Next Article in Journal
Sestrin2 Expression Has Regulatory Properties and Prognostic Value in Lung Cancer
Next Article in Special Issue
A Genotyping/Phenotyping Approach with Careful Clinical Monitoring to Manage the Fluoropyrimidines-Based Therapy: Clinical Cases and Systematic Review of the Literature
Previous Article in Journal
Efficacy of Incremental Next-Generation ALK Inhibitor Treatment in Oncogene-Addicted, ALK-Positive, TP53-Mutant NSCLC
Previous Article in Special Issue
Diagnostic and Prognostic Significance of MiR-150 in Colorectal Cancer: A Systematic Review and Meta-Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pharmacogenomics to Predict Tumor Therapy Response: A Focus on ATP-Binding Cassette Transporters and Cytochromes P450

1
Toxicogenomics Unit, National Institute of Public Health, 100 00 Prague, Czech Republic
2
Laboratory of Pharmacogenomics, Biomedical Center, Faculty of Medicine in Pilsen, Charles University, 306 05 Pilsen, Czech Republic
3
Third Faculty of Medicine, Charles University, 100 00 Prague, Czech Republic
*
Author to whom correspondence should be addressed.
J. Pers. Med. 2020, 10(3), 108; https://doi.org/10.3390/jpm10030108
Submission received: 31 July 2020 / Revised: 21 August 2020 / Accepted: 26 August 2020 / Published: 28 August 2020
(This article belongs to the Special Issue Pharmacogenomics of Oncology Therapies)

Abstract

:
Pharmacogenomics is an evolving tool of precision medicine. Recently, due to the introduction of next-generation sequencing and projects generating “Big Data”, a plethora of new genetic variants in pharmacogenes have been discovered. Cancer resistance is a major complication often preventing successful anticancer treatments. Pharmacogenomics of both somatic mutations in tumor cells and germline variants may help optimize targeted treatments and improve the response to conventional oncological therapy. In addition, integrative approaches combining copy number variations and long noncoding RNA profiling with germline and somatic variations seem to be a promising approach as well. In pharmacology, expression and enzyme activity are traditionally the more studied aspects of ATP-binding cassette transporters and cytochromes P450. In this review, we briefly introduce the field of pharmacogenomics and the advancements driven by next-generation sequencing and outline the possible roles of genetic variation in the two large pharmacogene superfamilies. Although the evidence needs further substantiation, somatic and copy number variants as well as rare variants and common polymorphisms in these genes could all affect response to cancer therapy. Regulation by long noncoding RNAs has also been shown to play a role. However, in all these areas, more comprehensive studies on larger sets of patients are needed.

1. Introduction

Pharmacogenetics is not a novel term. It was first used in 1959 by a German geneticist Friedrich Vogel. The term pharmacogenetics was understood as “phenotypic variation in metabolism and response to drugs” [1]. Later on, advances in analytical methods and gene cloning resulted in an improved understanding of the genetic basis of this variation. The cloning and sequencing of the entire human genome in the late 1990s brought with them another term—pharmacogenomics. These two terms are now used interchangeably, although the term pharmacogenomics is broader, also comprising the development of new drugs targeting specific disease-causing genes [2].
Pharmacogenetics has always been an important part of personalized therapy. The absorption, distribution, metabolism, and excretion (ADME) of drugs, pharmacokinetics (PK) and pharmacodynamics (PD) affect drug efficacy and toxicity and can thereby cause adverse drug reactions (ADRs) or therapy failure. All of these factors are influenced by inter-individual variations largely dictated by genotype, and thus pharmacogenetics has also become important in the clinical setting [3].
Initially, only hereditary germline variation was the primary focus of pharmacogenetics. However, after breakthroughs in genotyping techniques and sequencing, such as capillary sequencing, cancer genetics/genomics became a viable direction for research [4]. Tumor somatic mutations were revealed to be an important aspect of cancer treatment, allowing for the design of a number of new drugs targeting specific gene mutations [5]. After commercial use of next-generation sequencing (NGS) began in 2005, many new variants were discovered in known pharmacogenes. Pharmacogenetics/pharmacogenomics rapidly expanded as new “Big Data” sequencing projects (e.g., 1000 Genomes Project) were established. It was soon evident that the number of variants in pharmacogenes would be much greater than previously thought. For example, the number of variants in the cytochrome P450 (CYP) gene superfamily in the European population was calculated to be 3.4 × 104 [6]. Rare variants (minor allele frequency (MAF) < 1%) are also an important part of pharmacogenetics and constitute the majority of variant alleles in pharmacogenes. Recent studies on Exome Aggregation Consortium (ExAC) and the 1000 Genomes Project corroborate the importance of rare genetic variants in the pharmacogenetic prediction of drug response [7,8].
The number of publications about pharmacogenetics/pharmacogenomics is growing rapidly. While in 2000 only 57 articles published that year in the PubMed database [9] mentioned the term “pharmacogen* AND cancer“—this number rose to 365 in 2010 and almost 700 articles came out in 2019. Therefore, summarizing all known pharmacogenes that could be of interest in cancer therapy is beyond the scope of this review. Advancements in personalized therapy and a comparison between the germline and somatic aspects of pharmacogenetics were summarized by Hertz and McLeod [10], McLeod [11] and recently by Hyman et al. [5].
Common pharmacogenes predominantly consist of phase I and phase II biotransformation enzymes, cytochrome reductases, membrane transporters, and nuclear receptors. The genetic variability in these genes can modulate ADME and may cause ADRs or changes in drug efficacy [8]. As a result, exposure of cancer cells to a lower therapeutic dose can lead to resistance to therapy. Cancer chemoresistance and suboptimal responses to therapy have been a major obstacle for successful cancer treatments and can stem from a number of factors.
ATP-binding cassette (ABC) transporters may induce drug resistance by transporting drugs across cellular membranes. CYP monooxygenases metabolize a broad range of drugs or activate prodrugs. Together, these groups represent a significant portion of known pharmacogenes. The concept of genetic variants in ABCs and CYPs affecting the response of a cancer cell to an anticancer drug (e.g., paclitaxel) is presented in Figure 1. This article will focus solely on the genomic variation in ABC transporters and CYPs and their possible roles in tumor therapy responses and will outline the most recent findings, summarizing the perspectives of their potential use in personalized therapies on germline, somatic and copy number levels. In addition to genomics, an emerging direction of studying noncoding regulatory molecules will be discussed.
The cartoon displays the roles of ABC and CYP pharmacogenomics in anticancer treatments. Paclitaxel serves as an example. P450 variant CYP2C8*3 lowers the activity of the enzyme, resulting in more paclitaxel in the nucleus with a higher ability to cause mitotic catastrophe. CYP2C8*3 was found to be associated with an increased risk of taxane-related sensory neuropathy [12]. Simultaneously, paclitaxel might be the subject of efflux outside of the cell, e.g., by p-glycoproteins coded by ABCB1, lowering its therapeutic efficacy. Variant rs3213619 could mean a higher paclitaxel efflux and worse predicted prognosis. On the other hand, variant rs3213619 was associated with a decreased risk of taxane-related sensory neuropathy [13].

2. Germline Variants in ABC Transporters and CYP Genes

Germline single nucleotide variants in CYPs may modulate the metabolic activity in the liver and thus affect the prodrug activation or detoxification of active drugs in the organism. Secondly, hereditary genetic variability in membrane efflux transporters may alter the effective concentration of drugs in tumor cells. Furthermore, heterozygotic alterations may couple with a loss in the heterozygosity of the second allele in tumor cells or epigenetic dysregulation and impair cellular transport. Thus, assessing hereditary genetic information before prescribing anticancer drugs might be beneficial for patients’ life expectancy and quality of life and for health care system sustainability.
According to the U.S. Food and Drug Administration (FDA) Table of Pharmacogenomic Biomarkers in Drug Labeling [14], several diagnostic tests are mandatory before the prescription of anticancer drugs. Most of these drugs are tyrosine kinase inhibitors (TKIs) [15]. Focusing on germline variants, among the anticancer drugs found in drug labels are the EGFR inhibitor gefitinib, FGRF inhibitor erdafitinib, PARP inhibitor rucaparib, and tamoxifen, a selective modulator of the estrogen receptor [14]. European Medicines Agency (EMA) has a drug label also for the TKI sunitinib [16], but it is only related to drug–drug interactions with CYP3A4 (avoidance of CYP3A4 inhibitors), not its genetics. FDA drug labels for these compounds are described in Table 1. In the FDA drug labels, no dose recommendations are applicable but “The exposure of erdafitinib is predicted to be 50% higher in subjects with the CYP2C9*3/*3 genotype, estimated to be present in 0.4% to 3% of the population among various ethnic groups”. The rest is limited to the recommendations for the monitoring of ADR in patients with a poor metabolizer genotype (gefitinib) or information about metabolizer phenotypes (rucaparib and tamoxifen).
In the Pharmacogenomics Knowledgebase [17] multiple members of ABC and CYP superfamilies are indicated as “Very Important Pharmacogenes” (VIP) [18]. These are: ABCB1, ABCG2, CFTR, CYP1A2, CYP2A6, CYP2B6, CYP2C8, CYP2C9, CYP2C19, CYP2D6, CYP2E1, CYP2J2, CYP3A4, CYP3A5, and CYP4F2. However, most of the clinical associations are not related to anticancer drugs. For these, a high level of evidence is fulfilled only by tamoxifen dosage adjustment for CYP2D6 haplotypes (CYP2D6*1, CYP2D6*10, CYP2D6*3, CYP2D6*4, CYP2D6*41, CYP2D6*5, and CYP2D6*6) in breast cancer. Patients with the CYP2D6*1/*1 genotype, i.e., normal metabolizers, have an increased metabolism of tamoxifen. These patients show increased concentrations of tamoxifen’s active metabolite endoxifen, resulting in decreased cancer recurrence and improved disease-free survival during adjuvant tamoxifen treatment. Poor metabolizers carrying no functional alleles are recommended to use an alternate hormonal therapy, such as aromatase inhibitor anastrozole [19,20,21,22]. A moderate level of evidence (i.e., results are replicated, but some studies are not significant, and/or the effect size may be small) exists also for the variant rs3892097 in CYP2D6 (CYP2D6*4). Patients treated with tamoxifen who have the AA genotype, i.e., poor metabolizers, show an increased risk of relapse compared with patients with the GA or GG genotypes [18]. Finally, a well described variant rs4646 in the aromatase gene CYP19A1 is associated with the efficacy of anastrozole, letrozole, or tamoxifen in breast cancer patients [23,24]. Other studies concerning therapy efficacy or metabolism/PK which are described in the section of clinical annotations in PharmGKB have low (nonreplicated study or multiple studies lacking clear evidence of an association) or preliminary (in vitro, nonsignificant study or case-report) levels of evidence. Studies with levels of evidence from 1 to 3 (high–low) are summarized in Table 2. The studies with low levels of evidence will not be discussed further for the sake of brevity.
The role of rare germline variants in pharmacogenetics was recently underlined by Kozyra et al. [7]. Their finding was further confirmed by studies on ExAC data. Rare genetic variants constituted the vast majority of variants in pharmacogenes (98.5% out of 69,923 variants in 208 pharmacogenes or 97.5% out of 61,134 functional variants in 806 pharmacogenes) [8,25]. Rare pharmacogenetic variants were also highly enriched in those that were predicted to be functional alterations. In addition, rare variants were abundant in the genes of the irinotecan (topoisomerase inhibitor prodrug) pathway, accounting for 100% of variants in ABCB1, ABCC2, and ABCG2. Rare variants also represented more than 40% of variants in ABCC1 and CYP3A4 and to a lesser extent in CYP3A5 [8]. Quite recently, based on the data from The Cancer Genome Atlas (TCGA), it was discovered that a variant burden in ABCC1 substantially predicts disease-free survival in breast cancer patients and that this significance was even stronger in cyclophosphamide- and doxorubicin-treated subgroups [26]. Therefore, rare variants account for a substantial part of drug metabolism and may become an important factor in genotyping-based predictions of drug response.
In summary, germline variants in many of the VIPs of CYP and ABC superfamilies as well as in other genes (ABCB4, ABCC1-C4, CYP1A1, and CYP1B1) are associated with drug response in cancer. A recent study brought evidence that the germline genetic component was at least as important in the prediction of drug sensitivity as somatic mutations in 993 cell lines treated with 265 drugs [27]. Germline variants are thus a relevant part of pharmacogenetic studies in cancer therapy.
Table 2. Pharmacogenomics Knowledgebase (PharmGKB) clinical annotations of ABC and CYP genes.
Table 2. Pharmacogenomics Knowledgebase (PharmGKB) clinical annotations of ABC and CYP genes.
Gene 1VariantDrugTypeDiseaseReference
ABCB1rs10276036, rs2032582, rs1128503, rs1045642, rs2229109, rs4148737platinum compounds, cyclophosphamide, anthracyclines, methotrexate, vincristine, paclitaxel, sunitinib, imatinib, cytarabine, lenalidomide, 5-FU, tamoxifenefficacyOSRC, BC, OVC, RCC, AML, multiple myeloma, CML, NSCLC, CRC, EC[28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54]
ABCB4rs1202283imatinibefficacyGIST[55]
ABCC1rs6498588irinotecanPKCRC[56]
ABCC2rs2273697, rs3740065Imatinib, tamoxifenefficacyGIST[54,55,57]
ABCC3rs4148416, rs9895420cisplatin, methotrexateefficacyOSRC, ALL[28,58]
ABCC4rs9561765, rs16950650imatinib, cisplatinefficacyGIST, SCLC[59,60]
ABCG2rs2231142, rs2725252, rs7699188, rs13120400, rs12505410capecitabine, 5-FU, leukovorin, oxaliplatin, imatinib, gemcitabine, irinotecanefficacyCRC, AML, NSCLC, CML[61,62,63,64]
CYP1A1rs1048943capecitabine, docetaxelefficacyBC[65]
CYP1A2rs762551imatinibdosageGIST[66]
CYP1B1rs10568365-FU, anthracycline, cyclophosphamideefficacyBC[67]
CYP2A6*1A, *4AtegafurPKhealthy liver[68]
CYP2B6rs12721655, rs3745274, *1, *6cyclophosphamide, doxorubicin, imatinibefficacyBC, CML, CLL[43,69,70]
CYP2C19rs4244285, *1/ *17/*2cyclophosphamide, doxorubicin, tamoxifenefficacyBC[43,71]
CYP2D6*1/*10/*3/*4/*41/*5/*6 2, rs3892097 3, *1/*10/*114/*2/*3/*4/*41/*4xN/*5/*6tamoxifen, gefitinibefficacy, PKNSCLC, BC[19,20,21,22,72,73,74,75,76,77,78]
CYP2E1rs6413432, rs2070676cisplatin, cyclophosphamideefficacyOVC[79]
CYP3A4rs12721627paclitaxeldosageNSCLC[80]
CYP3A5rs776746imatinibPKCML[50]
CYP19A1rs4646 3anastrozole, letrozole, tamoxifenefficacyBC[23,24]
1 Very important pharmacogenes (VIP) are depicted in bold. 2 Level 1A association. 3 Level 2A/B associations (The rest of the associations are level 3, i.e., low level of evidence). Abbreviations: 5-FU—5-fluorouracil; ALL—acute lymphoblastic leukemia; AML—acute myeloid leukemia; BC—breast cancer; CLL—chronic lymphocytic leukemia; CML—chronic myeloid leukemia; CRC—colorectal cancer; EC—esophageal cancer; GIST—gastrointestinal stromal tumor; NSCLC—non-small cell lung cancer; OSRC—osteogenic sarcoma; OVC—ovarian cancer; PK—pharmacokinetics; RCC—renal cell carcinoma; SCLC—small cell lung cancer.

3. Somatic Variants in ABCs and CYPs in Solid Tumors

As outlined above, a mutation in a gene coding for a protein that metabolizes/transports a therapeutic drug could, in principle, affect the function of the protein, and thus lead to a change in the response of the cancer cell to treatment. However, while the area of germline variation and gene expression of ABC transporters and CYPs has been a major research focus for decades, only recently have somatic variations been given similar attention, owing to the advancements in knowledge and technology. The level of evidence (or lack thereof) is in stark contrast here.
While a number of pharmacogenes have been approved by the FDA as biomarkers and are used in drug labeling [14], none of the genes are ABC transporters and only three CYPs are on the list. More importantly, only the germline variation in population is considered, not variation due to somatic mutations in tumors, despite the reasonable expectation that it is indeed somatic, rather than germline, variation acquired by cancer cells that is in most cases behind the poor response of a tumor to therapy.
In TCGA (data release 23.0) [81], which collects data from dozens of large-scale genomic projects and currently contains mutation data from more than 83 thousand cases, including 67 primary tumor sites, there are in total 17,329 mutations indexed for 48 functional protein-coding ABC transporter genes and 9789 mutations for 57 CYPs. These variants can be further filtered by their Variant Effect Predictor (VEP), Sorting Intolerant From Tolerant (SIFT) of Polymorphism Phenotyping (PolyPhen) scores, representing their probable impact on the phenotype. For example, in ABC transporters, 1506 mutations in 48 genes are classified as having high impact according to the VEP score, meaning they cause a frameshift, stop gain, stop loss, start loss or a splice alteration. Hypothetically, any one of these variants could significantly affect the resulting protein and thus modulate its function. Since many ABCs are known to transport drugs out of cells and lower their efficacy, damaging the function of these proteins could lead to a higher sensitivity to therapeutics [82]. The function of CYPs, well-known drug metabolizers (794 high impact mutations in 57 genes) could be possibly altered in an analogous fashion. However, practically no high-level experimental studies exist that clearly confirm or disprove any potential roles of these variants in cancer therapy. The top 30 mutated (any predicted impact, incl. missense variants) ABC transporters (Figure 2) and CYPs (Figure 3) in TGCA in the top 200 mutated cases across the four most common cancers (colorectal, lung and bronchus, breast, and prostate) are presented for an overview.
A similar situation exists in the Catalogue of Somatic Mutations in Cancer (COSMIC) database (v91), a massive database that consists of more than 6 million coding mutations measured in 1.4 million tumor samples (including those in TCGA), manually curated from over 26 thousand publications [83]. In their list of genes whose somatic mutations have been implicated in drug resistance, no ABC transporters or CYPs are listed. Interestingly, CYP2C8 is included in their Cancer Gene Census [84], a list of genes which contain mutations causally implicated in cancer, with the level of evidence (Tier 2) described as “strong indications of a role in cancer, but with less extensive available evidence (than Tier 1)”. This finding was based on a 2018 study suggesting that CYP2C8 is an active driver of cancer after panel sequencing tumor samples from 273 patients with colorectal cancer (CRC) and an analysis of the somatic variants [85].
Mucinous CRC is known for a poorer response to chemotherapy and a worse prognosis than other types of CRC, owing largely to a higher rate of resistance [86]. Reynolds et al. [87] recently analyzed TCGA data for 67 mucinous CRC and compared them to 456 samples of non-mucinous CRC. They looked for associations of 26 pharmacogenes with response to irinotecan, oxaliplatin and 5-fluorouracil (5-FU). The somatic mutation rate between the two cohorts was compared and two genes were identified as having statistically significantly different rates: ABCB1 (p = 0.042), and ABCG2 (p = 0.010). The mutation rate of these two genes associated with resistance to irinotecan. The mutations did not lead to a statistically significant difference in the expression of corresponding proteins; however, the authors postulated that these mutations result in overactive proteins that more effectively export irinotecan out of the cell, lowering its efficacy [87].
While the ABC and CYP superfamilies are well established in pharmacology and oncology as influencers of drug responses, we are yet to see the vast data on their somatic genetic variation in tumors being utilized in a fashion similar to how gene expression and germline variation data are being used already in the clinic. This disparity is striking, considering the number of mutations in these important pharmacogenes that arise in tumors.

4. Copy Number Variants in ABCs and CYPs

Structural variation, including alterations in gene copy numbers, is a key mutational process in cancer development and progression [88,89]. In addition to their potential predictive value in terms of cancer onset, copy number variations (CNVs) can also help point out tumors with a poor prognosis and even provide therapeutically useful information [90].
The fact that germline CNVs are frequently observed in pharmacogenes (PharmGKB) was recently demonstrated by a retrospective study on data from the 1000 Genomes Project and ExAC. CNVs that are present in 97% of common pharmacogenes accounted for 5% of all the loss-of-function alleles in 42% of the studied pharmacogenes. Novel deletions were present with particularly high MAF in CYP2C19 and CYP4F2 [91]. The authors recommend the utilization of CNV detection assays for the testing of relevant genes in appropriate populations. Therefore, CNVs are another important source of drug response variation.
In oncology, CNVs in CYP2D6 are among the most frequently studied structural variations in connection with efficacy and ADR of tamoxifen treatment in breast cancer patients. A recent study reported that about 26% of Ethiopian breast cancer patients had germline CYP2D6 amplifications, predicting a significantly increased plasma concentration of endoxifen, the active tamoxifen metabolite, in most cases [92]. This study confirmed previous recommendations of the Clinical Pharmacogenetics Implementation Consortium (CPIC) suggesting that patients with gene amplifications should avoid moderate or strong CYP2D6 inhibitors [20], but a population-specific context warrants caution as even within Europe the inter-population variability in CYP2D6 genetics is high [93]. Despite notable progress in this area, some technical challenges persist, e.g., problems with an accurate assignment of the amplified allele in CNV heterozygotes [94,95] which have to be carefully standardized which includes the increasingly popular massive parallel sequencing methods [96]. In addition to the previously mentioned importance of germline single nucleotide polymorphisms in CYP19A1 for the treatment of breast cancer patients with aromatase inhibitors, a recent study observed that more than 20% of patients treated with inhibitors acquired CYP19A1 amplifications [97], suggesting a counter-attack of tumor cells against this therapy in a significant portion of patients and a need for further optimization.
In parallel with the FDA and EMA recommendations on the testing of germline polymorphisms, CNVs in CYP2A6, CYP2D6 and CYP2E1 (and GSTM1 or GSTT1) were found overrepresented among good responders to therapy with the TKIs imatinib, dasatinib or nilotinib in a small-scale study of patients with chronic myeloid leukemia [98]. However, this potentially useful observation for the optimization of targeted therapy awaits replication.
Among drug efflux transporters, few reports about the role of their CNVs in cancer are available. Recently reviewed information [99] suggested that an ABCB1 locus including another transporter ABCB4 is frequently amplified in cancer and mainly in tumors or in vitro models with induced drug resistance [100]. This could present an additional opportunity for intervention (in addition to clinically tested inhibitors) against hard to treat tumors. Outside of numerous studies on ABCB1 in drug resistance, ABCC1 was also found to be amplified in drug resistant in vitro models [101] and may be considered a target for future studies in this area.
A study of 128 discordant sibling pairs, validated in 1048 Chinese Han subjects, revealed an association of CNVs at 13q32.1, where ABCC4 is located, with an increased risk of esophageal squamous cell carcinoma [102], suggesting that CNVs should be investigated also in less studied transporters. The relevance of this observation to cancer progression or therapy awaits further elaboration.
The data paucity can be demonstrated by the analysis of publicly available databases (Table 3), where large germline CNVs (larger than 1 kb) can be identified only in CYP2D6, CYP4A11, CYP17A1, ABCD3, and ABCE1 (the Genome Aggregation Database—gnomAD [103]). None of these genes are either recognized as disease causing (in Database of Chromosomal Imbalance and Phenotype in Humans Using Ensembl Resources —DECIPHER [104]) or have pathogenic large CNVs of clinical significance in the ClinVar database [105]. Moreover, all germline CNVs, except MCNV_22_1026 in CYP2D6, are rather rare CNVs, with frequencies smaller than one allele per thousand (Table 3). Complementary analysis of somatic CNVs using the COSMIC database v91 (all databases accessed on 5 May 2020) demonstrates that highest counts of CNVs among all available cancers were observed in the less studied genes CYP11B1, CYP11B2, CYP4V2, and ABCA5-A12. In contrast to the pharmacogenes presented in Table 1, somatic CNVs were listed only in CYP2C9 at approximately half the frequency compared to CYP11B1/2. The same basically applies to ABC transporters listed in Table 2, except for ABCB1, ABCB4 and ABCC2. However, unlike the CYPs, the length of individual ABC genes markedly differs within the superfamily (e.g., the length of ABCC2 is 70kb while the length of ABCB1 is 210kb), a fact that calls for cautious interpretation in this particular case.
From the above information, it is obvious that CNVs represent valuable information in addition to single nucleotide polymorphisms and rare variants, and, after a thorough standardization of methodology for their assessment and interpretation, should supplement a predictive, diagnostic, and prognostic exploitation of gene panels in oncology and therapeutics.

5. Long Noncoding RNA Regulation

A recently discovered source of inter-individual differences in pharmacogenomics are noncoding regulatory molecules, an extensive system representing more than 80% of the human transcriptome. To underline their importance, a recent large-scale study, screening a multitude of cell lines, has discovered the significance of long noncoding RNA (lncRNA) in anticancer treatments. LncRNA transcriptome appears to be as important as the protein-coding transcriptome in the analysis of drug responses to hundreds of compounds [106].
LncRNAs affect drug metabolism and disposition through regulating phase I and phase II biotransformation enzymes and ABC transporters [107]. LncRNAs have been found to regulate the expression of ABCB1, ABCC1 and ABCG2 and modulate therapy responses in breast cancer, lung cancer and gastric cancer (reviewed in [108]). As a further example, lncRNAs have been implicated in the chemoresistance of triple-negative breast cancer (TNBC). In addition to their effects on drug metabolism and efflux, they may also facilitate TNBC’s escape from apoptosis and evasion of the immune system (reviewed in Kansara et al. [109]).
LncRNA MALAT1 positively regulates the expression of ABCB1 and ABCC1 transporters via the activation of the STAT3 transcription factor [110]. In a French study using real-time PCR and reverse phase protein array (RPPA) techniques, MALAT1 was overexpressed in mammary tumors. Interestingly, an alternatively spliced transcript Δsv-MALAT1 appeared as an independent factor for a good prognosis in breast cancer patients [111]. Next, LINC00963 was found upregulated in head and neck carcinomas. As was further shown, a knockdown of LINC00963 resulted in a decreased invasion, colony formation and self-renewal via the regulation of ABCB5 [112]. A knockout of another lncRNA HNF1α-AS1 decreased the mRNA expression of various CYPs including CYP2B6, 2C8, 2C9, 2C19, 2E1, and 3A4, while the knockout of HNF4α-AS1 increased the CYP mRNA levels in HepaRG cells in vitro. CYPs are regulated by HNF1α-AS1 and HNF4α-AS1 via nuclear receptors PXR and CAR [113,114]. Noteworthy are also the roles of MDR1-targeting lncRNA HOTAIR and of other lncRNAs in the development of cisplatin chemoresistance in various cancers as was recently reviewed by Hu et al. [115]. A number of lncRNAs were markedly differentially expressed in 33 Czech hepatocellular carcinoma patients between patients with downregulated CYP expression vs. patients with normal CYP levels [116].
Thus, lncRNAs are a new and rapidly expanding area of pharmacogenomics. Despite its relatively short history, the accumulating evidence points to a highly valuable source for predicting cancer responsiveness in the future, especially if combined with gene variant and expression data.

6. Conclusions

The emerging role of pharmacogenomics in the NGS era provides valuable tools for determining inter-individual variations and should be soon taken into account in drug prescribing. The sources of “Big Data” in public databases originating from large genotyping projects might become a part of medical records.
Promising examples of clinical applications of pharmacogenomics may be seen in the area of targeted therapy of selected cancers with TKIs and the list is envisaged to expand soon. Historically, research of ABCs and CYPs in pharmacology and oncology was directed at their expression and germline genetic variation. More focus on somatic variations is needed. Databases like TCGA are vast repositories of tens of thousands of whole exome or genome sequences of tumor and germline DNA from patients with practically all types of cancers. A substantial portion of this potential still remains unexplored and unexploited. Rare variants in pharmacogenes and epigenetic features such as lncRNA represent another treasury of pharmacogenetic variation in a population, apart from the long-established hereditary germline polymorphisms and somatic variants in tumors. In addition, CNVs represent another piece of this puzzle. On the other hand, it has to be kept in mind that pharmacogenomics does not provide information about the post-translational modifications of encoded proteins and thus the role of this aspect for cancer therapy also needs further assessment.
Although the pharmacogenomics data on CYPs and ABCs reviewed above represent the forefront of experimental oncology, the lack of a thorough validation of published observations and standardization of new research studies in a way similar to clinical trials impedes further progress in clinical applications. Future studies should bear in mind that cancer therapy is an extremely complex process. Clinical data and all other interacting factors, for example, lifestyle and the resulting drug–drug and drug–dietary interactions, should be included as well as the real-time monitoring of patients’ response to therapy which sometimes exceeds a decade, e.g., in breast cancer. Truly precise personalized medicine may be enabled only after considering the aforementioned facts, robust replication of contemporary findings and methodological standardization, careful exploration of emerging NGS data and sophisticated utilization of integrative approaches.

Author Contributions

Conceptualization of the article, V.H.; V.H., P.H. and P.S. contributed to literature search and critical revision of the work, as well as writing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Czech Medical Council, project no. AZV NV19-08-00113 (P.S.) and Charles University Grant Agency projects no. UNCE/MED/006 (V.H.) and no. 698119 (P.H.)

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Vogel, F. Moderne Probleme der Humangenetik. Ergeb. Inn. Med. Kinderheilkd 1959, 12, 52–125. [Google Scholar]
  2. Daly, A.K. Pharmacogenetics: A general review on progress to date. Br. Med. Bull. 2017, 124, 65–79. [Google Scholar] [CrossRef] [PubMed]
  3. Barbarino, J.M.; Whirl-Carrillo, M.; Altman, R.B.; Klein, T.E. PharmGKB: A worldwide resource for pharmacogenomic information. Wiley Interdiscip. Rev. Syst. Biol. Med. 2018, 10, e1417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Relling, M.V.; Evans, W.E. Pharmacogenomics in the clinic. Nature 2015, 526, 343–350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Hyman, D.M.; Taylor, B.S.; Baselga, J. Implementing Genome-Driven Oncology. Cell 2017, 168, 584–599. [Google Scholar] [CrossRef]
  6. Fujikura, K.; Ingelman-Sundberg, M.; Lauschke, V.M. Genetic variation in the human cytochrome P450 supergene family. Pharmacogenet. Genom. 2015, 25, 584–594. [Google Scholar] [CrossRef]
  7. Kozyra, M.; Ingelman-Sundberg, M.; Lauschke, V.M. Rare genetic variants in cellular transporters, metabolic enzymes, and nuclear receptors can be important determinants of interindividual differences in drug response. Genet. Med. 2017, 19, 20–29. [Google Scholar] [CrossRef] [Green Version]
  8. Ingelman-Sundberg, M.; Mkrtchian, S.; Zhou, Y.; Lauschke, V.M. Integrating rare genetic variants into pharmacogenetic drug response predictions. Hum. Genom. 2018, 12, 26. [Google Scholar] [CrossRef]
  9. PubMed. Available online: https://pubmed.ncbi.nlm.nih.gov/ (accessed on 31 July 2020).
  10. Hertz, D.L.; McLeod, H.L. Use of pharmacogenetics for predicting cancer prognosis and treatment exposure, response and toxicity. J. Hum. Genet. 2013, 58, 346–352. [Google Scholar] [CrossRef] [Green Version]
  11. McLeod, H.L. Cancer pharmacogenomics: Early promise, but concerted effort needed. Science 2013, 339, 1563–1566. [Google Scholar] [CrossRef] [Green Version]
  12. Lam, S.W.; Frederiks, C.N.; van der Straaten, T.; Honkoop, A.H.; Guchelaar, H.J.; Boven, E. Genotypes of CYP2C8 and FGD4 and their association with peripheral neuropathy or early dose reduction in paclitaxel-treated breast cancer patients. Br. J. Cancer Suppl. 2016, 115, 1335–1342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Abraham, J.E.; Guo, Q.; Dorling, L.; Tyrer, J.; Ingle, S.; Hardy, R.; Vallier, A.L.; Hiller, L.; Burns, R.; Jones, L.; et al. Replication of Genetic Polymorphisms Reported to Be Associated with Taxane-Related Sensory Neuropathy in Patients with Early Breast Cancer Treated with Paclitaxel. Clin. Cancer Res. 2014, 20, 2466–2475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. FDA. Available online: https://www.fda.gov/Drugs/ScienceResearch/ucm572698.htm (accessed on 4 May 2020).
  15. Cascorbi, I.; Werk, A.N. Advances and challenges in hereditary cancer pharmacogenetics. Expert Opin. Drug Metab. Toxicol. 2017, 13, 73–82. [Google Scholar] [CrossRef] [PubMed]
  16. EMA. Available online: https://www.ema.europa.eu/ (accessed on 4 May 2020).
  17. Pharmacogenomics Knowledgebase. Available online: https://www.pharmgkb.org/ (accessed on 31 July 2020).
  18. Whirl-Carrillo, M.; McDonagh, E.M.; Hebert, J.M.; Gong, L.; Sangkuhl, K.; Thorn, C.F.; Altman, R.B.; Klein, T.E. Pharmacogenomics knowledge for personalized medicine. Clin. Pharmacol. Ther. 2012, 92, 414–417. [Google Scholar] [CrossRef]
  19. Drogemoller, B.I.; Wright, G.E.B.; Shih, J.; Monzon, J.G.; Gelmon, K.A.; Ross, C.J.D.; Amstutz, U.; Carleton, B.C. CYP2D6 as a treatment decision aid for ER-positive non-metastatic breast cancer patients: A systematic review with accompanying clinical practice guidelines. Breast Cancer Res. Treat. 2019, 173, 521–532. [Google Scholar] [CrossRef]
  20. Goetz, M.P.; Sangkuhl, K.; Guchelaar, H.J.; Schwab, M.; Province, M.; Whirl-Carrillo, M.; Symmans, W.F.; McLeod, H.L.; Ratain, M.J.; Zembutsu, H.; et al. Clinical Pharmacogenetics Implementation Consortium (CPIC) Guideline for CYP2D6 and Tamoxifen Therapy. Clin. Pharmacol. Ther. 2018, 103, 770–777. [Google Scholar] [CrossRef] [Green Version]
  21. Cronin-Fenton, D.P.; Damkier, P.; Lash, T.L. Metabolism and transport of tamoxifen in relation to its effectiveness: New perspectives on an ongoing controversy. Future Oncol. 2014, 10, 107–122. [Google Scholar] [CrossRef] [Green Version]
  22. Lyon, E.; Gastier Foster, J.; Palomaki, G.E.; Pratt, V.M.; Reynolds, K.; Sabato, M.F.; Scott, S.A.; Vitazka, P. Laboratory testing of CYP2D6 alleles in relation to tamoxifen therapy. Genet. Med. 2012, 14, 990–1000. [Google Scholar] [CrossRef] [Green Version]
  23. Shao, X.; Cai, J.; Zheng, Y.; Wang, J.; Feng, J.; Huang, Y.; Shi, L.; Chen, Z.; Guo, Y.; Wang, X. S4646 polymorphism in CYP19A1 gene is associated with the efficacy of hormone therapy in early breast cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 5309–5317. [Google Scholar]
  24. Shao, X.; Guo, Y.; Xu, X.; Zheng, Y.; Wang, J.; Chen, Z.; Huang, J.; Huang, P.; Cai, J.; Wang, X. The CYP19 RS4646 polymorphism IS related to the prognosis of stage I-II and operable stage III breast cancer. PLoS ONE 2015, 10, e0121535. [Google Scholar] [CrossRef] [Green Version]
  25. Scharfe, C.P.I.; Tremmel, R.; Schwab, M.; Kohlbacher, O.; Marks, D.S. Genetic variation in human drug-related genes. Genome Med. 2017, 9, 117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Xiao, Q.; Zhou, Y.; Winter, S.; Buttner, F.; Schaeffeler, E.; Schwab, M.; Lauschke, V.M. Germline variant burden in multidrug resistance transporters is a therapy-specific predictor of survival in breast cancer patients. Int. J. Cancer 2020, 146, 2475–2487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Menden, M.P.; Casale, F.P.; Stephan, J.; Bignell, G.R.; Iorio, F.; McDermott, U.; Garnett, M.J.; Saez-Rodriguez, J.; Stegle, O. The germline genetic component of drug sensitivity in cancer cell lines. Nat. Commun. 2018, 9, 3385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Caronia, D.; Patino-Garcia, A.; Perez-Martinez, A.; Pita, G.; Moreno, L.T.; Zalacain-Diez, M.; Molina, B.; Colmenero, I.; Sierrasesumaga, L.; Benitez, J.; et al. Effect of ABCB1 and ABCC3 polymorphisms on osteosarcoma survival after chemotherapy: A pharmacogenetic study. PLoS ONE 2011, 6, e26091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Kalra, S.; Kaur, R.P.; Ludhiadch, A.; Shafi, G.; Vashista, R.; Kumar, R.; Munshi, A. Association of CYP2C19*2 and ALDH1A1*1/*2 variants with disease outcome in breast cancer patients: Results of a global screening array. Eur. J. Clin. Pharmacol. 2018, 74, 1291–1298. [Google Scholar] [CrossRef] [PubMed]
  30. Green, H.; Soderkvist, P.; Rosenberg, P.; Horvath, G.; Peterson, C. mdr-1 single nucleotide polymorphisms in ovarian cancer tissue: G2677T/A correlates with response to paclitaxel chemotherapy. Clin. Cancer Res. 2006, 12, 854–859. [Google Scholar] [CrossRef] [Green Version]
  31. Chang, H.; Rha, S.Y.; Jeung, H.C.; Im, C.K.; Ahn, J.B.; Kwon, W.S.; Yoo, N.C.; Roh, J.K.; Chung, H.C. Association of the ABCB1 gene polymorphisms 2677G>T/A and 3435C>T with clinical outcomes of paclitaxel monotherapy in metastatic breast cancer patients. Ann. Oncol. 2009, 20, 272–277. [Google Scholar] [CrossRef]
  32. Chu, Y.H.; Li, H.; Tan, H.S.; Koh, V.; Lai, J.; Phyo, W.M.; Choudhury, Y.; Kanesvaran, R.; Chau, N.M.; Toh, C.K.; et al. Association of ABCB1 and FLT3 Polymorphisms with Toxicities and Survival in Asian Patients Receiving Sunitinib for Renal Cell Carcinoma. PLoS ONE 2015, 10, e0134102. [Google Scholar] [CrossRef] [Green Version]
  33. Beuselinck, B.; Lambrechts, D.; Van Brussel, T.; Wolter, P.; Cardinaels, N.; Joniau, S.; Lerut, E.; Karadimou, A.; Couchy, G.; Sebe, P.; et al. Efflux pump ABCB1 single nucleotide polymorphisms and dose reductions in patients with metastatic renal cell carcinoma treated with sunitinib. Acta. Oncol. 2014, 53, 1413–1422. [Google Scholar] [CrossRef] [Green Version]
  34. Harivenkatesh, N.; Kumar, L.; Bakhshi, S.; Sharma, A.; Kabra, M.; Velpandian, T.; Gogia, A.; Shastri, S.S.; Gupta, Y.K. Do polymorphisms in MDR1 and CYP3A5 genes influence the risk of cytogenetic relapse in patients with chronic myeloid leukemia on imatinib therapy? Leuk. Lymphoma 2017, 58, 1–9. [Google Scholar] [CrossRef]
  35. Ben Hassine, I.; Gharbi, H.; Soltani, I.; Ben Hadj Othman, H.; Farrah, A.; Amouri, H.; Teber, M.; Ghedira, H.; Ben Youssef, Y.; Safra, I.; et al. Molecular study of ABCB1 gene and its correlation with imatinib response in chronic myeloid leukemia. Cancer Chemother. Pharmacol. 2017, 80, 829–839. [Google Scholar] [CrossRef] [PubMed]
  36. Zheng, Q.; Wu, H.; Yu, Q.; Kim, D.H.; Lipton, J.H.; Angelini, S.; Soverini, S.; Vivona, D.; Takahashi, N.; Cao, J. ABCB1 polymorphisms predict imatinib response in chronic myeloid leukemia patients: A systematic review and meta-analysis. Pharmacogenom. J. 2015, 15, 127–134. [Google Scholar] [CrossRef] [PubMed]
  37. Kim, D.H.; Park, J.Y.; Sohn, S.K.; Lee, N.Y.; Baek, J.H.; Jeon, S.B.; Kim, J.G.; Suh, J.S.; Do, Y.R.; Lee, K.B. Multidrug resistance-1 gene polymorphisms associated with treatment outcomes in de novo acute myeloid leukemia. Int. J. Cancer 2006, 118, 2195–2201. [Google Scholar] [CrossRef] [PubMed]
  38. Green, H.; Falk, I.J.; Lotfi, K.; Paul, E.; Hermansson, M.; Rosenquist, R.; Paul, C.; Nahi, H. Association of ABCB1 polymorphisms with survival and in vitro cytotoxicty in de novo acute myeloid leukemia with normal karyotype. Pharmacogenom. J. 2012, 12, 111–118. [Google Scholar] [CrossRef] [PubMed]
  39. Megias-Vericat, J.E.; Rojas, L.; Herrero, M.J.; Boso, V.; Montesinos, P.; Moscardo, F.; Poveda, J.L.; Sanz, M.A.; Alino, S.F. Influence of ABCB1 polymorphisms upon the effectiveness of standard treatment for acute myeloid leukemia: A systematic review and meta-analysis of observational studies. Pharmacogenom. J. 2015, 15, 109–118. [Google Scholar] [CrossRef]
  40. He, H.; Yin, J.; Li, X.; Zhang, Y.; Xu, X.; Zhai, M.; Chen, J.; Qian, C.; Zhou, H.; Liu, Z. Association of ABCB1 polymorphisms with prognostic outcomes of anthracycline and cytarabine in Chinese patients with acute myeloid leukemia. Eur. J. Clin. Pharmacol. 2015, 71, 293–302. [Google Scholar] [CrossRef]
  41. Yin, J.Y.; Huang, Q.; Zhao, Y.C.; Zhou, H.H.; Liu, Z.Q. Meta-analysis on pharmacogenetics of platinum-based chemotherapy in non small cell lung cancer (NSCLC) patients. PLoS ONE 2012, 7, e38150. [Google Scholar] [CrossRef] [Green Version]
  42. Jakobsen Falk, I.; Lund, J.; Green, H.; Gruber, A.; Alici, E.; Lauri, B.; Blimark, C.; Mellqvist, U.H.; Swedin, A.; Forsberg, K.; et al. Pharmacogenetic study of the impact of ABCB1 single-nucleotide polymorphisms on lenalidomide treatment outcomes in patients with multiple myeloma: Results from a phase IV observational study and subsequent phase II clinical trial. Cancer Chemother. Pharmacol. 2018, 81, 183–193. [Google Scholar] [CrossRef] [Green Version]
  43. Bray, J.; Sludden, J.; Griffin, M.J.; Cole, M.; Verrill, M.; Jamieson, D.; Boddy, A.V. Influence of pharmacogenetics on response and toxicity in breast cancer patients treated with doxorubicin and cyclophosphamide. Br. J. Cancer 2010, 102, 1003–1009. [Google Scholar] [CrossRef] [Green Version]
  44. Lal, S.; Wong, Z.W.; Sandanaraj, E.; Xiang, X.; Ang, P.C.; Lee, E.J.; Chowbay, B. Influence of ABCB1 and ABCG2 polymorphisms on doxorubicin disposition in Asian breast cancer patients. Cancer Sci. 2008, 99, 816–823. [Google Scholar] [CrossRef]
  45. Maggini, V.; Buda, G.; Martino, A.; Presciuttini, S.; Galimberti, S.; Orciuolo, E.; Barale, R.; Petrini, M.; Rossi, A.M. MDR1 diplotypes as prognostic markers in multiple myeloma. Pharm. Genom. 2008, 18, 383–389. [Google Scholar] [CrossRef] [PubMed]
  46. Wu, H.; Kang, H.; Liu, Y.; Xiao, Q.; Zhang, Y.; Sun, M.; Liu, D.; Wang, Z.; Zhao, H.; Yao, W.; et al. Association of ABCB1 genetic polymorphisms with susceptibility to colorectal cancer and therapeutic prognosis. Pharmacogenomics 2013, 14, 897–911. [Google Scholar] [CrossRef] [PubMed]
  47. Grau, J.J.; Caballero, M.; Campayo, M.; Jansa, S.; Vargas, M.; Alos, L.; Monzo, M. Gene single nucleotide polymorphism accumulation improves survival in advanced head and neck cancer patients treated with weekly paclitaxel. Laryngoscope 2009, 119, 1484–1490. [Google Scholar] [CrossRef]
  48. Hertz, D.L.; Motsinger-Reif, A.A.; Drobish, A.; Winham, S.J.; McLeod, H.L.; Carey, L.A.; Dees, E.C. CYP2C8*3 predicts benefit/risk profile in breast cancer patients receiving neoadjuvant paclitaxel. Breast Cancer Res. Treat. 2012, 134, 401–410. [Google Scholar] [CrossRef] [Green Version]
  49. Angelini, S.; Soverini, S.; Ravegnini, G.; Barnett, M.; Turrini, E.; Thornquist, M.; Pane, F.; Hughes, T.P.; White, D.L.; Radich, J.; et al. Association between imatinib transporters and metabolizing enzymes genotype and response in newly diagnosed chronic myeloid leukemia patients receiving imatinib therapy. Haematologica 2013, 98, 193–200. [Google Scholar] [CrossRef] [PubMed]
  50. Adeagbo, B.A.; Bolaji, O.O.; Olugbade, T.A.; Durosinmi, M.A.; Bolarinwa, R.A.; Masimirembwa, C. Influence of CYP3A5*3 and ABCB1 C3435T on clinical outcomes and trough plasma concentrations of imatinib in Nigerians with chronic myeloid leukaemia. J. Clin. Pharm. Ther. 2016, 41, 546–551. [Google Scholar] [CrossRef]
  51. Kafka, A.; Sauer, G.; Jaeger, C.; Grundmann, R.; Kreienberg, R.; Zeillinger, R.; Deissler, H. Polymorphism C3435T of the MDR-1 gene predicts response to preoperative chemotherapy in locally advanced breast cancer. Int. J. Oncol. 2003, 22, 1117–1121. [Google Scholar] [CrossRef]
  52. Narumiya, K.; Metzger, R.; Bollschweiler, E.; Alakus, H.; Brabender, J.; Drebber, U.; Holscher, A.H.; Warnecke-Eberz, U. Impact of ABCB1 C3435T polymorphism on lymph node regression in multimodality treatment of locally advanced esophageal cancer. Pharmacogenomics 2011, 12, 205–214. [Google Scholar] [CrossRef]
  53. McLeod, H.L.; Sargent, D.J.; Marsh, S.; Green, E.M.; King, C.R.; Fuchs, C.S.; Ramanathan, R.K.; Williamson, S.K.; Findlay, B.P.; Thibodeau, S.N.; et al. Pharmacogenetic predictors of adverse events and response to chemotherapy in metastatic colorectal cancer: Results from North American Gastrointestinal Intergroup Trial N9741. J. Clin. Oncol. 2010, 28, 3227–3233. [Google Scholar] [CrossRef]
  54. Sensorn, I.; Sirachainan, E.; Chamnanphon, M.; Pasomsub, E.; Trachu, N.; Supavilai, P.; Sukasem, C.; Pinthong, D. Association of CYP3A4/5, ABCB1 and ABCC2 polymorphisms and clinical outcomes of Thai breast cancer patients treated with tamoxifen. Pharmgenom. Pers. Med. 2013, 6, 93–98. [Google Scholar] [CrossRef] [Green Version]
  55. Ravegnini, G.; Urbini, M.; Simeon, V.; Genovese, C.; Astolfi, A.; Nannini, M.; Gatto, L.; Saponara, M.; Ianni, M.; Indio, V.; et al. An exploratory study by DMET array identifies a germline signature associated with imatinib response in gastrointestinal stromal tumor. Pharmacogenom. J. 2019, 19, 390–400. [Google Scholar] [CrossRef]
  56. Li, M.; Seiser, E.L.; Baldwin, R.M.; Ramirez, J.; Ratain, M.J.; Innocenti, F.; Kroetz, D.L. ABC transporter polymorphisms are associated with irinotecan pharmacokinetics and neutropenia. Pharmacogenom. J. 2018, 18, 35–42. [Google Scholar] [CrossRef]
  57. Kiyotani, K.; Mushiroda, T.; Imamura, C.K.; Hosono, N.; Tsunoda, T.; Kubo, M.; Tanigawara, Y.; Flockhart, D.A.; Desta, Z.; Skaar, T.C.; et al. Significant effect of polymorphisms in CYP2D6 and ABCC2 on clinical outcomes of adjuvant tamoxifen therapy for breast cancer patients. J. Clin. Oncol. 2010, 28, 1287–1293. [Google Scholar] [CrossRef] [Green Version]
  58. Ansari, M.; Sauty, G.; Labuda, M.; Gagne, V.; Rousseau, J.; Moghrabi, A.; Laverdiere, C.; Sinnett, D.; Krajinovic, M. Polymorphism in multidrug resistance-associated protein gene 3 is associated with outcomes in childhood acute lymphoblastic leukemia. Pharmacogenom. J. 2012, 12, 386–394. [Google Scholar] [CrossRef] [Green Version]
  59. Angelini, S.; Pantaleo, M.A.; Ravegnini, G.; Zenesini, C.; Cavrini, G.; Nannini, M.; Fumagalli, E.; Palassini, E.; Saponara, M.; Di Battista, M.; et al. Polymorphisms in OCTN1 and OCTN2 transporters genes are associated with prolonged time to progression in unresectable gastrointestinal stromal tumours treated with imatinib therapy. Pharmacol. Res. 2013, 68, 1–6. [Google Scholar] [CrossRef]
  60. Han, J.Y.; Lee, Y.S.; Shin, E.S.; Hwang, J.A.; Nam, S.; Hong, S.H.; Ghang, H.Y.; Kim, J.Y.; Yoon, S.J.; Lee, J.S. A genome-wide association study of survival in small-cell lung cancer patients treated with irinotecan plus cisplatin chemotherapy. Pharmacogenom. J. 2014, 14, 20–27. [Google Scholar] [CrossRef]
  61. Zhao, J.; Li, W.; Zhu, D.; Yu, Q.; Zhang, Z.; Sun, M.; Cai, S.; Zhang, W. Association of single nucleotide polymorphisms in MTHFR and ABCG2 with the different efficacy of first-line chemotherapy in metastatic colorectal cancer. Med. Oncol. 2014, 31, 802. [Google Scholar] [CrossRef] [Green Version]
  62. Delord, M.; Rousselot, P.; Cayuela, J.M.; Sigaux, F.; Guilhot, J.; Preudhomme, C.; Guilhot, F.; Loiseau, P.; Raffoux, E.; Geromin, D.; et al. High imatinib dose overcomes insufficient response associated with ABCG2 haplotype in chronic myelogenous leukemia patients. Oncotarget 2013, 4, 1582–1591. [Google Scholar] [CrossRef]
  63. De Mattia, E.; Toffoli, G.; Polesel, J.; D’Andrea, M.; Corona, G.; Zagonel, V.; Buonadonna, A.; Dreussi, E.; Cecchin, E. Pharmacogenetics of ABC and SLC transporters in metastatic colorectal cancer patients receiving first-line FOLFIRI treatment. Pharmacogenet. Genom. 2013, 23, 549–557. [Google Scholar] [CrossRef]
  64. Limviphuvadh, V.; Tan, C.S.; Konishi, F.; Jenjaroenpun, P.; Xiang, J.S.; Kremenska, Y.; Mu, Y.S.; Syn, N.; Lee, S.C.; Soo, R.A.; et al. Discovering novel SNPs that are correlated with patient outcome in a Singaporean cancer patient cohort treated with gemcitabine-based chemotherapy. BMC Cancer 2018, 18, 555. [Google Scholar] [CrossRef] [Green Version]
  65. Dong, N.; Yu, J.; Wang, C.; Zheng, X.; Wang, Z.; Di, L.; Song, G.; Zhu, B.; Che, L.; Jia, J.; et al. Pharmacogenetic assessment of clinical outcome in patients with metastatic breast cancer treated with docetaxel plus capecitabine. J. Cancer Res. Clin. Oncol. 2012, 138, 1197–1203. [Google Scholar] [CrossRef]
  66. Verboom, M.C.; Kloth, J.S.L.; Swen, J.J.; Sleijfer, S.; Reyners, A.K.L.; Steeghs, N.; Mathijssen, R.H.J.; Gelderblom, H.; Guchelaar, H.J. Genetic polymorphisms in ABCG2 and CYP1A2 are associated with imatinib dose reduction in patients treated for gastrointestinal stromal tumors. Pharmacogenom. J. 2019, 19, 473–479. [Google Scholar] [CrossRef]
  67. Le Morvan, V.; Litiere, S.; Laroche-Clary, A.; Ait-Ouferoukh, S.; Bellott, R.; Messina, C.; Cameron, D.; Bonnefoi, H.; Robert, J. Identification of SNPs associated with response of breast cancer patients to neoadjuvant chemotherapy in the EORTC-10994 randomized phase III trial. Pharmacogenom. J. 2015, 15, 63–68. [Google Scholar] [CrossRef]
  68. Wang, H.; Bian, T.; Liu, D.; Jin, T.; Chen, Y.; Lin, A.; Chen, C. Association analysis of CYP2A6 genotypes and haplotypes with 5-fluorouracil formation from tegafur in human liver microsomes. Pharmacogenomics 2011, 12, 481–492. [Google Scholar] [CrossRef]
  69. Kassogue, Y.; Quachouh, M.; Dehbi, H.; Quessar, A.; Benchekroun, S.; Nadifi, S. Functional polymorphism of CYP2B6 G15631T is associated with hematologic and cytogenetic response in chronic myeloid leukemia patients treated with imatinib. Med. Oncol. 2014, 31, 782. [Google Scholar] [CrossRef]
  70. Johnson, G.G.; Lin, K.; Cox, T.F.; Oates, M.; Sibson, D.R.; Eccles, R.; Lloyd, B.; Gardiner, L.J.; Carr, D.F.; Pirmohamed, M.; et al. CYP2B6*6 is an independent determinant of inferior response to fludarabine plus cyclophosphamide in chronic lymphocytic leukemia. Blood 2013, 122, 4253–4258. [Google Scholar] [CrossRef]
  71. Damkier, P.; Kjaersgaard, A.; Barker, K.A.; Cronin-Fenton, D.; Crawford, A.; Hellberg, Y.; Janssen, E.A.M.; Langefeld, C.; Ahern, T.P.; Lash, T.L. CYP2C19*2 and CYP2C19*17 variants and effect of tamoxifen on breast cancer recurrence: Analysis of the International Tamoxifen Pharmacogenomics Consortium dataset. Sci. Rep. 2017, 7, 7727. [Google Scholar] [CrossRef] [Green Version]
  72. Suzumura, T.; Kimura, T.; Kudoh, S.; Umekawa, K.; Nagata, M.; Matsuura, K.; Tanaka, H.; Mitsuoka, S.; Yoshimura, N.; Kira, Y.; et al. Reduced CYP2D6 function is associated with gefitinib-induced rash in patients with non-small cell lung cancer. BMC Cancer 2012, 12, 568. [Google Scholar] [CrossRef] [Green Version]
  73. Sugiyama, E.; Umemura, S.; Nomura, S.; Kirita, K.; Matsumoto, S.; Yoh, K.; Niho, S.; Ohmatsu, H.; Tsuboi, M.; Ohe, Y.; et al. Impact of single nucleotide polymorphisms on severe hepatotoxicity induced by EGFR tyrosine kinase inhibitors in patients with non-small cell lung cancer harboring EGFR mutations. Lung Cancer 2015, 90, 307–313. [Google Scholar] [CrossRef]
  74. Takimoto, T.; Kijima, T.; Otani, Y.; Nonen, S.; Namba, Y.; Mori, M.; Yokota, S.; Minami, S.; Komuta, K.; Uchida, J.; et al. Polymorphisms of CYP2D6 gene and gefitinib-induced hepatotoxicity. Clin. Lung Cancer 2013, 14, 502–507. [Google Scholar] [CrossRef]
  75. Kobayashi, H.; Sato, K.; Niioka, T.; Miura, H.; Ito, H.; Miura, M. Relationship Among Gefitinib Exposure, Polymorphisms of Its Metabolizing Enzymes and Transporters, and Side Effects in Japanese Patients with Non-Small-Cell Lung Cancer. Clin. Lung Cancer 2015, 16, 274–281. [Google Scholar] [CrossRef]
  76. Kijima, T.; Shimizu, T.; Nonen, S.; Furukawa, M.; Otani, Y.; Minami, T.; Takahashi, R.; Hirata, H.; Nagatomo, I.; Takeda, Y.; et al. Safe and successful treatment with erlotinib after gefitinib-induced hepatotoxicity: Difference in metabolism as a possible mechanism. J. Clin. Oncol. 2011, 29, e588–e590. [Google Scholar] [CrossRef]
  77. Hirose, T.; Fujita, K.; Kusumoto, S.; Oki, Y.; Murata, Y.; Sugiyama, T.; Ishida, H.; Shirai, T.; Nakashima, M.; Yamaoka, T.; et al. Association of pharmacokinetics and pharmacogenomics with safety and efficacy of gefitinib in patients with EGFR mutation positive advanced non-small cell lung cancer. Lung Cancer 2016, 93, 69–76. [Google Scholar] [CrossRef]
  78. Swaisland, H.C.; Cantarini, M.V.; Fuhr, R.; Holt, A. Exploring the relationship between expression of cytochrome P450 enzymes and gefitinib pharmacokinetics. Clin. Pharmacokinet. 2006, 45, 633–644. [Google Scholar] [CrossRef]
  79. Khrunin, A.; Ivanova, F.; Moisseev, A.; Khokhrin, D.; Sleptsova, Y.; Gorbunova, V.; Limborska, S. Pharmacogenomics of cisplatin-based chemotherapy in ovarian cancer patients of different ethnic origins. Pharmacogenomics 2012, 13, 171–178. [Google Scholar] [CrossRef]
  80. Nakajima, Y.; Yoshitani, T.; Fukushima-Uesaka, H.; Saito, Y.; Kaniwa, N.; Kurose, K.; Ozawa, S.; Aoyagi, N.; Kamatani, N.; Yamamoto, N.; et al. Impact of the haplotype CYP3A4*16B harboring the Thr185Ser substitution on paclitaxel metabolism in Japanese patients with cancer. Clin. Pharmacol. Ther. 2006, 80, 179–191. [Google Scholar] [CrossRef]
  81. The Cancer Genome Atlas (TCGA). Available online: https://www.cancer.gov/tcga/ (accessed on 31 July 2020).
  82. Robey, R.W.; Pluchino, K.M.; Hall, M.D.; Fojo, A.T.; Bates, S.E.; Gottesman, M.M. Revisiting the role of ABC transporters in multidrug-resistant cancer. Nat. Rev. Cancer 2018, 18, 452–464. [Google Scholar] [CrossRef]
  83. Tate, J.G.; Bamford, S.; Jubb, H.C.; Sondka, Z.; Beare, D.M.; Bindal, N.; Boutselakis, H.; Cole, C.G.; Creatore, C.; Dawson, E.; et al. COSMIC: The Catalogue Of Somatic Mutations In Cancer. Nucleic Acids Res. 2019, 47, D941–D947. [Google Scholar] [CrossRef] [Green Version]
  84. 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]
  85. Jensen, K.H.; Izarzugaza, J.M.G.; Juncker, A.S.; Hansen, R.B.; Hansen, T.F.; Timshel, P.; Blondal, T.; Jensen, T.S.; Rygaard-Hjalsted, E.; Mouritzen, P.; et al. Analysis of a gene panel for targeted sequencing of colorectal cancer samples. Oncotarget 2018, 9, 9043–9060. [Google Scholar] [CrossRef] [Green Version]
  86. McCawley, N.; Clancy, C.; O’Neill, B.D.; Deasy, J.; McNamara, D.A.; Burke, J.P. Mucinous Rectal Adenocarcinoma Is Associated with a Poor Response to Neoadjuvant Chemoradiotherapy: A Systematic Review and Meta-analysis. Dis. Colon Rectum 2016, 59, 1200–1208. [Google Scholar] [CrossRef]
  87. Reynolds, I.S.; O’Connell, E.; Fichtner, M.; McNamara, D.A.; Kay, E.W.; Prehn, J.H.M.; Furney, S.J.; Burke, J.P. Mucinous adenocarcinoma is a pharmacogenomically distinct subtype of colorectal cancer. Pharmacogenom. J. 2019, 20, 524–532. [Google Scholar] [CrossRef]
  88. Li, Y.; Roberts, N.D.; Wala, J.A.; Shapira, O.; Schumacher, S.E.; Kumar, K.; Khurana, E.; Waszak, S.; Korbel, J.O.; Haber, J.E.; et al. Patterns of somatic structural variation in human cancer genomes. Nature 2020, 578, 112–121. [Google Scholar] [CrossRef] [Green Version]
  89. Gerstung, M.; Jolly, C.; Leshchiner, I.; Dentro, S.C.; Gonzalez, S.; Rosebrock, D.; Mitchell, T.J.; Rubanova, Y.; Anur, P.; Yu, K.; et al. The evolutionary history of 2,658 cancers. Nature 2020, 578, 122–128. [Google Scholar] [CrossRef] [Green Version]
  90. Smith, J.C.; Sheltzer, J.M. Systematic identification of mutations and copy number alterations associated with cancer patient prognosis. eLife 2018, 7, e39217. [Google Scholar] [CrossRef]
  91. Santos, M.; Niemi, M.; Hiratsuka, M.; Kumondai, M.; Ingelman-Sundberg, M.; Lauschke, V.M.; Rodriguez-Antona, C. Novel copy-number variations in pharmacogenes contribute to interindividual differences in drug pharmacokinetics. Genet. Med. 2018, 20, 622–629. [Google Scholar] [CrossRef] [Green Version]
  92. Ahmed, J.H.; Makonnen, E.; Fotoohi, A.; Aseffa, A.; Howe, R.; Aklillu, E. CYP2D6 Genotype Predicts Plasma Concentrations of Tamoxifen Metabolites in Ethiopian Breast Cancer Patients. Cancers 2019, 11, 1353. [Google Scholar] [CrossRef] [Green Version]
  93. Petrovic, J.; Pesic, V.; Lauschke, V.M. Frequencies of clinically important CYP2C19 and CYP2D6 alleles are graded across Europe. Eur. J. Hum. Genet. 2020, 28, 88–94. [Google Scholar] [CrossRef] [Green Version]
  94. Cavallari, L.H.; Van Driest, S.L.; Prows, C.A.; Bishop, J.R.; Limdi, N.A.; Pratt, V.M.; Ramsey, L.B.; Smith, D.M.; Tuteja, S.; Duong, B.Q.; et al. Multi-site investigation of strategies for the clinical implementation of CYP2D6 genotyping to guide drug prescribing. Genet. Med. 2019, 21, 2255–2263. [Google Scholar] [CrossRef]
  95. Goetz, M.P.; Sun, J.X.; Suman, V.J.; Silva, G.O.; Perou, C.M.; Nakamura, Y.; Cox, N.J.; Stephens, P.J.; Miller, V.A.; Ross, J.S.; et al. Loss of heterozygosity at the CYP2D6 locus in breast cancer: Implications for germline pharmacogenetic studies. J. Natl. Cancer Inst. Monogr. 2014, 107. [Google Scholar] [CrossRef] [Green Version]
  96. Ng, D.; Hong, C.S.; Singh, L.N.; Johnston, J.J.; Mullikin, J.C.; Biesecker, L.G. Assessing the capability of massively parallel sequencing for opportunistic pharmacogenetic screening. Genet. Med. 2017, 19, 357–361. [Google Scholar] [CrossRef] [Green Version]
  97. Magnani, L.; Frige, G.; Gadaleta, R.M.; Corleone, G.; Fabris, S.; Kempe, M.H.; Verschure, P.J.; Barozzi, I.; Vircillo, V.; Hong, S.P.; et al. Acquired CYP19A1 amplification is an early specific mechanism of aromatase inhibitor resistance in ERalpha metastatic breast cancer. Nat. Genet. 2017, 49, 444–450. [Google Scholar] [CrossRef]
  98. Lavrov, A.V.; Ustaeva, O.A.; Adilgereeva, E.P.; Smirnikhina, S.A.; Chelysheva, E.Y.; Shukhov, O.A.; Shatokhin, Y.V.; Mordanov, S.V.; Turkina, A.G.; Kutsev, S.I. Copy number variation analysis in cytochromes and glutathione S-transferases may predict efficacy of tyrosine kinase inhibitors in chronic myeloid leukemia. PLoS ONE 2017, 12, e0182901. [Google Scholar] [CrossRef]
  99. Genovese, I.; Ilari, A.; Assaraf, Y.G.; Fazi, F.; Colotti, G. Not only P-glycoprotein: Amplification of the ABCB1-containing chromosome region 7q21 confers multidrug resistance upon cancer cells by coordinated overexpression of an assortment of resistance-related proteins. Drug Resist. Updat. 2017, 32, 23–46. [Google Scholar] [CrossRef]
  100. Hansen, S.N.; Ehlers, N.S.; Zhu, S.; Thomsen, M.B.; Nielsen, R.L.; Liu, D.; Wang, G.; Hou, Y.; Zhang, X.; Xu, X.; et al. The stepwise evolution of the exome during acquisition of docetaxel resistance in breast cancer cells. BMC Genom. 2016, 17, 442. [Google Scholar] [CrossRef] [Green Version]
  101. Spitzwieser, M.; Pirker, C.; Koblmuller, B.; Pfeiler, G.; Hacker, S.; Berger, W.; Heffeter, P.; Cichna-Markl, M. Promoter methylation patterns of ABCB1, ABCC1 and ABCG2 in human cancer cell lines, multidrug-resistant cell models and tumor, tumor-adjacent and tumor-distant tissues from breast cancer patients. Oncotarget 2016, 7, 73347–73369. [Google Scholar] [CrossRef] [Green Version]
  102. Sun, Y.; Shi, N.; Lu, H.; Zhang, J.; Ma, Y.; Qiao, Y.; Mao, Y.; Jia, K.; Han, L.; Liu, F.; et al. ABCC4 copy number variation is associated with susceptibility to esophageal squamous cell carcinoma. Carcinogenesis 2014, 35, 1941–1950. [Google Scholar] [CrossRef] [Green Version]
  103. Lek, M.; Karczewski, K.J.; Minikel, E.V.; Samocha, K.E.; Banks, E.; Fennell, T.; O’Donnell-Luria, A.H.; Ware, J.S.; Hill, A.J.; Cummings, B.B.; et al. Analysis of protein-coding genetic variation in 60,706 humans. Nature 2016, 536, 285–291. [Google Scholar] [CrossRef] [Green Version]
  104. Firth, H.V.; Richards, S.M.; Bevan, A.P.; Clayton, S.; Corpas, M.; Rajan, D.; Van Vooren, S.; Moreau, Y.; Pettett, R.M.; Carter, N.P. DECIPHER: Database of Chromosomal Imbalance and Phenotype in Humans Using Ensembl Resources. Am. J. Hum. Genet. 2009, 84, 524–533. [Google Scholar] [CrossRef] [Green Version]
  105. Clinvar. Available online: https://www.ncbi.nlm.nih.gov/clinvar/ (accessed on 31 July 2020).
  106. Nath, A.; Lau, E.Y.T.; Lee, A.M.; Geeleher, P.; Cho, W.C.S.; Huang, R.S. Discovering long noncoding RNA predictors of anticancer drug sensitivity beyond protein-coding genes. Proc. Natl. Acad. Sci. USA 2019, 116, 22020–22029. [Google Scholar] [CrossRef]
  107. Wang, Y.; Fang, Z.; Hong, M.; Yang, D.; Xie, W. Long-noncoding RNAs (lncRNAs) in drug metabolism and disposition, implications in cancer chemo-resistance. Acta. Pharm. Sin. B. 2020, 10, 105–112. [Google Scholar] [CrossRef]
  108. Zhang, X.; Yang, H. Research Progress on Long Non-coding RNAs and Drug Resistance of Breast Cancer. Clin. Breast Cancer 2020, 20, 275–282. [Google Scholar] [CrossRef]
  109. Kansara, S.; Pandey, V.; Lobie, P.E.; Sethi, G.; Garg, M.; Pandey, A.K. Mechanistic Involvement of Long Non-Coding RNAs in Oncotherapeutics Resistance in Triple-Negative Breast Cancer. Cells 2020, 9, 1511. [Google Scholar] [CrossRef]
  110. Fang, Z.; Chen, W.; Yuan, Z.; Liu, X.; Jiang, H. LncRNA-MALAT1 contributes to the cisplatin-resistance of lung cancer by upregulating MRP1 and MDR1 via STAT3 activation. Biomed. Pharmacother. 2018, 101, 536–542. [Google Scholar] [CrossRef]
  111. Meseure, D.; Vacher, S.; Lallemand, F.; Alsibai, K.D.; Hatem, R.; Chemlali, W.; Nicolas, A.; De Koning, L.; Pasmant, E.; Callens, C.; et al. Prognostic value of a newly identified MALAT1 alternatively spliced transcript in breast cancer. Br. J. Cancer Suppl. 2016, 114, 1395–1404. [Google Scholar] [CrossRef] [Green Version]
  112. Lee, S.P.; Hsieh, P.L.; Fang, C.Y.; Chu, P.M.; Liao, Y.W.; Yu, C.H.; Yu, C.C.; Tsai, L.L. LINC00963 Promotes Cancer Stemness, Metastasis, and Drug Resistance in Head and Neck Carcinomas via ABCB5 Regulation. Cancers 2020, 12, 1073. [Google Scholar] [CrossRef]
  113. Chen, L.; Bao, Y.; Piekos, S.C.; Zhu, K.; Zhang, L.; Zhong, X.B. A Transcriptional Regulatory Network Containing Nuclear Receptors and Long Noncoding RNAs Controls Basal and Drug-Induced Expression of Cytochrome P450s in HepaRG Cells. Mol. Pharmacol. 2018, 94, 749–759. [Google Scholar] [CrossRef]
  114. Chen, L.; Bao, Y.; Jiang, S.; Zhong, X.B. The Roles of Long Noncoding RNAs HNF1alpha-AS1 and HNF4alpha-AS1 in Drug Metabolism and Human Diseases. Non-Coding RNA 2020, 6, 24. [Google Scholar] [CrossRef]
  115. Hu, Y.; Zhu, Q.N.; Deng, J.L.; Li, Z.X.; Wang, G.; Zhu, Y.S. Emerging role of long non-coding RNAs in cisplatin resistance. Onco. Targets Ther. 2018, 11, 3185–3194. [Google Scholar] [CrossRef] [Green Version]
  116. Nekvindova, J.; Mrkvicova, A.; Zubanova, V.; Hyrslova Vaculova, A.; Anzenbacher, P.; Soucek, P.; Radova, L.; Slaby, O.; Kiss, I.; Vondracek, J.; et al. Hepatocellular carcinoma: Gene expression profiling and regulation of xenobiotic-metabolizing cytochromes P450. Biochem. Pharmacol. 2020, 177, 113912. [Google Scholar] [CrossRef]
Figure 1. Major functional aspects of ATP-binding cassette (ABC)/cytochrome P450 (CYP) in drug responses with the presumed role of their genetic variability.
Figure 1. Major functional aspects of ATP-binding cassette (ABC)/cytochrome P450 (CYP) in drug responses with the presumed role of their genetic variability.
Jpm 10 00108 g001
Figure 2. Top 30 mutated ATP-binding cassette transporters in top 200 mutated cases in The Cancer Genome Atlas (TCGA) across the four most common cancers types. The results shown here are in whole based upon data generated by the TCGA Research Network: [81].
Figure 2. Top 30 mutated ATP-binding cassette transporters in top 200 mutated cases in The Cancer Genome Atlas (TCGA) across the four most common cancers types. The results shown here are in whole based upon data generated by the TCGA Research Network: [81].
Jpm 10 00108 g002
Figure 3. Top 30 mutated cytochromes P450 in top 200 mutated cases in TCGA across the four most common cancers types. The results shown here are in whole based upon data generated by the TCGA Research Network: [81].
Figure 3. Top 30 mutated cytochromes P450 in top 200 mutated cases in TCGA across the four most common cancers types. The results shown here are in whole based upon data generated by the TCGA Research Network: [81].
Jpm 10 00108 g003
Table 1. Official U.S. Food and Drug Administration (FDA) pharmacogenomic biomarkers in CYPs for anticancer drugs.
Table 1. Official U.S. Food and Drug Administration (FDA) pharmacogenomic biomarkers in CYPs for anticancer drugs.
GeneDrugPharmacogenomics LevelRecommendationDisease
CYP1A2rucaparibinformativeno difference in PK among metabolizer phenotypesepithelial ovarian cancer, fallopian tube, primary peritoneal cancer
CYP2C9erdafitinibactionablelower dose for CYP2C9*3/*3 genotypeurothelial carcinoma
CYP2D6gefitinibactionableADR risk for poor metabolizersnon-small cell lung cancer
CYP2D6rucaparibinformativeno difference in PK among metabolizer phenotypesepithelial ovarian cancer, fallopian tube, primary peritoneal cancer
CYP2D6tamoxifenactionablepoor metabolizers have lower endoxifen concentrationsbreast cancer
PK—pharmacokinetics; ADR—adverse drug reactions.
Table 3. Information about germline copy number variations in ABC and CYP genes in Genome Aggregation Database (gnomAD) database.
Table 3. Information about germline copy number variations in ABC and CYP genes in Genome Aggregation Database (gnomAD) database.
GeneVariant IDConsequenceClassSizeAllele Count 1
ABCA3DEL_16_151377intronicdeletion122 bp1
DEL_16_151378intronicdeletion824 bp18
DEL_16_151379intronicdeletion434 bp208
DEL_16_151381intronicdeletion477 bp1
INS_16_98759intronicinsertion279 bp1
ABCD3INV_1_34inversion spaninversion19.4 Mb1
INV_1_35inversion spaninversion652 kb1
DUP_1_1758copy gainduplication321 kb2
DEL_1_6074intronicdeletion1.04 kb1
DEL_1_6075intronicdeletion6.0 kb1
DEL_1_6076intronicdeletion90 bp6408
INS_1_3772intronicinsertion281 bp2
DUP_1_1759partial duplicationduplication97 bp3
ABCE1CPX_4_1427inversion spancomplex87.0 Mb1
DEL_4_50987intronicdeletion485 bp1
CYP2D6MCNV_22_1026MCNV overlapmulti CNV12.2 kb2881
DEL_22_182964intronicdeletion105 bp1
CYP4A11INV_1_10inversion spaninversion58.3 Mb1
DUP_1_1171copy gainduplication118 kb3
DUP_1_1173partial duplicationduplication82.0 kb2
DUP_1_1174copy gainduplication201 kb6
CYP17A1CPX_10_3184inversion spancomplex106 Mb1
INV_10_527inversion spaninversion15.6 Mb1
DUP_10_30635partial duplicationduplication14.4 kb1
1 Out of 15,708 whole-genome sequences from unrelated individuals. MCNV—multiCNV.

Share and Cite

MDPI and ACS Style

Hlaváč, V.; Holý, P.; Souček, P. Pharmacogenomics to Predict Tumor Therapy Response: A Focus on ATP-Binding Cassette Transporters and Cytochromes P450. J. Pers. Med. 2020, 10, 108. https://doi.org/10.3390/jpm10030108

AMA Style

Hlaváč V, Holý P, Souček P. Pharmacogenomics to Predict Tumor Therapy Response: A Focus on ATP-Binding Cassette Transporters and Cytochromes P450. Journal of Personalized Medicine. 2020; 10(3):108. https://doi.org/10.3390/jpm10030108

Chicago/Turabian Style

Hlaváč, Viktor, Petr Holý, and Pavel Souček. 2020. "Pharmacogenomics to Predict Tumor Therapy Response: A Focus on ATP-Binding Cassette Transporters and Cytochromes P450" Journal of Personalized Medicine 10, no. 3: 108. https://doi.org/10.3390/jpm10030108

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