common cancers. Functional studies should be performed to validate druggable non-driver genes which are either essential for cell survival, proliferation or involved in vital metabolic cellular activities. Last but not the least, we have addressed several up-to-date findings relating to the importance of LOH for understanding the mechanisms underlying the progress of cancer evolution [104,128,149]. Methods of analyzing genomic data are continuously improving, and some cancer genes which have not been found based on previous methods have yet been discovered in recent work [104] and some basic questions which have not been fully addressed have currently been partially settled by a high-resolution allelic imbalance landscape [128]. Thus, studies relating to the LOH in cancer cells will help to further understand differences between tumor and normal cells as well the mechanism underlying cancer evolution and tumor immune evasion, opening novel avenues for cancer therapy development. At the same time, there is a highly need to utilize the results/findings from these studies, such as perform necessary investigations to look into the biological functions of the LOH genes among the top list [104,113] or novel strategies to target proteins which are translated from a gene harboring LOH loci [26,142,144,150].

Author Contributions

X.Z. and T.S. wrote the paper. Both authors have read and agreed to the published version of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grants to T.S. from the Swedish Cancer Foundation, the Swedish Research Council, the Swedish Foundation for Strategic Research, The Erling-Persson Family Foundation, and the Swedish Children’s Cancer Foundation.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vogelstein, B.; Papadopoulos, N.; Velculescu, V.E.; Zhou, S.; Diaz, L.A., Jr.; Kinzler, K.W. Cancer genome landscapes. Science 2013, 339, 1546–1558. [Google Scholar] [CrossRef] [PubMed]
  2. Cavenee, W.K.; Dryja, T.P.; Phillips, R.A.; Benedict, W.F.; Godbout, R.; Gallie, B.L.; Murphree, A.L.; Strong, L.C.; White, R.L. Expression of recessive alleles by chromosomal mechanisms in retinoblastoma. Nature 1983, 305, 779–784. [Google Scholar] [CrossRef] [PubMed]
  3. Fitzgibbon, J.; Smith, L.L.; Raghavan, M.; Smith, M.L.; Debernardi, S.; Skoulakis, S.; Lillington, D.; Lister, T.A.; Young, B.D. Association between acquired uniparental disomy and homozygous gene mutation in acute myeloid leukemias. Cancer Res. 2005, 65, 9152–9154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Ryland, G.L.; Doyle, M.A.; Goode, D.; Boyle, S.E.; Choong, D.Y.; Rowley, S.M.; Li, J.; Australian Ovarian Cancer Study, G.; Bowtell, D.D.; Tothill, R.W.; et al. Loss of heterozygosity: What is it good for? BMC Med. Genom. 2015, 8, 45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Tapial, S.; Garcia, J.L.; Corchete, L.; Holowatyj, A.N.; Perez, J.; Rueda, D.; Urioste, M.; Gonzalez-Sarmiento, R.; Perea, J. Copy neutral loss of heterozygosity (cnLOH) patterns in synchronous colorectal cancer. Eur. J. Hum. Genet. 2020. [Google Scholar] [CrossRef] [PubMed]
  6. Lo, H.S.; Wang, Z.; Hu, Y.; Yang, H.H.; Gere, S.; Buetow, K.H.; Lee, M.P. Allelic variation in gene expression is common in the human genome. Genom. Res. 2003, 13, 1855–1862. [Google Scholar] [CrossRef]
  7. O’Keefe, C.; McDevitt, M.A.; Maciejewski, J.P. Copy neutral loss of heterozygosity: A novel chromosomal lesion in myeloid malignancies. Blood 2010, 115, 2731–2739. [Google Scholar] [CrossRef] [Green Version]
  8. Lengauer, C.; Kinzler, K.W.; Vogelstein, B. Genetic instabilities in human cancers. Nature 1998, 396, 643–649. [Google Scholar] [CrossRef]
  9. Murphree, A.L.; Benedict, W.F. Retinoblastoma: Clues to human oncogenesis. Science 1984, 223, 1028–1033. [Google Scholar] [CrossRef]
  10. Jiang, G.; Zhang, S.; Yazdanparast, A.; Li, M.; Pawar, A.V.; Liu, Y.; Inavolu, S.M.; Cheng, L. Comprehensive comparison of molecular portraits between cell lines and tumors in breast cancer. BMC Genom. 2016, 17 (Suppl. 7), 525. [Google Scholar] [CrossRef] [Green Version]
  11. Takai, E.; Nakamura, H.; Chiku, S.; Kubo, E.; Ohmoto, A.; Totoki, Y.; Shibata, T.; Higuchi, R.; Yamamoto, M.; Furuse, J.; et al. Whole-exome Sequencing Reveals New Potential Susceptibility Genes for Japanese Familial Pancreatic Cancer. Ann. Surg. 2020. [Google Scholar] [CrossRef] [PubMed]
  12. Kaur, R.P.; Vasudeva, K.; Kumar, R.; Munshi, A. Role of p53 Gene in Breast Cancer: Focus on Mutation Spectrum and Therapeutic Strategies. Curr. Pharm. Des. 2018, 24, 3566–3575. [Google Scholar] [CrossRef]
  13. Lea, I.A.; Jackson, M.A.; Li, X.; Bailey, S.; Peddada, S.D.; Dunnick, J.K. Genetic pathways and mutation profiles of human cancers: Site- and exposure-specific patterns. Carcinogenesis 2007, 28, 1851–1858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Chauffaille, M.; Zalcberg, I.; Barreto, W.G.; Bendit, I. Detection of somatic TP53 mutations and 17p deletions in patients with chronic lymphocytic leukemia: A review of the current methods. Hematol. Transfus. Cell Ther. 2020, 42, 261–268. [Google Scholar] [CrossRef] [PubMed]
  15. Lebok, P.; Kopperschmidt, V.; Kluth, M.; Hube-Magg, C.; Ozden, C.; Taskin, B.; Hussein, K.; Mittenzwei, A.; Lebeau, A.; Witzel, I.; et al. Partial PTEN deletion is linked to poor prognosis in breast cancer. BMC Cancer 2015, 15, 963. [Google Scholar] [CrossRef] [Green Version]
  16. Neff, R.T.; Senter, L.; Salani, R. BRCA mutation in ovarian cancer: Testing, implications and treatment considerations. Ther. Adv. Med. Oncol 2017, 9, 519–531. [Google Scholar] [CrossRef] [Green Version]
  17. BRCA Gene Mutations: Cancer Risk and Genetic Testing. Available online: https://www.cancer.gov/about-cancer/causes-prevention/genetics/brca-fact-sheet (accessed on 19 November 2020).
  18. O’Brien, C. New tumor suppressor found in pancreatic cancer. Science 1996, 271, 294. [Google Scholar] [CrossRef]
  19. Hahn, S.A.; Schutte, M.; Hoque, A.T.; Moskaluk, C.A.; da Costa, L.T.; Rozenblum, E.; Weinstein, C.L.; Fischer, A.; Yeo, C.J.; Hruban, R.H.; et al. DPC4, a candidate tumor suppressor gene at human chromosome 18q21.1. Science 1996, 271, 350–353. [Google Scholar] [CrossRef]
  20. Lin, L.H.; Chang, K.W.; Cheng, H.W.; Liu, C.J. SMAD4 Somatic Mutations in Head and Neck Carcinoma Are Associated with Tumor Progression. Front. Oncol 2019, 9, 1379. [Google Scholar] [CrossRef] [Green Version]
  21. Hadžija, M.P.; Kapitanovic, S.; Radosevic, S.; Cacev, T.; Mirt, M.; Kovacevic, D.; Cacev, T.; Hadzija, M.; Spaventi, R.; Pavelic, K. Loss of heterozygosity of DPC4 tumor suppressor gene in human sporadic colon cancer. J. Mol. Med. (Berl.) 2001, 79, 128–132. [Google Scholar] [CrossRef]
  22. Boulay, J.L.; Stiefel, U.; Taylor, E.; Dolder, B.; Merlo, A.; Hirth, F. Loss of heterozygosity of TRIM3 in malignant gliomas. BMC Cancer 2009, 9, 71. [Google Scholar] [CrossRef] [Green Version]
  23. Boultwood, J. CUX1 in leukemia: Dosage matters. Blood 2013, 121, 869–871. [Google Scholar] [CrossRef] [PubMed]
  24. Mertens, F.; Heim, S.; Mandahl, N.; Johansson, B.; Mertens, O.; Persson, B.; Salemark, L.; Wennerberg, J.; Jonsson, N.; Mitelman, F. Cytogenetic analysis of 33 basal cell carcinomas. Cancer Res. 1991, 51, 954–957. [Google Scholar] [PubMed]
  25. Nichols, C.A.; Gibson, W.J.; Brown, M.S.; Kosmicki, J.A.; Busanovich, J.P.; Wei, H.; Urbanski, L.M.; Curimjee, N.; Berger, A.C.; Gao, G.F.; et al. Loss of heterozygosity of essential genes represents a widespread class of potential cancer vulnerabilities. Nat. Commun. 2020, 11, 2517. [Google Scholar] [CrossRef] [PubMed]
  26. Rendo, V.; Stoimenov, I.; Mateus, A.; Sjoberg, E.; Svensson, R.; Gustavsson, A.L.; Johansson, L.; Ng, A.; O’Brien, C.; Giannakis, M.; et al. Exploiting loss of heterozygosity for allele-selective colorectal cancer chemotherapy. Nat. Commun. 2020, 11, 1308. [Google Scholar] [CrossRef] [PubMed]
  27. Nguyen, L.; Martens, J.; Van Hoeck, A.; Cuppen, E. Pan-cancer landscape of homologous recombination deficiency. Nat. Commun. 2020, 11, 5584. [Google Scholar] [CrossRef]
  28. Vosoughi, A.; Zhang, T.; Shohdy, K.S.; Vlachostergios, P.J.; Wilkes, D.C.; Bhinder, B.; Tagawa, S.T.; Nanus, D.M.; Molina, A.M.; Beltran, H.; et al. Common germline-somatic variant interactions in advanced urothelial cancer. Nat. Commun. 2020, 11, 6195. [Google Scholar] [CrossRef] [PubMed]
  29. Pedersen, B.S.; De, S. Loss of heterozygosity preferentially occurs in early replicating regions in cancer genomes. Nucleic Acids Res. 2013, 41, 7615–7624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Stafford, W.C.; Peng, X.; Olofsson, M.H.; Zhang, X.; Luci, D.K.; Lu, L.; Cheng, Q.; Tresaugues, L.; Dexheimer, T.S.; Coussens, N.P.; et al. Irreversible inhibition of cytosolic thioredoxin reductase 1 as a mechanistic basis for anticancer therapy. Sci. Transl. Med. 2018, 10. [Google Scholar] [CrossRef] [Green Version]
  31. Tantin, D. Oct transcription factors in development and stem cells: Insights and mechanisms. Development 2013, 140, 2857–2866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Shakya, A.; Cooksey, R.; Cox, J.E.; Wang, V.; McClain, D.A.; Tantin, D. Oct1 loss of function induces a coordinate metabolic shift that opposes tumorigenicity. Nat. Cell Biol. 2009, 11, 320–327. [Google Scholar] [CrossRef] [PubMed]
  33. Kang, J.; Gemberling, M.; Nakamura, M.; Whitby, F.G.; Handa, H.; Fairbrother, W.G.; Tantin, D. A general mechanism for transcription regulation by Oct1 and Oct4 in response to genotoxic and oxidative stress. Genes Dev. 2009, 23, 208–222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Moscow, J.A.; Schmidt, L.; Ingram, D.T.; Gnarra, J.; Johnson, B.; Cowan, K.H. Loss of heterozygosity of the human cytosolic glutathione peroxidase I gene in lung cancer. Carcinogenesis 1994, 15, 2769–2773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Hu, Y.; Benya, R.V.; Carroll, R.E.; Diamond, A.M. Allelic loss of the gene for the GPX1 selenium-containing protein is a common event in cancer. J. Nutr. 2005, 135, 3021S–3024S. [Google Scholar] [CrossRef] [PubMed]
  36. Lenders, J.W.; Duh, Q.Y.; Eisenhofer, G.; Gimenez-Roqueplo, A.P.; Grebe, S.K.; Murad, M.H.; Naruse, M.; Pacak, K.; Young, W.F., Jr.; Endocrine, S. Pheochromocytoma and paraganglioma: An endocrine. society clinical practice guideline. J. Clin. Endocrinol. Metab. 2014, 99, 1915–1942. [Google Scholar] [CrossRef]
  37. Jochmanova, I.; Yang, C.; Zhuang, Z.; Pacak, K. Hypoxia-inducible factor signaling in pheochromocytoma: Turning the rudder in the right direction. J. Natl. Cancer Inst. 2013, 105, 1270–1283. [Google Scholar] [CrossRef]
  38. Zhuang, Z.; Yang, C.; Lorenzo, F.; Merino, M.; Fojo, T.; Kebebew, E.; Popovic, V.; Stratakis, C.A.; Prchal, J.T.; Pacak, K. Somatic HIF2A gain-of-function mutations in paraganglioma with polycythemia. N. Engl. J. Med. 2012, 367, 922–930. [Google Scholar] [CrossRef] [Green Version]
  39. Sawyers, C. Targeted cancer therapy. Nature 2004, 432, 294–297. [Google Scholar] [CrossRef]
  40. Yang, B. Targeted therapy for cancer. Discov. Med. 2004, 4, 18–21. [Google Scholar]
  41. Martincorena, I.; Raine, K.M.; Gerstung, M.; Dawson, K.J.; Haase, K.; Van Loo, P.; Davies, H.; Stratton, M.R.; Campbell, P.J. Universal Patterns of Selection in Cancer and Somatic Tissues. Cell 2017, 171, 1029–1041. [Google Scholar] [CrossRef]
  42. Vousden, K.H.; Ryan, K.M. p53 and metabolism. Nat. Rev. Cancer 2009, 9, 691–700. [Google Scholar] [CrossRef] [PubMed]
  43. Baugh, E.H.; Ke, H.; Levine, A.J.; Bonneau, R.A.; Chan, C.S. Why are there hotspot mutations in the TP53 gene in human cancers? Cell Death Differ. 2018, 25, 154–160. [Google Scholar] [CrossRef] [PubMed]
  44. Litviakov, N.V.; Denisov, E.V.; Takhauov, R.M.; Karpov, A.B.; Skobel’skaja, E.V.; Vasil’eva, E.O.; Goncharik, O.O.; Ageeva, A.M.; Mamonova, N.V.; Mezheritskiy, S.A.; et al. Association between TP53 gene ARG72PRO polymorphism and chromosome aberrations in human cancers. Mol. Carcinog. 2010, 49, 521–524. [Google Scholar] [CrossRef] [PubMed]
  45. Beroukhim, R.; Mermel, C.H.; Porter, D.; Wei, G.; Raychaudhuri, S.; Donovan, J.; Barretina, J.; Boehm, J.S.; Dobson, J.; Urashima, M.; et al. The landscape of somatic copy-number alteration across human cancers. Nature 2010, 463, 899–905. [Google Scholar] [CrossRef]
  46. Ciriello, G.; Miller, M.L.; Aksoy, B.A.; Senbabaoglu, Y.; Schultz, N.; Sander, C. Emerging landscape of oncogenic signatures across human cancers. Nat. Genet. 2013, 45, 1127–1133. [Google Scholar] [CrossRef] [Green Version]
  47. Carter, D.R.; Murray, J.; Cheung, B.B.; Gamble, L.; Koach, J.; Tsang, J.; Sutton, S.; Kalla, H.; Syed, S.; Gifford, A.J.; et al. Therapeutic targeting of the MYC signal by inhibition of histone chaperone FACT in neuroblastoma. Sci. Transl. Med. 2015, 7, 312ra176. [Google Scholar] [CrossRef] [Green Version]
  48. Rickman, D.S.; Schulte, J.H.; Eilers, M. The Expanding World of N-MYC-Driven Tumors. Cancer Discov. 2018, 8, 150–163. [Google Scholar] [CrossRef] [Green Version]
  49. Fan-Minogue, H.; Bodapati, S.; Solow-Cordero, D.; Fan, A.; Paulmurugan, R.; Massoud, T.F.; Felsher, D.W.; Gambhir, S.S. A c-Myc activation sensor-based high-throughput drug screening identifies an antineoplastic effect of nitazoxanide. Mol. Cancer Ther. 2013, 12, 1896–1905. [Google Scholar] [CrossRef] [Green Version]
  50. Bandopadhayay, P.; Bergthold, G.; Nguyen, B.; Schubert, S.; Gholamin, S.; Tang, Y.; Bolin, S.; Schumacher, S.E.; Zeid, R.; Masoud, S.; et al. BET bromodomain inhibition of MYC-amplified medulloblastoma. Clin. Cancer Res. 2014, 20, 912–925. [Google Scholar] [CrossRef] [Green Version]
  51. Demma, M.J.; Mapelli, C.; Sun, A.; Bodea, S.; Ruprecht, B.; Javaid, S.; Wiswell, D.; Muise, E.; Chen, S.; Zelina, J.; et al. Omomyc Reveals New Mechanisms To Inhibit the MYC Oncogene. Mol. Cell Biol. 2019, 39, e00248-19. [Google Scholar] [CrossRef] [Green Version]
  52. Villanueva, M.T. Long path to MYC inhibition approaches clinical trials. Nat. Rev. Cancer 2019, 19, 252. [Google Scholar] [CrossRef] [PubMed]
  53. Zinzalla, G. Targeting MYC: Is it getting any easier? Future Med. Chem. 2016, 8, 1899–1902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. He, Z.; Henricksen, L.A.; Wold, M.S.; Ingles, C.J. RPA involvement in the damage-recognition and incision steps of nucleotide excision repair. Nature 1995, 374, 566–569. [Google Scholar] [CrossRef]
  55. Brill, S.J.; Stillman, B. Replication factor-A from Saccharomyces cerevisiae is encoded by three essential genes coordinately expressed at S phase. Genes Dev. 1991, 5, 1589–1600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Basilion, J.P.; Schievella, A.R.; Burns, E.; Rioux, P.; Olson, J.C.; Monia, B.P.; Lemonidis, K.M.; Stanton, V.P., Jr.; Housman, D.E. Selective killing of cancer cells based on loss of heterozygosity and normal variation in the human genome: A new paradigm for anticancer drug therapy. Mol. Pharmacol. 1999, 56, 359–369. [Google Scholar] [CrossRef] [PubMed]
  57. Courtney, D.G.; Moore, J.E.; Atkinson, S.D.; Maurizi, E.; Allen, E.H.A.; Pedrioli, D.M.L.; McLean, W.H.I.; Nesbit, M.A.; Moore, C.B.T. CRISPR/Cas9 DNA cleavage at SNP-derived PAM enables both in vitro and in vivo KRT12 mutation-specific targeting. Gene Ther. 2015, 23, 108–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Kilchert, C.; Wittmann, S.; Vasiljeva, L. The regulation and functions of the nuclear RNA exosome complex. Nat. Rev. Mol. Cell Biol. 2016, 17, 227–239. [Google Scholar] [CrossRef]
  59. Halling-Brown, M.D.; Bulusu, K.C.; Patel, M.; Tym, J.E.; Al-Lazikani, B. canSAR: An integrated cancer public translational research and drug discovery resource. Nucleic Acids Res. 2012, 40, D947–D956. [Google Scholar] [CrossRef] [Green Version]
  60. Bulusu, K.C.; Tym, J.E.; Coker, E.A.; Schierz, A.C.; Al-Lazikani, B. canSAR: Updated cancer research and drug discovery knowledgebase. Nucleic Acids Res. 2014, 42, D1040–D1047. [Google Scholar] [CrossRef] [Green Version]
  61. Komotar, R.J.; Otten, M.L.; Starke, R.M.; Anderson, R.C. Chromosome 1p and 11q deletions and outcome in neuroblastoma-a critical review. Clin. Med. Oncol. 2008, 2, 419–420. [Google Scholar] [CrossRef]
  62. Attiyeh, E.F.; London, W.B.; Mosse, Y.P.; Wang, Q.; Winter, C.; Khazi, D.; McGrady, P.W.; Seeger, R.C.; Look, A.T.; Shimada, H.; et al. Chromosome 1p and 11q deletions and outcome in neuroblastoma. N. Engl. J. Med. 2005, 353, 2243–2253. [Google Scholar] [CrossRef]
  63. Caron, H.; van Sluis, P.; de Kraker, J.; Bokkerink, J.; Egeler, M.; Laureys, G.; Slater, R.; Westerveld, A.; Voute, P.A.; Versteeg, R. Allelic loss of chromosome 1p as a predictor of unfavorable outcome in patients with neuroblastoma. N. Engl. J. Med. 1996, 334, 225–230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Spitz, R.; Hero, B.; Westermann, F.; Ernestus, K.; Schwab, M.; Berthold, F. Fluorescence in situ hybridization analyses of chromosome band 1p36 in neuroblastoma detect two classes of alterations. Genes Chromosomes Cancer 2002, 34, 299–305. [Google Scholar] [CrossRef] [PubMed]
  65. Stein, C.K.; Glover, T.W.; Palmer, J.L.; Glisson, B.S. Direct correlation between FRA3B expression and cigarette smoking. Genes Chromosomes Cancer 2002, 34, 333–340. [Google Scholar] [CrossRef]
  66. Pylkkanen, L.; Wolff, H.; Stjernvall, T.; Tuominen, P.; Sioris, T.; Karjalainen, A.; Anttila, S.; Husgafvel-Pursiainen, K. Reduced Fhit protein expression and loss of heterozygosity at FHIT gene in tumours from smoking and asbestos-exposed lung cancer patients. Int. J. Oncol. 2002, 20, 285–290. [Google Scholar] [CrossRef] [PubMed]
  67. Wong, A.M.; Kong, K.L.; Chen, L.; Liu, M.; Wong, A.M.; Zhu, C.; Tsang, J.W.; Guan, X.Y. Characterization of CACNA2D3 as a putative tumor suppressor gene in the development and progression of nasopharyngeal carcinoma. Int. J. Cancer 2013, 133, 2284–2295. [Google Scholar] [CrossRef] [PubMed]
  68. Qin, Y.R.; Fu, L.; Sham, P.C.; Kwong, D.L.; Zhu, C.L.; Chu, K.K.; Li, Y.; Guan, X.Y. Single-nucleotide polymorphism-mass array reveals commonly deleted regions at 3p22 and 3p14.2 associate with poor clinical outcome in esophageal squamous cell carcinoma. Int. J. Cancer 2008, 123, 826–830. [Google Scholar] [CrossRef] [PubMed]
  69. Verbiest, A.; Lambrechts, D.; Van Brussel, T.; Couchy, G.; Wozniak, A.; Mejean, A.; Lerut, E.; Oudard, S.; Verkarre, V.; Job, S.; et al. Polymorphisms in the Von Hippel-Lindau Gene Are Associated with Overall Survival in Metastatic Clear-Cell Renal-Cell Carcinoma Patients Treated with VEGFR Tyrosine Kinase Inhibitors. Clin. Genitourin Cancer 2018, 16, 266–273. [Google Scholar] [CrossRef]
  70. Newey, A.; Griffiths, B.; Michaux, J.; Pak, H.S.; Stevenson, B.J.; Woolston, A.; Semiannikova, M.; Spain, G.; Barber, L.J.; Matthews, N.; et al. Immunopeptidomics of colorectal cancer organoids reveals a sparse HLA class I neoantigen landscape and no increase in neoantigens with interferon or MEK-inhibitor treatment. J. Immunother. Cancer 2019, 7, 309. [Google Scholar] [CrossRef]
  71. Rosenbaum, E.; Seier, K.; Bandlamudi, C.; Dickson, M.; Gounder, M.; Keohan, M.L.; Chi, P.; Kelly, C.; Movva, S.; Nacev, B.; et al. HLA Genotyping in Synovial Sarcoma: Identifying HLA-A*02 and Its Association with Clinical Outcome. Clin. Cancer Res. 2020, 26, 5448–5455. [Google Scholar] [CrossRef]
  72. Liang, Y.; Diehn, M.; Watson, N.; Bollen, A.W.; Aldape, K.D.; Nicholas, M.K.; Lamborn, K.R.; Berger, M.S.; Botstein, D.; Brown, P.O.; et al. Gene expression profiling reveals molecularly and clinically distinct subtypes of glioblastoma multiforme. Proc. Natl. Acad. Sci. USA 2005, 102, 5814–5819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Goto, Y.; Koyanagi, K.; Narita, N.; Kawakami, Y.; Takata, M.; Uchiyama, A.; Nguyen, L.; Nguyen, T.; Ye, X.; Morton, D.L.; et al. Aberrant fatty acid-binding protein-7 gene expression in cutaneous malignant melanoma. J. Investig. Dermatol. 2010, 130, 221–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Huang, T.; Li, S.; Yang, Z.; Liu, J.; Han, Y. Loss of Heterozygosity of 9p Is Associated with Poorer Survival in Patients with Gliomas. Mol. Neurobiol. 2016, 53, 6407–6412. [Google Scholar] [CrossRef]
  75. Garcia, J.M.; Silva, J.M.; Dominguez, G.; Gonzalez, R.; Navarro, A.; Carretero, L.; Provencio, M.; Espana, P.; Bonilla, F. Allelic loss of the PTEN region (10q23) in breast carcinomas of poor pathophenotype. Breast Cancer Res. Treat. 1999, 57, 237–243. [Google Scholar] [CrossRef]
  76. Dolan, K.; Garde, J.; Gosney, J.; Sissons, M.; Wright, T.; Kingsnorth, A.N.; Walker, S.J.; Sutton, R.; Meltzer, S.J.; Field, J.K. Allelotype analysis of oesophageal adenocarcinoma: Loss of heterozygosity occurs at multiple sites. Br. J. Cancer 1998, 78, 950–957. [Google Scholar] [CrossRef] [Green Version]
  77. Orlando, E.J.; Han, X.; Tribouley, C.; Wood, P.A.; Leary, R.J.; Riester, M.; Levine, J.E.; Qayed, M.; Grupp, S.A.; Boyer, M.; et al. Genetic mechanisms of target antigen loss in CAR19 therapy of acute lymphoblastic leukemia. Nat. Med. 2018, 24, 1504–1506. [Google Scholar] [CrossRef] [PubMed]
  78. Vousden, K.H.; Lane, D.P. p53 in health and disease. Nat. Rev. Mol. Cell Biol. 2007, 8, 275–283. [Google Scholar] [CrossRef]
  79. Aylon, Y.; Oren, M. New plays in the p53 theater. Curr. Opin. Genet. Dev. 2011, 21, 86–92. [Google Scholar] [CrossRef]
  80. Merajver, S.D.; Frank, T.S.; Xu, J.; Pham, T.M.; Calzone, K.A.; Bennett-Baker, P.; Chamberlain, J.; Boyd, J.; Garber, J.E.; Collins, F.S.; et al. Germline BRCA1 mutations and loss of the wild-type allele in tumors from families with early onset breast and ovarian cancer. Clin. Cancer Res. 1995, 1, 539–544. [Google Scholar]
  81. Sarli, L.; Bottarelli, L.; Bader, G.; Iusco, D.; Pizzi, S.; Costi, R.; D’Adda, T.; Bertolani, M.; Roncoroni, L.; Bordi, C. Association between recurrence of sporadic colorectal cancer, high level of microsatellite instability, and loss of heterozygosity at chromosome 18q. Dis. Colon Rectum 2004, 47, 1467–1482. [Google Scholar] [CrossRef]
  82. Popat, S.; Houlston, R.S. A systematic review and meta-analysis of the relationship between chromosome 18q genotype, DCC status and colorectal cancer prognosis. Eur. J. Cancer 2005, 41, 2060–2070. [Google Scholar] [CrossRef] [PubMed]
  83. Fischer, M.; Spitz, R.; Oberthur, A.; Westermann, F.; Berthold, F. Risk estimation of neuroblastoma patients using molecular markers. Klin. Padiatr. 2008, 220, 137–146. [Google Scholar] [CrossRef]
  84. Fischer, M.; Oberthuer, A.; Brors, B.; Kahlert, Y.; Skowron, M.; Voth, H.; Warnat, P.; Ernestus, K.; Hero, B.; Berthold, F. Differential expression of neuronal genes defines subtypes of disseminated neuroblastoma with favorable and unfavorable outcome. Clin. Cancer Res. 2006, 12, 5118–5128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Wang, Q.; Diskin, S.; Rappaport, E.; Attiyeh, E.; Mosse, Y.; Shue, D.; Seiser, E.; Jagannathan, J.; Shusterman, S.; Bansal, M.; et al. Integrative genomics identifies distinct molecular classes of neuroblastoma and shows that multiple genes are targeted by regional alterations in DNA copy number. Cancer Res. 2006, 66, 6050–6062. [Google Scholar] [CrossRef] [Green Version]
  86. Hibi, K.; Takahashi, T.; Yamakawa, K.; Ueda, R.; Sekido, Y.; Ariyoshi, Y.; Suyama, M.; Takagi, H.; Nakamura, Y.; Takahashi, T. Three distinct regions involved in 3p deletion in human lung cancer. Oncogene 1992, 7, 445–449. [Google Scholar] [PubMed]
  87. Lee, Y.C.; Wu, C.T.; Shih, J.Y.; Jou, Y.S.; Chang, Y.L. Frequent allelic deletion at the FHIT locus associated with p53 overexpression in squamous cell carcinoma subtype of Taiwanese non-small-cell lung cancers. Br. J. Cancer 2004, 90, 2378–2383. [Google Scholar] [CrossRef] [Green Version]
  88. Sozzi, G.; Veronese, M.L.; Negrini, M.; Baffa, R.; Cotticelli, M.G.; Inoue, H.; Tornielli, S.; Pilotti, S.; De Gregorio, L.; Pastorino, U.; et al. The FHIT gene 3p14.2 is abnormal in lung cancer. Cell 1996, 85, 17–26. [Google Scholar] [CrossRef] [Green Version]
  89. Kurtz, A.; Zimmer, A.; Schnutgen, F.; Bruning, G.; Spener, F.; Muller, T. The expression pattern of a novel gene encoding brain-fatty acid binding protein correlates with neuronal and glial cell development. Development 1994, 120, 2637–2649. [Google Scholar]
  90. Anthony, T.E.; Mason, H.A.; Gridley, T.; Fishell, G.; Heintz, N. Brain lipid-binding protein is a direct target of Notch signaling in radial glial cells. Genes Dev. 2005, 19, 1028–1033. [Google Scholar] [CrossRef] [Green Version]
  91. Scheuermann, R.H.; Racila, E. CD19 antigen in leukemia and lymphoma diagnosis and immunotherapy. Leuk Lymphoma 1995, 18, 385–397. [Google Scholar] [CrossRef]
  92. Maude, S.L.; Frey, N.; Shaw, P.A.; Aplenc, R.; Barrett, D.M.; Bunin, N.J.; Chew, A.; Gonzalez, V.E.; Zheng, Z.; Lacey, S.F.; et al. Chimeric antigen receptor T cells for sustained remissions in leukemia. N. Engl. J. Med. 2014, 371, 1507–1517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Maude, S.L.; Laetsch, T.W.; Buechner, J.; Rives, S.; Boyer, M.; Bittencourt, H.; Bader, P.; Verneris, M.R.; Stefanski, H.E.; Myers, G.D.; et al. Tisagenlecleucel in Children and Young Adults with B-Cell Lymphoblastic Leukemia. N. Engl. J. Med. 2018, 378, 439–448. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, Z.; Wu, Z.; Liu, Y.; Han, W. New development in CAR-T cell therapy. J. Hematol. Oncol. 2017, 10, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Riester, M.; Singh, A.P.; Brannon, A.R.; Yu, K.; Campbell, C.D.; Chiang, D.Y.; Morrissey, M.P. PureCN: Copy number calling and SNV classification using targeted short read sequencing. Source Code Biol. Med. 2016, 11, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Benchimol, S.; Lamb, P.; Crawford, L.V.; Sheer, D.; Shows, T.B.; Bruns, G.A.; Peacock, J. Transformation associated p53 protein is encoded by a gene on human chromosome 17. Somat. Cell Mol. Genet. 1985, 11, 505–510. [Google Scholar] [CrossRef]
  97. El-Deiry, W.S.; Kern, S.E.; Pietenpol, J.A.; Kinzler, K.W.; Vogelstein, B. Definition of a consensus binding site for p53. Nat. Genet. 1992, 1, 45–49. [Google Scholar] [CrossRef]
  98. Bouaoun, L.; Sonkin, D.; Ardin, M.; Hollstein, M.; Byrnes, G.; Zavadil, J.; Olivier, M. TP53 Variations in Human Cancers: New Lessons from the IARC TP53 Database and Genomics Data. Hum. Mutat. 2016, 37, 865–876. [Google Scholar] [CrossRef]
  99. Silwal-Pandit, L.; Vollan, H.K.; Chin, S.F.; Rueda, O.M.; McKinney, S.; Osako, T.; Quigley, D.A.; Kristensen, V.N.; Aparicio, S.; Borresen-Dale, A.L.; et al. TP53 mutation spectrum in breast cancer is subtype specific and has distinct prognostic relevance. Clin. Cancer Res. 2014, 20, 3569–3580. [Google Scholar] [CrossRef] [Green Version]
  100. Cho, Y.; Gorina, S.; Jeffrey, P.D.; Pavletich, N.P. Crystal structure of a p53 tumor suppressor-DNA complex: Understanding tumorigenic mutations. Science 1994, 265, 346–355. [Google Scholar] [CrossRef]
  101. Kyndi, M.; Alsner, J.; Hansen, L.L.; Sorensen, F.B.; Overgaard, J. LOH rather than genotypes of TP53 codon 72 is associated with disease-free survival in primary breast cancer. Acta Oncol. 2006, 45, 602–609. [Google Scholar] [CrossRef]
  102. Thanendrarajan, S.; Tian, E.; Qu, P.; Mathur, P.; Schinke, C.; van Rhee, F.; Zangari, M.; Rasche, L.; Weinhold, N.; Alapat, D.; et al. The level of deletion 17p and bi-allelic inactivation of TP53 has a significant impact on clinical outcome in multiple myeloma. Haematologica 2017, 102, e364–e367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Bepler, G.; Gautam, A.; McIntyre, L.M.; Beck, A.F.; Chervinsky, D.S.; Kim, Y.C.; Pitterle, D.M.; Hyland, A. Prognostic significance of molecular genetic aberrations on chromosome segment 11p15.5 in non-small-cell lung cancer. J. Clin. Oncol. 2002, 20, 1353–1360. [Google Scholar] [CrossRef] [PubMed]
  104. Lopez, S.; Lim, E.L.; Horswell, S.; Haase, K.; Huebner, A.; Dietzen, M.; Mourikis, T.P.; Watkins, T.B.K.; Rowan, A.; Dewhurst, S.M.; et al. Interplay between whole-genome doubling and the accumulation of deleterious alterations in cancer evolution. Nat. Genet. 2020, 52, 283–293. [Google Scholar] [CrossRef] [PubMed]
  105. Bielski, C.M.; Zehir, A.; Penson, A.V.; Donoghue, M.T.A.; Chatila, W.; Armenia, J.; Chang, M.T.; Schram, A.M.; Jonsson, P.; Bandlamudi, C.; et al. Genome doubling shapes the evolution and prognosis of advanced cancers. Nat. Genet. 2018, 50, 1189–1195. [Google Scholar] [CrossRef] [PubMed]
  106. Nei, M.; Gojobori, T. Simple methods for estimating the numbers of synonymous and nonsynonymous nucleotide substitutions. Mol. Biol. Evol. 1986, 3, 418–426. [Google Scholar] [CrossRef]
  107. Yang, Z.; Bielawski, J.P. Statistical methods for detecting molecular adaptation. Trends Ecol. Evol. 2000, 15, 496–503. [Google Scholar] [CrossRef]
  108. Tamborero, D.; Gonzalez-Perez, A.; Lopez-Bigas, N. OncodriveCLUST: Exploiting the positional clustering of somatic mu-tations to identify cancer genes. Bioinformatics 2013, 29, 2238–2244. [Google Scholar] [CrossRef]
  109. Zapata, L.; Pich, O.; Serrano, L.; Kondrashov, F.A.; Ossowski, S.; Schaefer, M.H. Negative selection in tumor genome evo-lution acts on essential cellular functions and the immunopeptidome. Genome Biol. 2018, 19, 67. [Google Scholar] [CrossRef]
  110. Nambara, S.; Masuda, T.; Tobo, T.; Kidogami, S.; Komatsu, H.; Sugimachi, K.; Saeki, H.; Oki, E.; Maehara, Y.; Mimori, K. Clinical significance of ZNF750 gene expression, a novel tumor suppressor gene, in esophageal squamous cell carcinoma. Oncol. Lett. 2017, 14, 1795–1801. [Google Scholar] [CrossRef] [Green Version]
  111. Hazawa, M.; Lin, D.C.; Handral, H.; Xu, L.; Chen, Y.; Jiang, Y.Y.; Mayakonda, A.; Ding, L.W.; Meng, X.; Sharma, A.; et al. ZNF750 is a lineage-specific tumour suppressor in squamous cell carcinoma. Oncogene 2017, 36, 2243–2254. [Google Scholar] [CrossRef] [Green Version]
  112. Sondka, Z.; Bamford, S.; Cole, C.G.; Ward, S.A.; Dunham, I.; Forbes, S.A. The COSMIC Cancer Gene Census: Describing genetic dysfunction across all human cancers. Nat. Rev. Cancer 2018, 18, 696–705. [Google Scholar] [CrossRef] [PubMed]
  113. Zaccaria, S.; Raphael, B.J. Characterizing allele- and haplotype-specific copy numbers in single cells with CHISEL. Nat. Biotechnol. 2020. [Google Scholar] [CrossRef] [PubMed]
  114. Nik-Zainal, S.; Davies, H.; Staaf, J.; Ramakrishna, M.; Glodzik, D.; Zou, X.; Martincorena, I.; Alexandrov, L.B.; Martin, S.; Wedge, D.C.; et al. Landscape of somatic mutations in 560 breast cancer whole-genome sequences. Nature 2016, 534, 47–54. [Google Scholar] [CrossRef]
  115. Aso, T.; Uozaki, H.; Morita, S.; Kumagai, A.; Watanabe, M. Loss of ARID1A, ARID1B, and ARID2 Expression during Progression of Gastric Cancer. Anticancer Res. 2015, 35, 6819–6827. [Google Scholar]
  116. Sim, J.C.; White, S.M.; Lockhart, P.J. ARID1B-mediated disorders: Mutations and possible mechanisms. Intractable Rare Dis. Res. 2015, 4, 17–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Zhu, W.; Xu, B. Overcoming resistance to endocrine therapy in hormone receptor-positive human epidermal growth factor receptor 2-negative (HR+/HER2) advanced breast cancer: A meta-analysis and systemic review of randomized clinical trials. Front. Med. 2020. [Google Scholar] [CrossRef]
  118. Hanker, A.B.; Sudhan, D.R.; Arteaga, C.L. Overcoming Endocrine Resistance in Breast Cancer. Cancer Cell 2020, 37, 496–513. [Google Scholar] [CrossRef]
  119. Nakao, M.; Fujiwara, S.; Iwase, H. Cancer Navigation Strategy for Endocrine Therapy-Resistant Breast Tumors. Trends Cancer 2018, 4, 404–407. [Google Scholar] [CrossRef]
  120. Lawrence, B.; Blenkiron, C.; Parker, K.; Tsai, P.; Fitzgerald, S.; Shields, P.; Robb, T.; Yeong, M.L.; Kramer, N.; James, S.; et al. Recurrent loss of heterozygosity correlates with clinical outcome in pancreatic neuroendocrine cancer. NPJ Genom. Med. 2018, 3, 18. [Google Scholar] [CrossRef] [Green Version]
  121. Amin, R.; Tripathi, K.; Sanderson, R.D. Nuclear Heparanase Regulates Chromatin Remodeling, Gene Expression and PTEN Tumor Suppressor Function. Cells 2020, 9, 2038. [Google Scholar] [CrossRef]
  122. Yang, J.; Yin, Y. PTEN in Chromatin Remodeling. Cold Spring Harb Perspect Med. 2020, 10, a036160. [Google Scholar] [CrossRef] [PubMed]
  123. Ding, Y.; Li, N.; Dong, B.; Guo, W.; Wei, H.; Chen, Q.; Yuan, H.; Han, Y.; Chang, H.; Kan, S.; et al. Chromatin remodeling ATPase BRG1 and PTEN are synthetic lethal in prostate cancer. J. Clin. Investig. 2019, 129, 759–773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Taylor, R.A.; Fraser, M.; Rebello, R.J.; Boutros, P.C.; Murphy, D.G.; Bristow, R.G.; Risbridger, G.P. The influence of BRCA2 mutation on localized prostate cancer. Nat. Rev. Urol 2019, 16, 281–290. [Google Scholar] [CrossRef] [PubMed]
  125. Kwong, A.; Shin, V.Y.; Ho, J.C.; Kang, E.; Nakamura, S.; Teo, S.H.; Lee, A.S.; Sng, J.H.; Ginsburg, O.M.; Kurian, A.W.; et al. Comprehensive spectrum of BRCA1 and BRCA2 deleterious mutations in breast cancer in Asian countries. J. Med. Genet. 2016, 53, 15–23. [Google Scholar] [CrossRef] [Green Version]
  126. Berry, J.L.; Polski, A.; Cavenee, W.K.; Dryja, T.P.; Murphree, A.L.; Gallie, B.L. The RB1 Story: Characterization and Cloning of the First Tumor Suppressor Gene. Genes 2019, 10, 879. [Google Scholar] [CrossRef] [Green Version]
  127. Dyson, N.J. RB1: A prototype tumor suppressor and an enigma. Genes Dev. 2016, 30, 1492–1502. [Google Scholar] [CrossRef] [Green Version]
  128. Palin, K.; Pitkanen, E.; Turunen, M.; Sahu, B.; Pihlajamaa, P.; Kivioja, T.; Kaasinen, E.; Valimaki, N.; Hanninen, U.A.; Cajuso, T.; et al. Contribution of allelic imbalance to colorectal cancer. Nat. Commun. 2018, 9, 3664. [Google Scholar] [CrossRef]
  129. Bhagirath, D.; Yang, T.L.; Tabatabai, Z.L.; Shahryari, V.; Majid, S.; Dahiya, R.; Tanaka, Y.; Saini, S. Role of a novel race-related tumor suppressor microRNA located in frequently deleted chromosomal locus 8p21 in prostate cancer progression. Carcinogenesis 2019, 40, 633–642. [Google Scholar] [CrossRef]
  130. Kang, X.L.; Zou, H.; Pang, L.J.; Hu, W.H.; Zhao, J.; Qi, Y.; Liu, C.X.; Hu, J.M.; Tang, J.X.; Li, H.A.; et al. Chromosomal imbalances revealed in primary renal cell carcinomas by comparative genomic hybridization. Int. J. Clin. Exp. Pathol. 2015, 8, 3636–3647. [Google Scholar]
  131. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  132. Sima, J.; Gilbert, D.M. Complex correlations: Replication timing and mutational landscapes during cancer and genome evolution. Curr. Opin. Genet. Dev. 2014, 25, 93–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Guren, M.G. The global challenge of colorectal cancer. Lancet Gastroenterol. Hepatol. 2019, 4, 894–895. [Google Scholar] [CrossRef] [Green Version]
  134. Dekker, E.; Tanis, P.J.; Vleugels, J.L.A.; Kasi, P.M.; Wallace, M.B. Colorectal cancer. Lancet 2019, 394, 1467–1480. [Google Scholar] [CrossRef]
  135. Ongen, H.; Andersen, C.L.; Bramsen, J.B.; Oster, B.; Rasmussen, M.H.; Ferreira, P.G.; Sandoval, J.; Vidal, E.; Whiffin, N.; Planchon, A.; et al. Putative cis-regulatory drivers in colorectal cancer. Nature 2014, 512, 87–90. [Google Scholar] [CrossRef]
  136. Rooney, M.S.; Shukla, S.A.; Wu, C.J.; Getz, G.; Hacohen, N. Molecular and genetic properties of tumors associated with local immune cytolytic activity. Cell 2015, 160, 48–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Schumacher, T.N.; Schreiber, R.D. Neoantigens in cancer immunotherapy. Science 2015, 348, 69–74. [Google Scholar] [CrossRef] [Green Version]
  138. Campoli, M.; Ferrone, S. HLA antigen changes in malignant cells: Epigenetic mechanisms and biologic significance. Oncogene 2008, 27, 5869–5885. [Google Scholar] [CrossRef] [Green Version]
  139. Seliger, B.; Ferrone, S. HLA Class I Antigen Processing Machinery Defects in Cancer Cells-Frequency, Functional Significance, and Clinical Relevance with Special Emphasis on Their Role in T Cell-Based Immunotherapy of Malignant Disease. Methods Mol. Biol. 2020, 2055, 325–350. [Google Scholar] [CrossRef]
  140. Durgeau, A.; Virk, Y.; Gros, G.; Voilin, E.; Corgnac, S.; Djenidi, F.; Salmon, J.; Adam, J.; de Montpreville, V.; Validire, P.; et al. Human preprocalcitonin self-antigen generates TAP-dependent and -independent epitopes triggering optimised T-cell responses toward immune-escaped tumours. Nat. Commun. 2018, 9, 5097. [Google Scholar] [CrossRef] [Green Version]
  141. Anichini, A.; Perotti, V.E.; Sgambelluri, F.; Mortarini, R. Immune Escape Mechanisms in Non Small Cell Lung Cancer. Cancers 2020, 12, 3605. [Google Scholar] [CrossRef]
  142. Yu, S.; Zhao, Z.; Chen, L.; Gu, T.; Yu, H.; Tang, H.; Wang, Q.; Wu, Y. HLA loss of heterozygosity-mediated discordant responses to immune checkpoint blockade in squamous cell lung cancer with renal metastasis. Immunotherapy 2020. [Google Scholar] [CrossRef]
  143. Tran, E.; Robbins, P.F.; Lu, Y.C.; Prickett, T.D.; Gartner, J.J.; Jia, L.; Pasetto, A.; Zheng, Z.; Ray, S.; Groh, E.M.; et al. T-Cell Transfer Therapy Targeting Mutant KRAS in Cancer. N. Engl. J. Med. 2016, 375, 2255–2262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. McGranahan, N.; Rosenthal, R.; Hiley, C.T.; Rowan, A.J.; Watkins, T.B.K.; Wilson, G.A.; Birkbak, N.J.; Veeriah, S.; Van Loo, P.; Herrero, J.; et al. Allele-Specific HLA Loss and Immune Escape in Lung Cancer Evolution. Cell 2017, 171, 1259–1271.e11. [Google Scholar] [CrossRef]
  145. Negrao, M.V.; Quek, K.; Zhang, J.; Sepesi, B. TRACERx: Tracking tumor evolution to impact the course of lung cancer. J. Thorac. Cardiovasc. Surg. 2018, 155, 1199–1202. [Google Scholar] [CrossRef] [Green Version]
  146. Jamal-Hanjani, M.; Hackshaw, A.; Ngai, Y.; Shaw, J.; Dive, C.; Quezada, S.; Middleton, G.; de Bruin, E.; Le Quesne, J.; Shafi, S.; et al. Tracking genomic cancer evolution for precision medicine: The lung TRACERx study. PLoS Biol. 2014, 12, e1001906. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Brastianos, P.K.; Carter, S.L.; Santagata, S.; Cahill, D.P.; Taylor-Weiner, A.; Jones, R.T.; Van Allen, E.M.; Lawrence, M.S.; Horowitz, P.M.; Cibulskis, K.; et al. Genomic Characterization of Brain Metastases Reveals Branched Evolution and Potential Therapeutic Targets. Cancer Discov. 2015, 5, 1164–1177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Vogelstein, B.; Fearon, E.R.; Kern, S.E.; Hamilton, S.R.; Preisinger, A.C.; Nakamura, Y.; White, R. Allelotype of colorectal carcinomas. Science 1989, 244, 207–211. [Google Scholar] [CrossRef] [PubMed]
  149. Watkins, T.B.K.; Lim, E.L.; Petkovic, M.; Elizalde, S.; Birkbak, N.; Wilson, G.A.; Moore, D.A.; Grönroos, E.; Rowan, A.; Dewhurst, S.M.; et al. Pervasive chromosomal instability and karyotype order in tumour evolution. Nat. Cell Biol. 2020, 587, 126–132. [Google Scholar] [CrossRef] [PubMed]
  150. Linjama, T.; Impola, U.; Niittyvuopio, R.; Kuittinen, O.; Kaare, A.; Rimpilainen, J.; Volin, L.; Perasaari, J.; Jaatinen, T.; Lauronen, J.; et al. Conflicting HLA assignment by three different typing methods due to the ap-parent loss of heterozygosity in the MHC region. HLA 2016, 87, 350–355. [Google Scholar] [CrossRef]
Figure 1. Two types of loss of heterozygosity. There are two main types of loss of heterozygosity (LOH), LOH with copy number losses (CNL-LOH) and copy number neutral LOH (CNN-LOH). Mitotic recombination is considered a major contributor to acquired LOH. During cancer progression, tumor cells may lose one allele (here referred to as A) as a consequence of partial chromosome deletion which is defined as LOH with copy number losses (CNL-LOH). CNL-LOH could further undergo recombination using the homolog (here refers as allele B) as a template for correction which is defined as LOH with copy number neutral (CNN-LOH). Here, orange and green regions represent either of two homologous chromosomes.
Figure 1. Two types of loss of heterozygosity. There are two main types of loss of heterozygosity (LOH), LOH with copy number losses (CNL-LOH) and copy number neutral LOH (CNN-LOH). Mitotic recombination is considered a major contributor to acquired LOH. During cancer progression, tumor cells may lose one allele (here referred to as A) as a consequence of partial chromosome deletion which is defined as LOH with copy number losses (CNL-LOH). CNL-LOH could further undergo recombination using the homolog (here refers as allele B) as a template for correction which is defined as LOH with copy number neutral (CNN-LOH). Here, orange and green regions represent either of two homologous chromosomes.
Pharmaceuticals 14 00057 g001
Figure 2. Targeting LOH in tumors for cancer treatment. When a cancer cell undergoes loss of heterozygosity of non-driver genes (A), further loss or inhibition of a specific allele (B) still retained in the tumor should not be tolerated, whereas normal cells will be able to survive relying solely on the retained allele. Treatment with an agent (purple dots) that is either a selective inhibitor of protein product B or a cytotoxic substrate metabolized by A but not B will result in selective killing of the tumor cells. Thus, approaches focusing on non-driver genes represent an alternative target class that merits further exploration. Here, orange and green regions represent either of two homologous chromosomes and ’X’ indicates the selective kill of cancer cells retaining only allele B.
Figure 2. Targeting LOH in tumors for cancer treatment. When a cancer cell undergoes loss of heterozygosity of non-driver genes (A), further loss or inhibition of a specific allele (B) still retained in the tumor should not be tolerated, whereas normal cells will be able to survive relying solely on the retained allele. Treatment with an agent (purple dots) that is either a selective inhibitor of protein product B or a cytotoxic substrate metabolized by A but not B will result in selective killing of the tumor cells. Thus, approaches focusing on non-driver genes represent an alternative target class that merits further exploration. Here, orange and green regions represent either of two homologous chromosomes and ’X’ indicates the selective kill of cancer cells retaining only allele B.
Pharmaceuticals 14 00057 g002
Table 1. Biomarkers on Chromosomes associated with LOH.
Table 1. Biomarkers on Chromosomes associated with LOH.
Chromosome
Associated with LOH
Gene NameLOH PositionPredictive BiomarkerReference
1pN/A1p36Significantly associated with prognostic markers of aggressive neuroblastoma when patients are diagnosed[61,62,63,64]
3pFRA3B3p13–3p21Predicted early event in the genesis of smoking-related cancers.[65,66]
CACNA2D33p21Associated with poor clinical outcome in esophageal cancer.[67,68]
VHL3p25Predictive biomarker for clinical outcome in clear-cell renal-cell carcinoma (ccRCC) patients.[69]
6qHLA6p21Leads to a poor response to immune checkpoint inhibitors.[70]
Predictive biomarker for patients with synovial sarcoma and is prognostic of poor clinical outcome.[71]
FABP76q22Correlated with survival in patients with glioblastoma.[72]
A potential diagnostic biomarker of early-stage melanoma systemic spreading in blood.[73]
9qN/A9p13Significantly associated with poorer prognosis of glioma patients.[74]
10qPTEN10q23Functionally related to the development of breast cancer, associated with poor prognosis[75]
11qNELL111p15An association between LOH at 11p15.5 and poor survival in 180 lung cancer patients.[76]
16qCD1916p11Irreversible loss of heterozygosity in CD19 could be used as a biomarker for an outcome prediction after the CAR T cells therapy CTL019.[77]
17qTP5317p13One of the most frequent genetic alterations leading to human cancers.[78,79]
BRCA117q21A germline LOH on BRCA1 confront an 85% lifetime risk of breast cancer and a greatly elevated risk of ovarian cancer[80]
18qDPC4 (SMAD4)18q21Assist in predicting the clinical outcome after therapies in colorectal cancer (CRC) patients.[81,82]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, X.; Sjöblom, T. Targeting Loss of Heterozygosity: A Novel Paradigm for Cancer Therapy. Pharmaceuticals 2021, 14, 57. https://0-doi-org.brum.beds.ac.uk/10.3390/ph14010057

AMA Style

Zhang X, Sjöblom T. Targeting Loss of Heterozygosity: A Novel Paradigm for Cancer Therapy. Pharmaceuticals. 2021; 14(1):57. https://0-doi-org.brum.beds.ac.uk/10.3390/ph14010057

Chicago/Turabian Style

Zhang, Xiaonan, and Tobias Sjöblom. 2021. "Targeting Loss of Heterozygosity: A Novel Paradigm for Cancer Therapy" Pharmaceuticals 14, no. 1: 57. https://0-doi-org.brum.beds.ac.uk/10.3390/ph14010057

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

Article Metrics

Back to TopTop