Next Article in Journal
Acknowledgment to Reviewers of Life in 2020
Previous Article in Journal
Smart Nanoformulation Based on Polymeric Magnetic Nanoparticles and Vincristine Drug: A Novel Therapy for Apoptotic Gene Expression in Tumors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cancer, Retrogenes, and Evolution

Institute of Human Biology and Evolution, Faculty of Biology, Adam Mickiewicz University in Poznan, 61-614 Poznan, Poland
*
Author to whom correspondence should be addressed.
Submission received: 19 December 2020 / Revised: 14 January 2021 / Accepted: 15 January 2021 / Published: 19 January 2021
(This article belongs to the Section Evolutionary Biology)

Abstract

:
This review summarizes the knowledge about retrogenes in the context of cancer and evolution. The retroposition, in which the processed mRNA from parental genes undergoes reverse transcription and the resulting cDNA is integrated back into the genome, results in additional copies of existing genes. Despite the initial misconception, retroposition-derived copies can become functional, and due to their role in the molecular evolution of genomes, they have been named the “seeds of evolution”. It is convincing that retrogenes, as important elements involved in the evolution of species, also take part in the evolution of neoplastic tumors at the cell and species levels. The occurrence of specific “resistance mechanisms” to neoplastic transformation in some species has been noted. This phenomenon has been related to additional gene copies, including retrogenes. In addition, the role of retrogenes in the evolution of tumors has been described. Retrogene expression correlates with the occurrence of specific cancer subtypes, their stages, and their response to therapy. Phylogenetic insights into retrogenes show that most cancer-related retrocopies arose in the lineage of primates, and the number of identified cancer-related retrogenes demonstrates that these duplicates are quite important players in human carcinogenesis.

1. Introduction

A large part of the eukaryotic genome contains sequences that result from the activity of transposable elements. Until recently, most of them were considered insignificant. It turned out that transposable elements importantly influenced the evolution of genomes. Identifying the possible functions of “junk DNA” constitutes one of the greatest discoveries in genomic analyses [1,2,3]. The term “junk DNA” also refers to pseudogenes; however, an increasing number of studies support the functionality of many pseudogenes and their role in various human diseases. Findings from studies in cell culture, animal models, and clinical samples confirm the role of pseudogenes in tumorigenesis [4,5,6]. This includes retrocopies that are usually called “processed pseudogenes” or “retropseudogenes” and have been classified as gene copies with no functional significance. Nevertheless, as the quantity of data from high-throughput experiments increases, new functionally important retrocopies are identified, including those associated with various diseases [7,8], especially with many types of cancer [9,10].
The phenomenon of neoplasm is widespread across the evolutionary tree, but its incidence varies between species. Surprisingly, the risk of cancer transformation does not seem to depend strongly on individual species’ body size or life expectancy [11]. Studies on a group of elephants have shown that among them, there is a low rate of tumor transformation compared to other mammals. Elephants have 20 copies of the TP53 gene, a well-known oncosuppressor, 19 of which are retropseudogenes. The presence of this gene’s additional copies is proposed to be related to an increased apoptotic response in the elephant population [12]. The occurrence of resistance mechanisms to cancer has also been described in other animal species. A good example is the unusual tolerance to hypoxia in naked mole-rats [13].
Many studies have described the relationship between the expression level of retrogenes and the incidence of specific neoplasms [14,15], and the connection between these two appears to go beyond this. Retrogenes are known as important evolutionary players that can affect genome diversity [16,17,18,19,20], which is due to, among other things, their much faster evolution than in the case of protein-coding genes. In cancer, there is also a rapid accumulation of mutations and the formation of qualitatively different populations of cancer cells. In addition, it has been demonstrated that internal diversity is the result of natural selection, which shapes the tumor during its development [11]. Natural selection evidently also determines the fate of retrocopies. These connections between retrogenes and cancer gave reasons for an evolutionary-based overview of retrogenes in cancerous contexts (Figure 1).

2. Retrocopies and Their Functions

New gene copies may be obtained by polyploidization, irregular crossing over, or DNA- or RNA-mediated duplication [21]. Until recently, only DNA-based duplication was considered to be functionally relevant. However, later studies have revealed that RNA-based duplication provides copies that may play a vital role in the cell [3,22,23,24].
The formation of retrocopy begins with the transcription of a parental gene. The processed mRNA goes to the cytoplasm, where L1 retrotransposon-derived proteins bind to its polyA tail. The process takes place with the participation of reverse transcriptase, endonuclease, and chaperones. Parental gene’s mRNA anneals to the broken DNA ends, undergoing reverse transcription, and the resulting cDNA is integrated back into the genome in the form of a retrocopy (Figure 2a). Retrocopies are devoid of introns and regulatory elements. They are equipped with a poly-A tail along with flanking repeats. These copies were long considered to be “dead-on-arrival” and classified as transcriptional noise due to their high similarity to the parental genes. Nevertheless, to promote transcription, retrotransposed transcripts can take advantage of adjacent gene regulatory regions and use distant CpG sequences or sometimes even parts of their own sequences. Furthermore, retrocopy insertion into the intron of another gene often leads to the acquisition of the host’s regulatory machinery [22,25].
Retrocopies that acquired transcriptional ability are called retrogenes. In the process of evolution, they may be subject to subfunctionalization and take over some of the parental genes’ functions (Figure 2b). A good example is retrogene SLC5A3 and the parental gene SLC5A1. Proteins encoded by these genes contain solute binding domains, but they differ in activity. Retrocopy-derived protein is a sodium-dependent myo-inositol transporter. In turn, parental-derived proteins participate in the transport of glucose and galactose [24]. Another functional evolution path of retrocopies is neofunctionalization. As a result, they can encode proteins, novel or similar to those encoded by the parental gene, or they can obtain regulatory functions and be involved in transcriptional regulation of parental counterparts or other genes. They can also participate in transcriptional interference, be a source of different small RNAs, or act as miRNA sponges [24,26,27]. Retrogene-derived RNAs can also be involved in epigenetic regulation [28] or function as trans-NATs (natural antisense transcripts) [24,29]. Moreover, it was demonstrated that retrogenes can functionally replace their parental genes [8]. Retrocopies may also contribute to other genes and/or transcripts; they can create chimeric transcripts, act as recombination hot spots [24], or provide a sequence for alternative exons [30].

3. Retrogenes and Evolution

It is widely understood that adaptive features, lineage-specific phenotypic traits, are associated with the formation of new genes [22]. Gene duplication is a primary mechanism of new gene formation by providing a substrate for natural selection. The additional copies of ancestral genes are subject to less evolutionary restriction to develop a novel feature [21]. They accumulate mutations faster than protein-coding genes and thus evolve faster [31]. Within the different types of duplication, there are differences in the susceptibility to evolutionary changes. Duplication at the DNA level results in daughter copies with full equipment (core promoters and gene organization). Therefore, these duplicates mostly mirror the protein function and expression pattern of their ancestor [25]. In contrast, analysis of retrogenes has pointed to their significant contribution to molecular evolution as a source of genomic novelties, and they are called “seeds of evolution” [1]. Due to the lack of regulatory elements, transcribed retrocopies must acquire regulatory regions. Thus, retrocopies are probably more predisposed to evolve a novel expression pattern and functional role than copies emerging from segmental duplication. Moreover, retrogenes play a role in gene structure evolution by mediating the decline of introns [22]. Nevertheless, retrocopies may gain introns or additional exons over time. Szcześniak et al. reported two retrogenes, RNF113B and DCAF12, where introns were created through mutations and the appearance of new splice sites [19]. On the other hand, Vinckenbosch et al. identified 27 intergenic retrogenes that acquired de novo exons [25].
Several studies support the hypothesis that splicing signal conservation constrains the rate of protein evolution [22], it has been suggested that the evolution rate is lower within the exon-intron boundaries and for intron-rich genes [32]. Therefore, splicing constraints impose some limitations on parental gene evolution. However, such constraints should not apply to single exon retrocopies. Interestingly, it has been noted that within the retrocopy sequence, the rate of protein evolution is in fact the strongest within previous splicing junctions in the ancestor gene. Consequently, a more effective adaptation of retrogene-derived protein in comparison with the parental gene’s protein can be speculated as a result of relaxing splicing constraints [22].
Many studies emphasize the role of retrogenes in the differentiation and molecular evolution of genomes [20,24,30] and, as a result, are a source of species-specific features as well as interspecies variation. For example, a retrocopy of the cyclophilin A gene within the owl monkey genome is associated with resistance to HIV [20]. Another example constitutes the rodent-specific retrogene Rps23r1, which reduces Alzheimer’s β-amyloid levels and may cause discrepancies between animal model studies and results of clinical trials, for example [33]. Furthermore, the fgf4 retrogene that is responsible for chondrodysplasia is found only in short-legged dog breeds [17]. Finally, as previously mentioned, the increased number of TP53 gene retrocopies results in a lower cancer transformation rate in the elephant population [12]. Moreover, retrocopy number variation was also observed across human populations. This includes transcriptionally active retrogenes like EIF4A1P10 or TCF3P lost in some members of African populations, for example [34,35].
Numerous analyses have shown that the retroposition process was particularly intensive during the evolution of primates. The intensity of this phenomenon is associated with the occurrence of many retrocopies specific for this order of mammals [36,37]. As a result of this “burst of retroposition”, retropseudogenes belong to the largest group within all human pseudogenes [38]. Many retrogenes have been linked to cancer and a lot of them are human and primate-specific as it is demonstrated further down. Therefore, the question arises whether a large number of human retrogenes are associated with a high risk of neoplastic transformation in our species.

4. Cancer and Evolution

Mutational events form the basis of species evolution as well as cancer development [9]. Cancer tumors are highly dynamic and adaptive systems and evolve very quickly. Evolutionary processes play a role in the progression of cancer on two levels, at the level of species evolution and the level of individual cancer development (Figure 3). The phenomenon of natural selection operates on specific features of the population associated with cancer promotion/suppression [11]. Mechanisms of resistance to tumor transformation have been described in several species [39]. Cancer tumors are also subject to natural selection, and the “branched evolution” of species is reflected in the evolutionary trajectories of cancer cell populations [40].

4.1. Species-Specific Features of Cancer Suppression

The riddle “Peto’s paradox” indicates that the incidence of cancer among animals does not increase with body size and length of life. A good example constitutes already mentioned studies on a group of African elephants (lat. Loxodonta africana) and Asian elephants (lat. Elephas maximus). Among elephants, in comparison to other species, the tumor transformation rate is lower than expected. Unlike human cells, with one copy of the TP53 gene, African elephants have 20 copies, 19 of which arose from retroposition [12]. The TP53 gene encodes the p53 protein, which is called the “genome guardian” [41]. It belongs to key suppressor genes, and TP53 mutations have been observed in most human cancers [42,43]. Disruption of p53 protein function causes the occurrence of cancer cell features [12]. The presence of extra copies results in an effective DNA damage response through the hyperactive TP53 pathway (Figure 3a) [44].
A similar phenomenon has been observed in the population of the long-lived (approximately 200 years) bowhead whale (lat. Balaena mysticetus), although the resistance mechanism is not entirely clear. This has been linked to the positive selection force acting on cancer and aging genes involved in DNA damage repair and thermoregulation, ERCC1 and UCP1, respectively. Duplication of the PCNA gene, one of the essential repair mechanism genes, has been reported. This can reduce the frequency of mutations and thus prevent tumorigenesis [45,46].
A species-specific cancer defense mechanism was also uncovered in the long-lived rodent lineage, including the naked mole-rat (lat. Heterocephalus glaber) [13]. The naked mole-rat resistance mechanism is based on the limitation of cell proliferation through the expression of high molecular mass hyaluronan (HMM-HA). The longer variant of HMM-HA inhibits divisions, inflammation, and metastatic processes [47]. It is quite interesting that in the naked mole-rat genome, 17 additional copies of PTEN, an important tumor suppressor gene, have been also reported. This may additionally contribute to such strong resistance to cancer [48]. Another example of a rodent that has evolved a way to suppress cancer is the blind mole-rat (lat. Spalax ehrenbergi). In this case, a subterranean lifestyle is connected to unusual tolerance to hypoxia. This has been associated with alterations in the TP53 gene sequence. Similar changes have been identified in hypoxia-tolerant human tumors [13,49]. Interestingly, despite many studies, no cases of malignant neoplasm have been found in this species [44].
The literature data also report bats as a relatively long-lived species. It has been suggested that the ability to fly as an energy-intensive activity has caused the evolution of mechanisms that inhibit oxidative stress. Furthermore, DNA damage checkpoint genes are under positive selection in this case. These effects may be related to cancer resistance in the bat population [39].

4.2. Evolution within Cancer Tumors

The impact of evolutionary forces is also visible in cancer cells. It is supposed that a tumor basically consists of copies of a single cell. During tumor development, neoplastic changes (e.g., mutations) create heterogeneous masses—the starting point for the operation of evolutionary pressure. However, mutations are not the only forces shaping cancer evolution. Successive changes lead to the formation of different genetic subclones (Figure 3b) [11,40]. At this point, the evolutionary selection is also starting to play a role, and one group of tumor cells may be evolutionarily favored. This situation may occur when some cells develop traits that give them an advantage in a particular tumor environment. As a result, these cells will have more “offspring” than others. Furthermore, the high degree of tumor diversity and genomic instability results in a high risk of an adaptive mutation, which in turn is related to a faster progression of the disease [50].
Additional confirmation of the evolutionary forces acting within cancer comes from genetic changes in the response to a particular drug (Figure 3c). The diagnosed tumor may consist only of cells that are sensitive to treatment, and the patient has a good prognosis. However, among some fraction of patients, after applying certain therapies, the response is observed only at the initial phase and, unfortunately, tumor progression unexpectedly accelerates. This may indicate the presence of therapy-resistant subpopulations in the pretreated tumor. It has been also noted that therapy-derived selective pressure can determine the growth of therapy-resistant populations and induce the onset of “acquired resistance”. Therefore, adaptive therapy has been proposed, whereby maintaining a population of drug-sensitive cells limits the growth of populations resistant to treatment. The evolution-based approach relies on the combination of different drugs or their doses to slow tumor proliferation. It is suggested to use repeated optimal doses to reduce tumor volume rather than destroying it. This less aggressive approach may allow better control for the tumor and prevent the development or widespread of more aggressive, treatment-resistant form. This switch from the traditional approach that bases on maximal cell death to maximum progression-free survival could improve cancer treatment outcomes [40,51,52,53].
Just as the evolutionary history of a given species has led to differences in susceptibility to cancer, so does the history of tumor development influence the response to applied oncological treatment. Thus, determining the course of the evolutionary history of the tumor is important for establishing the best oncological treatment for a particular type of tumor and the stage of its development. The evolution of species-specific resistance mechanisms to cancer and the occurrence of specific tumor types and patient responses to treatment have been linked in some cases with retroposed genes as described below.

5. Retrogenes in Cancer

Identification of new biomarkers that will help predict a series of events in cancer evolution would certainly lead to more effective diagnostics and treatment. Retrocopies seem to be perfect candidates. The literature has shown a relationship between the expression level of some retrogenes and the occurrence of specific cancers [14,15]. Retrocopies involved in the response to a particular treatment, such as radiation [10] or paclitaxel [54], have also been reported [10,54]. Moreover, many studies describe retropseudogenes associated with the occurrence of a particular stage or form of the tumor [55,56,57]. It turns out that they can play multifaceted roles within tumor cells, and the literature reports retrogenes that are both oncogenes and tumor suppressors. A list of the cancer-related retrocopies that have been described so far in the literature is presented in Table 1.

5.1. Increased Expression in Cancer

Retrogenes with elevated expression in cancer constitute a large group, and in many cases, increased expression of these retrocopies promotes cancer development. The expression levels of retrocopy KRASP1 are correlated with the prostate cancer phenotype. Its parental gene—KRAS—belongs to one of the most well-known oncogenes. Cancer cell line studies have shown that KRASP1 overexpression causes increased parental gene expression and cell proliferation [59]. It has also been hypothesized that predisposition to ovarian cancer is associated with the expression of the small subunit processome component UTP14C, a protein-coding retrocopy of the UTP14A gene. This was explained by UTP14C downregulation of TP53 levels, which leads to the prevention of cell cycle arrest and apoptosis [60]. Upregulated expression of MSL3P1, male-specific lethal-3 homolog pseudogene 1, has been correlated with renal cell carcinoma [61]. Other examples of retrocopies overexpressed in cancer tissues include ANXA2P2 [62], CSDAP1 [63], LGMNP1 [64], UBE2CP3 [65], RACGAP1P [66], PTTG3P [67], and CKS1BP7 [68].
Analyses of RNA-seq data performed in our group revealed that more retrogenes may be associated with cancer. Differential expression analysis allowed the identification of 3 potential markers with increased expression levels characteristic of breast cancer, RPL5P4, ASS1P2, and AC007731.2, and 8 retrocopies with elevated expression in lung adenocarcinoma, PTBP1P, AL121949.1, HNRNPA3P9, retro_hsap_4319, AC090695.2, CDK8P2, MSL3P1, and POLR3GP1 [89]. Retro_hsap_4319 is a novel retrogene, i.e., not annotated in the reference genome, placed in the RetrogeneDB database [58].

5.2. Decreased Expression in Cancer

Retrogenes associated with tumor suppression have also been reported. Downregulation levels of PTENP1 have been associated with gastric cancer and renal cell carcinoma. The PTENP1 functions as a miRNA sponge. A decreased level of PTENP1 contributes to increased degradation of its oncosuppressive parental gene, PTEN, which exerts a growth-inhibitory role within the tumor [15,59,69,70,71]. Another example is INTS6P1 retrocopy. Its lower serum levels correspond to hepatocellular carcinoma. Interestingly, the diagnostic power of this retrocopy is comparable to the most common biomarker for hepatocellular carcinoma—α-fetoprotein [15,72]. In turn, the expression of TUSC2P1 retrocopy suppresses the proliferation and migration of cancer cells and promotes apoptosis. This duplicate share sites for miRNAs with its progenitor TUSC2 gene, thereby regulating its expression. The interaction with common miRNAs promotes parental gene expression and results in inhibition of proliferation, migration restriction, and apoptosis induction [73]. An additional example of a tumor suppressor is NKPL. Downregulation of this retrocopy is connected with lower overall survival in several cancers, including kidney renal papillary cell carcinoma, pancreatic adenocarcinoma, and adenoid cystic carcinoma [74]. A decreased expression level of CTNNA1P1 has been associated with the pathogenesis of colorectal cancer. Suppressive action of the cognate gene CTNNA1 has also been shown in several tumors [75]. An example of the well-described retrogene in the cancer literature is RHOB exhibiting suppression activity. Decreased expression of RHOB has been reported in many cancer studies [76,77,78,90].
The previously mentioned analysis of breast cancer samples led to the identification of 17 additional retrocopies with decreased expression levels (AC104212.2, RHOQP2, NKAPL, RPL21P16, RBMS1P1, retro_hsap_2623, DIO3, FAM122A, RPSAP70, PTENP1, AC138392.1, DHFR2, CTB-50E14.5, AK4P1, RAB43P1, PSMA2P1, and RBMXL1). In the lung cancer cohort, 13 retrogenes showed decreased expression in cancer (RPL13AP17, HNRNPA1P33, SIRPAP1, AL136982.4, AL136452.1, AC084880.1, HMGN2P15, CDC20P1, AC022217.1, DIO3, HMGB3P10, BET1P1, and TMED10P2) [89].

5.3. Subtype-Specific Retrogenes

Differences in the expression level of retrogenes were also observed depending on the subtype of cancer. Retrocopies of the HMGA1 gene have been related to the occurrence of anaplastic thyroid carcinoma. HMGA1P6 and HMGA1P7 have oncogenic activity and contribute to cancer progression. In well-differentiated and weakly aggressive papillary thyroid carcinoma, HMGA1P6 and HMGA1P7 were not identified. In turn, anaplastic thyroid carcinoma, one of the most malignant cancers in humans, expresses high levels of these retrogenes [14]. Interestingly, a similar relationship has been noted among patients with endometrial cancer—increased expression levels of HMGA1P6 and HMGA1P7 correlate with the malignant phenotype [80]. Another good example is the upregulated SUMO1P3 retrocopy in gastric cancer patients, which has the marker potential to differentiate between cancer and benign gastric disease [15,82].
In our laboratory breast cancer analysis, two retrocopies with differential expression characteristics for the ER+ (estrogen receptor-positive) subtype, AC098591.2, and PABPC4L, and 7 downregulated retrocopies in the TNBC (triple-negative breast cancer) subtype, RAB6C, RPS16P5, RHOB, MEIS3P2, PGAM1P5, HMGN2P15, and KRT8P13, were identified [89].

5.4. Stage-Specific Retrogenes

Increased expression of NANOGP8 and POU5F1P4/P5 retrogenes has been correlated with the phenotype of cancer stem cells (CSCs). The occurrence of this subpopulation, with high metastatic capacity, heralds intensive tumor expansion. In addition, the altered expression of a retrocopy of parental genes associated with maintaining pluripotency (NANOG and POU5F1) may also be a sign of early disease relapse [27,55]. It is worth noting that the knockdown of NANOG and NANOGP8 reduces the malignant transformation in prostate cancer cells [85]. Another example of cancer stage-specific retrocopy is SLC6A6P1, also known as SLC6A610P, associated with recurrence in high-grade serous ovarian cancer. This subtype of ovarian cancer is very common (over 70% of affected women) [91]. Moreover, due to the lack of reliable diagnostics, it is usually detected at an advanced stage [87].
The expression of the PDIA3P1 retrocopy was significantly increased in hepatocellular carcinoma. Interestingly, it has been demonstrated that the PDIA3P1 expression level is related to metastasis and TNM stage and that a knockdown of retrocopy causes reduced migration and invasion of cancer cells [56]. One of the metastasis-related retrocopies is the previously described CTNNA1P1, whose expression has been significantly correlated with node metastasis in colorectal cancer patients [75]. Cooke et al., sequenced samples from different stages of lung and colon cancer. Their analysis revealed retrocopies unique for a given stage. Nevertheless, they have also found several processed pseudogenes that are expressed in both the primary tumor and metastasis [9].

5.5. Treatment Response-Related Retrogenes

Examples of retrogenes that are associated with the response to a particular treatment can also be found in the literature. The expression of the retrocopy PPIAP43, for instance, has been correlated with radiosensitivity in a patient with small-cell lung cancer [10]. This discovery is quite important since radiation constitutes the main strategy in the case of this cancer. The sensitivity to radiation differs among oncological patients, but to date, there is no suitable biomarker. Another example is the FTH1P3 retrocopy, which promotes ABCB1 expression by sponging miR-206. As a result, resistance to paclitaxel is activated in breast cancer patients [54].
The relationship between the sequence variant of a given retrocopy and the individual’s prognosis has also been described. The occurrence of the E2F3P1 GA/AA allele at the rs9909601 locus has been associated with higher overall survival among hepatocellular carcinoma patients [15,88].

5.6. mRNA Sponging as the Main Retrocopy Mechanism in Cancer

A leading role of retrogenes described in the literature is sponging the miRNAs. This posttranscriptional process regulates parental or other genes when they share binding sites for miRNAs [6]. Recently, a genome-wide analysis demonstrated that as many as 181 retrocopies putatively regulate 250 transcripts of 187 genes [24].
Under normal conditions, there is a balance in the expression level of retrocopies. Sufficient expression of retrocopies regulates suppressor genes by competing for shared miRNAs. This prevents suppressor gene transcript degradation and enables the translation process (Figure 4a). Consequently, the low expression level of retrogenes contributes to increased miRNA binding to suppressor genes, which drives them on the path of degradation and promotes cancer transformation (Figure 4c). A good example represents the decreased expression of the PTENP1 retrocopy in cancer [15].
The opposite is true in the case of oncogenes. In a normal cell, low expression of a retrogene that shares binding sites with oncogenes results in a lack of competition for common miRNAs. As a result, miRNAs bind to oncogene mRNAs and direct them on the degradation path (Figure 4b). Under cancer conditions, elevated expression of a given retrocopy causes sponging miRNAs and prevents oncogene degradation (Figure 4d), which leads to cancer development. This type of relationship occurs in the abovementioned HMGA1 gene and its retrocopies [14].

6. Phylogeny of Cancer-Related Retrogenes

We used the GenTree database (http://gentree.ioz.ac.cn/) [92] to determine the time of cancer-related retrogene origination. Figure 5 represents the estimated point of the origin of earlier described retrogenes (no data for TUSC2P1, retro_hsap_2623, or retro_hsap_4319). Some of them are characterized by heterogeneous origins (“patchy tree”), but future research is needed to establish whether this results from independent retroposition events or loss of retrocopy in some species.
The oldest retrocopies recognized as cancer-related are DIO3 and RHOB. Both arose during the early evolution of vertebrates and represent protein-coding retrogenes. They are widely distributed in animal genomes and well described in the literature. Conservation of ORFs in these quite old retrogenes may indicate that they acquired transcriptional capabilities very quickly and the propensity to accumulate mutations, typical for retrocopies, was “locked” due to the functional importance of gene products. RHOB is an important oncosuppressor, and a decrease in its expression promotes cancer, as described earlier. Increased expression of the DIO3 retrocopy has been related to tumor progression in papillary thyroid cancer and colon cancer [93]. In turn, a decrease in the level of DIO3 expression has been described in lung and breast cancer [89]. Old retrocopies, well recognized in human cancers and present in the genomes of all bony vertebrates, are great candidates for studies of the origins of neoplastic processes. The study of these genes may also be valuable for uncovering common features of tumors among species that are far away in the evolutionary tree. Nevertheless, a lack of data regarding species other than humans and mice seems to be the greatest difficulty in performing such research.
The majority of human cancer-related retrocopies are specific for primates. A large part of these groups arose before the split of New and Old-World Monkeys. There is also a group of retrocopies that are present in the human genome only. One of these is NANOGP8, the retrocopy of the NANOG gene that has 11 pseudogenes. Ten of them are derived from retroposition, and NANOGP8 is evolutionarily the youngest. Interestingly, in the chimpanzee genome, all NANOG copies can be found, except NANOGP8 [55]. Other cancer-related retrocopies unique to humans include AK4P1, RAB43P1, RPL21P16, and AC138392.1. Changes in the expression level in cancer have also been detected in the case of their parental genes, AK4 in lung cancer [94], RAB43 in gliomas [95], and RPL21 in breast cancer [96].
The role of the newly arose genes is intriguing from an evolutionary point of view. It has been suggested that the presence of genes characteristic of a given lineage is related to phenotypic adaptation [97]. Furthermore, it was noted that the expression of some evolutionarily young genes occurs specifically or preferentially in tumors. These genes were termed tumor-specifically expressed, evolutionarily novel (TSEEN) [38,98]. Moreover, it has been reported that new genes are also overrepresented among the testis and brain. It has been hypothesized that new genes can be recruited to processes under strong selection pressure (e.g., spermatogenesis, immune response) or processes involving novel organ development (placenta, expanded brain) [92]. We searched the GTEx database to assess the expression levels of cancer-related retrocopies in human normal tissues (Table S1) [99]. The results of the GTEx analysis are in agreement with the abovementioned statements. In the analyzed cohort, there were 25 retrocopies with low or no expression levels in normal tissues (max median TPM < 1). These genes are good candidates for potential TSEEN genes. Additionally, eight neoplastic retrocopies are active mainly in testis and brain tissues: CSDAP1, AC022217.1, SLC6A6P1, KRT8P13, MSL3P1, HMGA1P7, RACGAP1P, and NKAPL.
As we stated in the previous chapters, the process of tumorigenesis is common across the evolutionary tree. The study of human-specific retrocopies and differences in the retrocopy repertoire across vertebrate species may be essential for better understanding the high rate of neoplastic processes in our species. Moreover, retrogenes included in the TSEEN group, due to their species and tissue specificity, represent great potential as new tumor biomarkers.

7. Conclusions

This review aims to gain insight into the link between the evolutionary origin of high cancer incidence among humans and protein-coding gene retroposition. Retrocopies are known evolutionary players in the context of species-specific traits and are also involved in the carcinogenesis process. It is possible that retrogenes, as important elements involved in the evolution of species, also count in the “microevolution” of cancer tumors. The studies discussed herein underline that investigation of retrogene expression can be a useful diagnostic tool, especially when conventional clinical methods are not sufficient. Nevertheless, retrocopies are often omitted as candidate genes and considered artifacts. Low interest in these sequences in cancer studies may also come from the fact that analysis of their expression is a significant challenge due to their low level compared to genes encoding proteins and high level of sequence similarity with their progenitors. However, the number of retrocopies identified thus far demonstrates that these gene duplicates are quite important actors in carcinogenesis.
Considering the expanding research in the field of evolutionary medicine, retrogene analysis seems to be a promising direction for future investigation. The studies summarized here open the way for using retrogene expression evaluation both to indicate cancer subtype/stage and to predict patient treatment responses. This may improve cancer diagnostics, customize more tailored therapy, and affect the prognosis of oncological patients. In addition, the evolutionary-based view of cancer can provide information about human-specific traits and the direction of our evolution.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2075-1729/11/1/72/s1, Table S1: GTex analysis.

Author Contributions

Conceptualization, I.M. and K.S.; investigation, K.S.; writing—original draft preparation, K.S.; writing—review and editing, I.M. and K.S.; visualization, K.S.; supervision, I.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brosius, J. Retroposons—Seeds of evolution. Science 1991, 251, 753. [Google Scholar] [CrossRef] [PubMed]
  2. Brosius, J. The contribution of RNAs and retroposition to evolutionary novelties. Genetica 2003, 118, 99–115. [Google Scholar] [CrossRef] [PubMed]
  3. Kubiak, M.R.; Makałowska, I. Protein-coding genes’ retrocopies and their functions. Viruses 2017, 9, 80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kalyana-Sundaram, S.; Kumar-Sinha, C.; Shankar, S.; Robinson, D.R.; Wu, Y.M.; Cao, X.; Asangani, I.A.; Kothari, V.; Prensner, J.R.; Lonigro, R.J.; et al. Expressed pseudogenes in the transcriptional landscape of human cancers. Cell 2012, 149, 1622–1634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Han, L.; Yuan, Y.; Zheng, S.; Yang, Y.; Li, J.; Edgerton, M.E.; Diao, L.; Xu, Y.; Verhaak, R.G.W.; Liang, H. The Pan-Cancer analysis of pseudogene expression reveals biologically and clinically relevant tumour subtypes. Nat. Commun. 2014, 5, 3963. [Google Scholar] [CrossRef] [Green Version]
  6. Hu, X.; Yang, L.; Mo, Y.Y. Role of pseudogenes in tumorigenesis. Cancers 2018, 10, 256. [Google Scholar] [CrossRef] [Green Version]
  7. Grzybowska, E.A. Human intronless genes: Functional groups, associated diseases, evolution, and mRNA processing in absence of splicing. Biochem. Biophys. Res. Commun. 2012, 424, 1–6. [Google Scholar] [CrossRef]
  8. Ciomborowska, J.; Rosikiewicz, W.; Szklarczyk, D.; Makałowski, W.; Makałowska, I. “Orphan” retrogenes in the human genome. Mol. Biol. Evol. 2013, 30, 384–396. [Google Scholar] [CrossRef]
  9. Cooke, S.L.; Shlien, A.; Marshall, J.; Pipinikas, C.P.; Martincorena, I.; Tubio, J.M.C.; Li, Y.; Menzies, A.; Mudie, L.; Ramakrishna, M.; et al. Processed pseudogenes acquired somatically during cancer development. Nat. Commun. 2014, 5, 1–9. [Google Scholar] [CrossRef]
  10. Wang, S.; Yu, J. Long non-coding RNA transcribed from pseudogene PPIAP43 is associated with radiation sensitivity of small cell lung cancer cells. Oncol. Lett. 2019, 18, 4583–4592. [Google Scholar] [CrossRef] [Green Version]
  11. Donnem, T.; Micklem, K.; Pezzella, F. Evolution and cancer. In Oxford Textbook of Cancer Biology, 1st ed.; Pezzella, F., Tavassoli, M., Kerr, D.J., Eds.; Oxford University Press: Oxford, UK, 2019; pp. 33–41. [Google Scholar]
  12. Abegglen, L.M.; Caulin, A.F.; Chan, A.; Lee, K.; Robinson, R.; Campbell, M.S.; Kiso, W.K.; Schmitt, D.L.; Waddell, P.J.; Bhaskara, S.; et al. Potential mechanisms for cancer resistance in elephants and comparative cellular response to DNA Damage in Humans. JAMA 2015, 314, 1850–1860. [Google Scholar] [CrossRef] [PubMed]
  13. Bredberg, A.; Schmitz, B. Human cancer, the naked mole rat and faunal turnovers. Cancer Med. 2019, 8, 1652–1654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Esposito, F.; de Martino, M.; Petti, M.G.; Forzati, F.; Tornincasa, M.; Federico, A.; Arra, C.; Pierantoni, G.M.; Fusco, A. HMGA1 pseudogenes as candidate proto-oncogenic competitive endogenous RNAs. Oncotarget 2014, 5, 8341–8354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Poliseno, L.; Marranci, A.; Pandolfi, P.P. Pseudogenes in human cancer. Front. Med. 2015, 2, 68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Pan, D.; Zhang, L. Burst of young retrogenes and independent retrogene formation in mammals. PLoS ONE 2009, 4. [Google Scholar] [CrossRef] [Green Version]
  17. Parker, H.G.; Von Holdt, B.M.; Quignon, P.; Margulies, E.H.; Shao, S.; Mosher, D.S.; Spady, T.C.; Elkahloun, A.; Cargill, M.; Jones, P.G.; et al. An expressed fgf4 retrogene is associated with breed-defining chondrodysplasia in domestic dogs. Science 2009, 325, 995–998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Toups, M.A.; Hahn, M.W. Retrogenes reveal the direction of sex-chromosome evolution in mosquitoes. Genetics 2010, 186, 763–766. [Google Scholar] [CrossRef] [Green Version]
  19. Szcześniak, M.W.; Ciomborowska, J.; Nowak, W.; Rogozin, I.B.; Makałowska, I. Primate and rodent specific intron gains and the origin of retrogenes with splice variants. Mol. Biol. Evol. 2011, 28, 33–37. [Google Scholar] [CrossRef]
  20. Richardson, S.R.; Salvador-Palomeque, C.; Faulkner, G.J. Diversity through duplication: Whole-genome sequencing reveals novel gene retrocopies in the human population. BioEssays 2014, 36, 475–481. [Google Scholar] [CrossRef] [Green Version]
  21. Magadum, S.; Banerjee, U.; Murugan, P.; Gangapur, D.; Ravikesavan, R. Gene duplication as a major force in evolution. J. Genet. 2013, 92, 155–161. [Google Scholar] [CrossRef]
  22. Kaessmann, H.; Vinckenbosch, N.; Long, M. RNA-based gene duplication: Mechanistic and evolutionary insights. Nat. Rev. Genet. 2009, 10, 19–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Chen, S.; Zhang, Y.E.; Long, M. New genes in Drosophila quickly become essential. Science 2010, 330, 1682–1685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Kubiak, M.R.; Szcześniak, M.W.; Makałowska, I. Complex analysis of retroposed genes’ contribution to human genome, proteome and transcriptome. Genes 2020, 11, 542. [Google Scholar] [CrossRef]
  25. Vinckenbosch, N.; Dupanloup, I.; Kaessmann, H. Evolutionary fate of retroposed gene copies in the human genome. Proc. Natl. Acad. Sci. USA 2006, 103, 3220–3225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Wen, Y.Z.; Zheng, L.L.; Liao, J.Y.; Wang, M.H.; Wei, Y.; Guo, X.M.; Qu, L.H.; Ayala, F.J.; Lun, Z.R. Pseudogene-derived small interference RNAs regulate gene expression in African Trypanosoma brucei. Proc. Natl. Acad. Sci. USA 2011, 108, 8345–8350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Wang, L.; Guo, Z.-Y.; Zhang, R.; Xin, B.; Chen, R.; Zhao, J.; Wang, T.; Wen, W.-H.; Jia, L.-T.; Yao, L.-B.; et al. Pseudogene OCT4-pg4 functions as a natural microRNA sponge to regulate OCT4 expression by competing for miR-145 in hepatocellular carcinoma. Carcinogenesis 2013, 34, 1773–1781. [Google Scholar] [CrossRef] [Green Version]
  28. Wood, A.J.; Schulz, R.; Woodfine, K.; Koltowska, K.; Beechey, C.V.; Peters, J.; Bourc’his, D.; Oakey, R.J. Regulation of alternative polyadenylation by genomic imprinting. Genes Dev. 2008, 22, 1141–1146. [Google Scholar] [CrossRef] [Green Version]
  29. Bryzghalov, O.; Szcześniak, M.W.; Makałowska, I. Retroposition as a source of antisense long non-coding RNAs with possible regulatory functions. Acta Biochim. Pol. 2016, 63, 825–833. [Google Scholar] [CrossRef]
  30. Baertsch, R.; Diekhans, M.; James, W.J.; Haussler, D.; Brosius, J. Retrocopy contributions to the evolution of the human genome. BMC Genomics 2008, 9, 466. [Google Scholar] [CrossRef] [Green Version]
  31. Tutar, Y. Pseudogenes. Comp. Funct. Genomics 2012. [Google Scholar] [CrossRef] [Green Version]
  32. Parmley, J.L.; Urrutia, A.O.; Potrzebowski, L.; Kaessmann, H.; Hurst, L.D. Splicing and the evolution of proteins in mammals. PLoS Biol. 2007, 5, 0343–0353. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, Y.W.; Liu, S.; Zhang, X.; Li, W.B.; Chen, Y.; Huang, X.; Sun, L.; Luo, W.; Netzer, W.J.; Threadgill, R.; et al. A Functional Mouse Retroposed Gene Rps23r1 Reduces Alzheimer’s β-Amyloid Levels and Tau Phosphorylation. Neuron 2009, 64, 328–340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Abyzov, A.; Iskow, R.; Gokcumen, O.; Radke, D.W.; Balasubramanian, S.; Pei, B.; Habegger, L.; Lee, C.; Gerstein, M. Analysis of variable retroduplications in human populations suggests coupling of retrotransposition to cell division. Genome Res. 2013, 23, 2042–2052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Kabza, M.; Kubiak, M.R.; Danek, A.; Rosikiewicz, W.; Deorowicz, S.; Polański, A.; Makałowska, I. Inter-population Differences in Retrogene Loss and Expression in Humans. PLoS Genet. 2015, 11, 1005579. [Google Scholar] [CrossRef] [PubMed]
  36. Marques, A.C.; Dupanloup, I.; Vinckenbosch, N.; Reymond, A.; Kaessmann, H. Emergence of young human genes after a burst of retroposition in primates. PLoS Biol. 2005, 3, 1970–1979. [Google Scholar] [CrossRef] [Green Version]
  37. Navarro, F.C.P.; Galante, P.A.F. A genome-wide landscape of retrocopies in primate genomes. Genome Biol. Evol. 2015, 7, 2265–2275. [Google Scholar] [CrossRef] [Green Version]
  38. Kozlov, A.P. Expression of evolutionarily novel genes in tumors. Infect. Agent. Cancer 2016, 11. [Google Scholar] [CrossRef] [Green Version]
  39. Seluanov, A.; Gladyshev, V.N.; Vijg, J.; Gorbunova, V. Mechanisms of cancer resistance in long-lived mammals. Nat. Rev. Cancer 2018, 18, 433–441. [Google Scholar] [CrossRef]
  40. McPherson, A.W.; Chan, F.C.; Shah, S.P. Observing clonal dynamics across spatiotemporal axes: A prelude to quantitative fitness models for cancer. Cold Spring Harb. Perspect. Med. 2018, 8. [Google Scholar] [CrossRef] [Green Version]
  41. Efeyan, A.; Serrano, M. p53: Guardian of the genome and policeman of the oncogenes. Cell Cycle 2007, 6, 1006–1010. [Google Scholar] [CrossRef]
  42. Olivier, M.; Hollstein, M.; Hainaut, P. TP53 mutations in human cancers: Origins, consequences, and clinical use. Cold Spring Harb. Perspect. Biol. 2010, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Giacomelli, A.O.; Yang, X.; Lintner, R.E.; McFarland, J.M.; Duby, M.; Kim, J.; Howard, T.P.; Takeda, D.Y.; Ly, S.H.; Kim, E.; et al. Mutational processes shape the landscape of TP53 mutations in human cancer. Nat. Genet. 2018, 50, 1381–1387. [Google Scholar] [CrossRef] [PubMed]
  44. Kitsoulis, C.V.; Baxevanis, A.D.; Abatzopoulos, T.J. The occurrence of cancer in vertebrates: A mini review. J. Biol. Res. 2020, 27, 1–12. [Google Scholar] [CrossRef]
  45. Seim, I.; Ma, S.; Zhou, X.; Gerashchenko, M.V.; Lee, S.G.; Suydam, R.; George, J.C.; Bickham, J.W.; Gladyshev, V.N. The transcriptome of the bowhead whale Balaena mysticetus reveals adaptations of the longest-lived mammal. Aging 2014, 6, 879–899. [Google Scholar] [CrossRef] [Green Version]
  46. Keane, M.; Semeiks, J.; Webb, A.E.; Li, Y.I.; Quesada, V.; Craig, T.; Madsen, L.B.; van Dam, S.; Brawand, D.; Marques, P.I.; et al. Insights into the evolution of longevity from the bowhead whale genome. Cell Rep. 2015, 10, 112–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Tian, X.; Azpurua, J.; Hine, C.; Vaidya, A.; Myakishev-Rempel, M.; Ablaeva, J.; Mao, Z.; Nevo, E.; Gorbunova, V.; Seluanov, A. High-molecular-mass hyaluronan mediates the cancer resistance of the naked mole rat. Nature 2013, 499, 346–349. [Google Scholar] [CrossRef] [Green Version]
  48. Tang, J.; Ning, R.; Zeng, B.; Li, Y. Molecular evolution of PTEN pseudogenes in mammals. PLoS ONE 2016, 11, 1–12. [Google Scholar] [CrossRef]
  49. Gorbunova, V.; Hine, C.; Tian, X.; Ablaeva, J.; Gudkov, A.V.; Nevo, E.; Seluanov, A. Cancer resistance in the blind mole rat is mediated by concerted necrotic cell death mechanism. Proc. Natl. Acad. Sci. USA 2012, 109, 19392–19396. [Google Scholar] [CrossRef] [Green Version]
  50. Fortunato, A.; Boddy, A.; Mallo, D.; Aktipis, A.; Maley, C.C.; Pepper, J.W. Natural selection in cancer biology: From molecular snowflakes to trait hallmarks. Cold Spring Harb. Perspect. Med. 2017, 7, 1–14. [Google Scholar] [CrossRef] [Green Version]
  51. Russo, M.; Bardelli, A. Lesion-Directed Therapies and Monitoring Tumor Evolution Using Liquid Biopsies. Cold Spring Harb. Perspect. Med. 2017, 1–14. [Google Scholar] [CrossRef] [Green Version]
  52. Enriquez-Navas, P.M.; Wojtkowiak, J.W.; Gatenby, R.A. Application of Evolutionary Principles to Cancer Therapy. Cancer Res. 2015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Venkatesan, S.; Swanton, C.; Taylor, B.S.; Costello, J.F. Treatment-Induced Mutagenesis and Selective Pressures Sculpt Cancer Evolution. Cold Spring Harb. Perspect. Med. 2017, 7. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, R.; Zhang, T.; Yang, Z.; Jiang, C.; Seng, J. Long non-coding RNA FTH1P3 activates paclitaxel resistance in breast cancer through miR-206/ABCB1. J. Cell. Mol. Med. 2018, 22, 4068–4075. [Google Scholar] [CrossRef] [PubMed]
  55. Fairbanks, D.J.; Fairbanks, A.D.; Heath Ogden, T.; Parker, G.J.; Maughan, P.J. NANOGP8: Evolution of a human-specific retro-oncogene. G3 2012, 2, 1447–1457. [Google Scholar] [CrossRef] [Green Version]
  56. Kong, Y.; Zhang, L.; Huang, Y.; He, T.; Zhang, L.; Zhao, X.; Zhou, X.; Zhou, D.; Yan, Y.; Zhou, J.; et al. Pseudogene PDIA3P1 promotes cell proliferation, migration and invasion, and suppresses apoptosis in hepatocellular carcinoma by regulating the p53 pathway. Cancer Lett. 2017, 407, 76–83. [Google Scholar] [CrossRef]
  57. Chen, X.; Wan, L.; Wang, W.; Xi, W.J.; Yang, A.G.; Wang, T. Re-recognition of pseudogenes: From molecular to clinical applications. Theranostics 2020, 10, 1479–1499. [Google Scholar] [CrossRef]
  58. Rosikiewicz, W.; Kabza, M.; Kosinski, J.G.; Ciomborowska-Basheer, J.; Kubiak, M.R.; Makalowska, I. RetrogeneDB-a database of plant and animal retrocopies. Database 2017, 2007, 1–11. [Google Scholar] [CrossRef]
  59. Poliseno, L.; Salmena, L.; Zhang, J.; Carver, B.; Haveman, W.J.; Pandolfi, P.P. A coding-independent function of gene and pseudogene mRNAs regulates tumour biology. Nature 2010, 465, 1033–1038. [Google Scholar] [CrossRef] [Green Version]
  60. Rohozinski, J.; Edwards, C.L.; Anderson, M.L. Does expression of the retrogene UTP14c in the ovary pre-dispose women to ovarian cancer? Med. Hypotheses 2012, 78, 446–449. [Google Scholar] [CrossRef]
  61. Chen, B.; Wang, C.; Zhang, J.; Zhou, Y.; Hu, W.; Guo, T. New insights into long noncoding RNAs and pseudogenes in prognosis of renal cell carcinoma. Cancer Cell Int. 2018, 18. [Google Scholar] [CrossRef]
  62. Wang, Q.S.; Shi, L.L.; Sun, F.; Zhang, Y.F.; Chen, R.W.; Yang, S.L.; Hu, J.L. High Expression of ANXA2 Pseudogene ANXA2P2 Promotes an Aggressive Phenotype in Hepatocellular Carcinoma. Dis. Markers 2019, 2019. [Google Scholar] [CrossRef] [PubMed]
  63. Xu, T.; Li, D.; He, Y.; Zhang, F.; Qiao, M.; Chen, Y. The expression level of CSDAP1 in lung cancer and its clinical significance. Oncol. Lett. 2018, 16, 4361–4366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Xu, H.; Chen, B.; Xing, J.; Wei, Z.; Liu, C.; Qiu, Y.; Lin, Y.; Ren, L. Upregulation of LGMNP1 confers radiotherapy resistance in glioblastoma. Oncol. Rep. 2019, 41, 3435–3443. [Google Scholar] [CrossRef] [PubMed]
  65. Lin, J.; Cao, S.; Wang, Y.; Hu, Y.; Liu, H.; Li, J.; Chen, J.; Li, P.; Liu, J.; Wang, Q.; et al. Long non-coding RNA UBE2CP3 enhances HCC cell secretion of VEGFA and promotes angiogenesis by activating ERK1/2/HIF-1α/VEGFA signalling in hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 2018, 37. [Google Scholar] [CrossRef]
  66. Wang, M.Y.; Chen, D.P.; Qi, B.; Li, M.Y.; Zhu, Y.Y.; Yin, W.J.; He, L.; Yu, Y.; Li, Z.Y.; Lin, L.; et al. Pseudogene RACGAP1P activates RACGAP1/Rho/ERK signalling axis as a competing endogenous RNA to promote hepatocellular carcinoma early recurrence. Cell Death Dis. 2019, 10. [Google Scholar] [CrossRef] [PubMed]
  67. Lou, W.; Ding, B.; Fan, W. High Expression of Pseudogene PTTG3P Indicates a Poor Prognosis in Human Breast Cancer. Mol. Ther. Oncolytics 2019, 14, 15–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Liu, Y.; Wang, W.; Li, Y.; Sun, F.; Lin, J.; Li, L. CKS1BP7, a Pseudogene of CKS1B, is Co-Amplified with IGF1R in Breast Cancers. Pathol. Oncol. Res. 2018, 24, 223–229. [Google Scholar] [CrossRef]
  69. Qian, Y.Y.; Li, K.; Liu, Q.Y.; Liu, Z.S. Long non-coding RNA PTENP1 interacts with miR-193a-3p to suppress cell migration and invasion through the PTEN pathway in hepatocellular carcinoma. Oncotarget 2017, 8, 107859–107869. [Google Scholar] [CrossRef] [Green Version]
  70. Guo, X.; Deng, L.; Deng, K.; Wang, H.; Shan, T.; Zhou, H.; Liang, Z.; Xia, J.; Li, C. Pseudogene PTENP1 Suppresses Gastric Cancer Progression by Modulating PTEN. AntiCancer Agents Med. Chem. 2016, 16, 456–464. [Google Scholar] [CrossRef]
  71. Yu, G.; Yao, W.; Gumireddy, K.; Li, A.; Wang, J.; Xiao, W.; Chen, K.; Xiao, H.; Li, H.; Tang, K.; et al. Pseudogene PTENP1 functions as a competing endogenous RNA to suppress clear-cell renal cell carcinoma progression. Mol. Cancer Ther. 2014, 13, 3086–3097. [Google Scholar] [CrossRef] [Green Version]
  72. Peng, H.; Ishida, M.; Li, L.; Saito, A.; Kamiya, A.; Hamilton, J.P.; Fu, R.; Olaru, A.V.; An, F.; Popescu, I.; et al. Pseudogene INTS6P1 regulates its cognate gene INTS6 through competitive binding of miR-17-5p in hepatocellular carcinoma. Oncotarget 2015, 6, 5666–5677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Rutnam, Z.J.; Du, W.W.; Yang, W.; Yang, X.; Yang, B.B. The pseudogene TUSC2P promotes TUSC2 function by binding multiple microRNAs. Nat. Commun. 2014, 5. [Google Scholar] [CrossRef] [PubMed]
  74. Li, M.; Sun, Q.; Wang, X. Transcriptional landscape of human cancers. Oncotarget 2017, 8, 34534–34551. [Google Scholar] [CrossRef] [Green Version]
  75. Chen, X.; Zhu, H.; Wu, X.; Xie, X.; Huang, G.; Xu, X.; Li, S.; Xing, C. Downregulated pseudogene CTNNAP1 promote tumor growth in human cancer by downregulating its cognate gene CTNNA1 expression. Oncotarget 2016, 7, 55518–55528. [Google Scholar] [CrossRef] [Green Version]
  76. Chen, W.; Niu, S.; Ma, X.; Zhang, P.; Gao, Y.; Fan, Y.; Pang, H.; Gong, H.; Shen, D.; Gu, L.; et al. RhoB acts as a tumor suppressor that inhibits malignancy of clear cell renal cell carcinoma. PLoS ONE 2016, 11, e0157599. [Google Scholar] [CrossRef] [PubMed]
  77. Mazieres, J.; Antonia, T.; Daste, G.; Muro-Cacho, C.; Berchery, D.; Tillement, V.; Pradines, A.; Sebti, S.; Favre, G. Loss of RhoB Expression in Human Lung Cancer Progression. Clin. Cancer Res. 2004, 10, 2742–2750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Liu, M.; Tang, Q.; Qiu, M.; Lang, N.; Li, M.; Zheng, Y.; Bi, F. MiR-21 targets the tumor suppressor RhoB and regulates proliferation, invasion and apoptosis in colorectal cancer cells. FEBS Lett. 2011, 585, 2998–3005. [Google Scholar] [CrossRef]
  79. Júnior, A.P.; de Sousa, V.P.L.; Esposito, F.; de Martino, M.; Forzati, F.; de Moreira, F.C.B.; de Simão, T.A.; Nasciutti, L.E.; Fusco, A.; Pinto, L.F.R.; et al. Overexpression of HMGA1 figures as a potential prognostic factor in endometrioid endometrial carcinoma (EEC). Genes 2019, 10, 372. [Google Scholar] [CrossRef] [Green Version]
  80. De Martino, M.; Forzati, F.; Arra, C.; Fusco, A.; Esposito, F. HMGA1-pseudogenes and cancer. Oncotarget 2016, 7, 28724–28735. [Google Scholar] [CrossRef]
  81. Zhou, Y.; He, P.; Xie, X.; Sun, C. Knockdown of SUMO1P3 represses tumor growth and invasion and enhances radiosensitivity in hepatocellular carcinoma. Mol. Cell. Biochem. 2019, 450, 125–134. [Google Scholar] [CrossRef]
  82. Mei, D.; Song, H.; Wang, K.; Lou, Y.; Sun, W.; Liu, Z.; Ding, X.; Guo, J. Up-regulation of SUMO1 pseudogene 3 (SUMO1P3) in gastric cancer and its clinical association. Med. Oncol. 2013, 30. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, L.-M.; Wang, P.; Liu, X.-M.; Zhang, Y.-J. LncRNA SUMO1P3 drives colon cancer growth, metastasis and angiogenesis. Am. J. Trans. Res. 2017, 9, 5461. [Google Scholar]
  84. Zhang, J.; Wang, X.; Chen, B.; Xiao, Z.; Li, W.; Lu, Y.; Dai, J. The human pluripotency gene NANOG/NANOGP8 is expressed in gastric cancer and associated with tumor development. Oncol. Lett. 2010, 1, 457–463. [Google Scholar] [CrossRef] [Green Version]
  85. Kawamura, N.; Nimura, K.; Nagano, H.; Yamaguchi, S.; Nonomura, N.; Kaneda, Y. CRISPR/Cas9-mediated gene knockout of NANOG and NANOGP8 decreases the malignant potential of prostate cancer cells. Oncotarget 2015, 6, 22361–22374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Bai, M.; Yuan, M.; Liao, H.; Chen, J.; Xie, B.; Yan, D.; Xi, X.; Xu, X.; Zhang, Z.; Feng, Y. OCT4 pseudogene 5 upregulates OCT4 expression to promote proliferation by competing with miR-145 in endometrial carcinoma. Oncol. Rep. 2015, 33, 1745–1752. [Google Scholar] [CrossRef] [Green Version]
  87. Ganapathi, M.K.; Jones, W.D.; Sehouli, J.; Michener, C.M.; Braicu, I.E.; Norris, E.J.; Biscotti, C.V.; Vaziri, S.A.J.; Ganapathi, R.N. Expression profile of COL2A1 and the pseudogene SLC6A10P predicts tumor recurrence in high-grade serous ovarian cancer. Int. J. Cancer 2016, 138, 679–688. [Google Scholar] [CrossRef] [PubMed]
  88. Pan, Y.; Sun, C.; Huang, M.; Liu, Y.; Qi, F.; Liu, L.; Wen, J.; Liu, J.; Xie, K.; Ma, H.; et al. A genetic variant in pseudogene E2F3P1 contributes to prognosis of hepatocellular carcinoma. J. Biomed. Res. 2014, 28, 194–200. [Google Scholar] [CrossRef] [Green Version]
  89. Marciniak, P. Retropozycja Genów w Tkankach Nowotworowych. Master’s Thesis, Adam Mickiewicz University, Poznan, Poland, 2018. [Google Scholar]
  90. Huang, M.; Prendergast, G.C. RhoB in cancer suppression. Histol. Histopathol. 2006, 21, 213–218. [Google Scholar] [CrossRef]
  91. Yamulla, R.J.; Nalubola, S.; Flesken-Nikitin, A.; Nikitin, A.Y.; Schimenti, J.C. Most Commonly Mutated Genes in High-Grade Serous Ovarian Carcinoma Are Nonessential for Ovarian Surface Epithelial Stem Cell Transformation. Cell Rep. 2020, 32. [Google Scholar] [CrossRef]
  92. Shao, Y.; Chen, C.; Shen, H.; He, B.Z.; Yu, D.; Jiang, S.; Zhao, S.; Gao, Z.; Zhu, Z.; Chen, X.; et al. GenTree, an integrated resource for analyzing the evolution and function of primate-specific coding genes. Genome Res. 2019, 29, 682–696. [Google Scholar] [CrossRef]
  93. Goemann, I.M.; Marczyk, V.R.; Recamonde-Mendoza, M.; Wajner, S.M.; Graudenz, M.S.; Maia, A.L. Decreased expression of the thyroid hormone-inactivating enzyme type 3 deiodinase is associated with lower survival rates in breast cancer. Sci. Rep. 2020, 10, 13914. [Google Scholar] [CrossRef] [PubMed]
  94. Jan, Y.H.; Tsai, H.Y.; Yang, C.J.; Huang, M.S.; Yang, Y.F.; Lai, T.C.; Lee, C.H.; Jeng, Y.M.; Huang, C.Y.; Su, J.L.; et al. Adenylate kinase-4 is a marker of poor clinical outcomes that promotes metastasis of lung cancer by downregulating the transcription factor ATF3. Cancer Res. 2012, 72, 5119–5129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Han, M.Z.; Bin, H.; Chen, A.J.; Xin, Z.; Xu, R.; Jian, W.; Li, X.G. High expression of RAB43 predicts poor prognosis and is associated with epithelial-mesenchymal transition in gliomas. Oncol. Rep. 2017, 37, 903–912. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Guimaraes, J.C.; Zavolan, M. Patterns of ribosomal protein expression specify normal and malignant human cells. Genome Biol. 2016, 17, 236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Cai, J.J.; Petrov, D.A. Relaxed purifying selection and possibly high rate of adaptation in primate lineage-specific genes. Genome Biol. Evol. 2010, 2, 393–409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Matyunina, E.A.; Emelyanov, A.V.; Kurbatova, T.V.; Makashov, A.A.; Mizgirev, I.V.; Kozlov, A.P. Evolutionarily novel genes are expressed in transgenic fish tumors and their orthologs are involved in development of progressive traits in humans. Infect. Agent. Cancer 2019, 14, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. GTEx Portal. Available online: https://gtexportal.org/home/ (accessed on 14 December 2020).
Figure 1. Schematic view of the relationship between retrogenes, cancer, and evolution.
Figure 1. Schematic view of the relationship between retrogenes, cancer, and evolution.
Life 11 00072 g001
Figure 2. Mechanism of retrocopy formation (a) and possible functions in the cell (b) (based on mechanisms collectively described by Kubiak et al. [3]).
Figure 2. Mechanism of retrocopy formation (a) and possible functions in the cell (b) (based on mechanisms collectively described by Kubiak et al. [3]).
Life 11 00072 g002
Figure 3. Evolutionary cancer mechanisms in the context of species-specific features in cancer suppression (a), clonal subpopulations within a tumor (b), and cancer treatment (c).
Figure 3. Evolutionary cancer mechanisms in the context of species-specific features in cancer suppression (a), clonal subpopulations within a tumor (b), and cancer treatment (c).
Life 11 00072 g003
Figure 4. Cancer-related retrocopies as miRNA sponges in normal and cancer cells. Binding miRNAs to the highly expressed retrogene and translation of the suppressor gene (a). Degradation of the oncogene mRNA by miRNAs binding due to a low retrogene expression (b). Degradation of the suppressor gene mRNA by miRNAs binding because of low retrogene level (c). Binding miRNAs to the high expressed retrogene and translation of the oncogene (d).
Figure 4. Cancer-related retrocopies as miRNA sponges in normal and cancer cells. Binding miRNAs to the highly expressed retrogene and translation of the suppressor gene (a). Degradation of the oncogene mRNA by miRNAs binding due to a low retrogene expression (b). Degradation of the suppressor gene mRNA by miRNAs binding because of low retrogene level (c). Binding miRNAs to the high expressed retrogene and translation of the oncogene (d).
Life 11 00072 g004
Figure 5. Schematic tree illustrating the estimated time of the origin of cancer-related retrocopies during animal evolution.
Figure 5. Schematic tree illustrating the estimated time of the origin of cancer-related retrocopies during animal evolution.
Life 11 00072 g005
Table 1. Characteristics of the cancer-related retrogenes that have been described in the literature (based on [15,57]).
Table 1. Characteristics of the cancer-related retrogenes that have been described in the literature (based on [15,57]).
RetrocopyEnsembl IDRetrogeneDB ID [58]ChromosomeParental GeneCancer Type
KRASP1ENSG00000220635retro_hsap_34746KRASprostate cancer [59]
UTP14CENSG00000253797retro_hsap_2913UTP14Aovarian cancer [60]
MSL3P1ENSG00000224287retro_hsap_24012MSL3renal cell carcinoma [61]
ANXA2P2ENSG00000231991retro_hsap_41509ANXA2hepatocellular carcinoma [62]
CSDAP1 (YBX3P1)ENSG00000261614retro_hsap_167416YBX3lung adenocarcinoma [63]
LGMNP1ENSG00000214269retro_hsap_127213LGMNglioblastoma [64]
UBE2CP3ENSG00000250384retro_hsap_29354UBE2Chepatocellular carcinoma [65]
RACGAP1PENSG00000257331-12RACGAP1hepatocellular carcinoma [66]
PTTG3PENSG00000213005-8PTTG1breast cancer [67]
CKS1BP7ENSG00000254331-8CKS1Bbreast cancer [68]
PTENP1ENSG00000237984retro_hsap_42459 PTENhepatocellular carcinoma [69], gastric cancer [70], renal cell carcinoma [71]
INTS6P1ENSG00000250492retro_hsap_33075INTS6hepatocellular carcinoma [72]
TUSC2P1ENSG00000285470-YTUSC2esophageal squamous cell carcinoma [73]
NKAPLENSG00000189134retro_hsap_156NKAPkidney renal papillary cell carcinoma, pancreatic adenocarcinoma, adenoid cystic carcinoma [74]
CTNNA1P1ENSG00000249026-5CTNNA1colorectal cancer [75]
RHOBENSG00000143878retro_hsap_1082RHOArenal cell carcinoma [76], lung cancer [77], colorectal cancer [78]
HMGA1P6ENSG00000233440retro_hsap_117513HMGA1endometrial carcinoma [79], ovarian carcinosarcoma, thyroid carcinoma [14]
HMGA1P7ENSG00000216753-6HMGA1endometrial carcinoma [79], ovarian carcinosarcoma, thyroid carcinoma [14], breast cancer [80]
SUMO1P3ENSG00000235082retro_hsap_2401SUMO1hepatocellular carcinoma [81], gastric cancer [82], colorectal cancer [83]
NANOGP8ENSG00000255192retro_hsap_154915NANOGgastric cancer [84], prostate cancer [85]
POU5F1P4 (OCT4-pg4)ENSG00000237872-1POU5F1hepatocellular carcinoma [27]
POU5F1P5 (OCT4-pg5)ENSG00000236375-10POU5F1endometrial carcinoma [86]
SLC6A6P1ENSG00000226818retro_hsap_249821SLC6A6ovarian cancer [87]
PDIA3P1ENSG00000180867retro_hsap_2171PDIA3multiple myeloma [56]
PPIAP43ENSG00000255059retro_hsap_81611PPIAsmall cell lung cancer [10]
FTH1P3ENSG00000213453retro_hsap_22402FTH1breast cancer [54]
E2F3P1ENSG00000267046retro_hsap_174917E2F3hepatocellular carcinoma [88]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Staszak, K.; Makałowska, I. Cancer, Retrogenes, and Evolution. Life 2021, 11, 72. https://0-doi-org.brum.beds.ac.uk/10.3390/life11010072

AMA Style

Staszak K, Makałowska I. Cancer, Retrogenes, and Evolution. Life. 2021; 11(1):72. https://0-doi-org.brum.beds.ac.uk/10.3390/life11010072

Chicago/Turabian Style

Staszak, Klaudia, and Izabela Makałowska. 2021. "Cancer, Retrogenes, and Evolution" Life 11, no. 1: 72. https://0-doi-org.brum.beds.ac.uk/10.3390/life11010072

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