Next Article in Journal
Role of N-Cadherin in Epithelial-to-Mesenchymal Transition and Chemosensitivity of Colon Carcinoma Cells
Next Article in Special Issue
Current State and Future Challenges for PI3K Inhibitors in Cancer Therapy
Previous Article in Journal
Sex-Related Differences in Metastatic Melanoma Patients Treated with Immune Checkpoint Inhibition
Previous Article in Special Issue
IFI35 Promotes Renal Cancer Progression by Inhibiting pSTAT1/pSTAT6-Dependent Autophagy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Genomics and Proteomics in Lung Cancer Early Detection and Treatment †

by
Mohammad Hadi Abbasian
1,
Ali M. Ardekani
2,
Navid Sobhani
3,‡ and
Raheleh Roudi
4,*,‡
1
Department of Medical Genetics, National Institute of Genetic Engineering and Biotechnology (NIGEB), Tehran 1497716316, Iran
2
Department of Medical Biotechnology, National Institute of Genetic Engineering and Biotechnology (NIGEB), Tehran 1497716316, Iran
3
Department of Medicine, Section of Epidemiology and Population Sciences, Baylor College of Medicine, Houston, TX 77030, USA
4
Department of Radiology, Molecular Imaging Program at Stanford, Stanford University, Stanford, CA 94305, USA
*
Author to whom correspondence should be addressed.
Genomics and proteomics in lung cancer.
These authors contributed equally to this work.
Cancers 2022, 14(20), 5144; https://doi.org/10.3390/cancers14205144
Submission received: 7 September 2022 / Revised: 10 October 2022 / Accepted: 18 October 2022 / Published: 20 October 2022
(This article belongs to the Special Issue PI3K Pathway in Cancer)

Abstract

:

Simple Summary

Biomarkers for the prediction and efficacy of therapies are an urgent necessity for lung cancer patients. In this article, we summarize genomic and proteomic biomarkers utilized for the early detection and treatment of lung cancer, with a focus on immune checkpoint and PI3K pathways.

Abstract

Lung cancer is the leading cause of cancer-related death worldwide, with non-small-cell lung cancer (NSCLC) being the primary type. Unfortunately, it is often diagnosed at advanced stages, when therapy leaves patients with a dismal prognosis. Despite the advances in genomics and proteomics in the past decade, leading to progress in developing tools for early diagnosis, targeted therapies have shown promising results; however, the 5-year survival of NSCLC patients is only about 15%. Low-dose computed tomography or chest X-ray are the main types of screening tools. Lung cancer patients without specific, actionable mutations are currently treated with conventional therapies, such as platinum-based chemotherapy; however, resistances and relapses often occur in these patients. More noninvasive, inexpensive, and safer diagnostic methods based on novel biomarkers for NSCLC are of paramount importance. In the current review, we summarize genomic and proteomic biomarkers utilized for the early detection and treatment of NSCLC. We further discuss future opportunities to improve biomarkers for early detection and the effective treatment of NSCLC.

1. Introduction

Worldwide, the second most common type of cancer diagnosed in men and women is lung cancer, which is the most common cause of cancer-related deaths. There are two main types of lung cancers: non-small-cell lung cancer (NSCLC) and small-cell lung cancer (SCLC). Adenocarcinoma and squamous cell carcinoma are the main histological subtypes of NSCLC. Prostate, colon, and breast cancers have screening tools for early detection; however, there are generally less accurate early detection biomarkers for lung cancer [1]. GLOBOCAN estimated approximately 2.2 million new lung cancer cases and 1,796,144 deaths [2].
The five-year survival rate in stage I NSCLC patients is 70–90%. However, over 75% of lung cancer patients are diagnosed at advanced or metastatic stages when the five-year survival rates are significantly lower [2,3]. Its diagnosis at advanced stages is associated with significantly lower five-year survival.
The first strategies for lung cancer detection for many years had been annual chest radiography and sputum cytology. It has not been until recently, as medicine has advanced, that these strategies have been updated. The results of several studies, such as two studies conducted by Johns Hopkins University and the Memorial Sloan Kettering Center, revealed that chest X-ray (CXR) and sputum cytology had not been promising screening tools for reducing lung cancer mortality [4,5]. The Early Lung Cancer Action Project (ELCAP) was initiated in 1993 and evaluated high-risk persons for lung cancer by low-radiation-dose computed tomography (low-dose C.T.) [6]. In addition, the National Lung Screening Trial (NLST) was initiated in 2002. These studies showed that the use of low-dose C.T. reduced mortality from lung cancer; however, cumulative radiation exposure, a high-false positive rate, and costs are some disadvantages of low-dose C.T. [7]. More specific, less invasive, and cost-effective biomarkers are therefore urgently needed.
In the past several decades, the molecular profiling of lung cancer patients has shown that many mutations correlate with NSCLC [8,9,10]. Molecular profiling helps stratify patients with specific mutations to use several lines of targeted therapies. EGFR, KRAS, MET, BRAF, ERBB2, ALK, ROS1, and RET gene rearrangements are the main actionable genetic alterations in NSCLC [11,12,13,14]. Several generations of agents have been developed to target these mutations, improving the patients’ outcomes. However, disease progression and resistance to targeted drugs remain major challenges in NSCLC cancer treatment. In our opinion, the efficacy of these drugs depends on how accurate biomarkers could help stratify patients that would best respond to such targeted therapies. As medicine advances, more accurate biomarkers have been discovered that could better help predict therapy responses. Proteins are functional products from the genome. The dynamic range of protein concentration differs from cell to cell and from time to time. Therefore, proteomics has dynamic and complex entities. A quantitative understanding of proteomics provides the exact knowledge about protein networks of human cells and may improve our cancer detection and treatment tools [15].
The application of proteomics has resulted in subtype-specific, stage-specific, and metastasis-specific tumor biomarkers in lung cancer. Numerous studies have demonstrated that a significant number of proteins, including ENO1, selenium-binding protein 1 (SELENBP1), carbonic anhydrase (CA), heat shock 20KD-like protein, and transgelin (SM22-alpha), are associated with poor prognosis [16,17].
This review summarizes the current knowledge on genomics and proteomics biomarkers for the early detection and treatment of NSCLC. Furthermore, we focus on new biomarkers predicting the efficacy of immunotherapies and the PI3K–mTOR–AKT axis. Finally, we summarize the ongoing clinical trials investigating the AKT–PI3K–mTOR pathway in NSCLC.

2. Early Detection

Cancer early detection increases the opportunity for more successful treatment. Five years of survival in NSCLC patients are approximately 63% and 7% for localized and distant tumors, respectively [18]. Early detection of lung cancer will dramatically reduce the mortality rate of patients. Many genomic and proteomic biomarkers, such as DNA methylation, microRNA (miRNA), antitumor antibodies, and plasma proteins, have been investigated for early cancer detection, including lung cancer. Liquid biopsy comprises various sources, including sputum, blood, exhaled breath, sputum, bronchial aspirate, and bronchoalveolar lavages for the early detection of lung cancer [19,20,21] (Figure 1).

2.1. Genomic and Epigenomes

2.1.1. Methylation

DNA methylation (hypermethylation or hypomethylation) is an epigenetic modification that plays a pivotal role in silencing gene transcription and is crucial in many biological processes, including embryonic development, X-chromosome inactivation, genomic imprinting, chromatin structure, normal growth, and cellular proliferation. Abnormal DNA methylation has been observed in tumor initiation and proliferation of all forms of cancers [22]. The epigenetic pathway is potentially reversible and is involved in tumorigenesis by three main mechanisms: DNA hypermethylation, global hypomethylation of the genome, and histone modifications.
DNA methylation can be assessed by different techniques, such as digital PCR (dPCR), methylation-specific PCR (MSP), methyl-CpG-binding domain (MBD), real-time quantitative MSP, quantitative methylation-specific PCR (qMSP), multiplex nested methylation-specific PCR, and whole-genome bisulfite sequencing (WGBS) [23]. Aberrant DNA methylation is catalyzed by DNA methyltransferases (DNMTs) through the promoter methylation of tumor suppressor genes. It is associated with many types of cancers, including NSCLC, autoimmune diseases, diabetes, and multiple sclerosis [24,25,26]. Hypermethylation of DNA can silence tumor suppressor genes, and hypomethylation of DNA can lead to transcriptional activation proto-oncogenes. Methylation patterns have been associated with different cancer aspects, such as disease stage, survival, and response to treatment. DNA methylation could predict the response to platinum-based chemotherapy in lung cancer patients [27,28].
Methylation of single or panel genes is reported in lung cancer and is associated with response to therapies (Table 1). Several studies have shown that the methylation of RASSF1A, SHOX2, and PTGER4 genes is greatly helpful in the early diagnosis of lung cancer [29,30,31]. RASSF1A is a tumor-suppressor gene, and its methylation-associated inactivation is reported in NSCLC, breast, and gastrointestinal cancer (G.C.) [32]. Combining two or more gene methylation detection methods have improved sensitivity and specificity compared to a single-gene methylation detection method. The promoter methylation TAC1, HOXA17, and SOX17 in sputum obtained from a prospective cohort of 150 NSCLC patients and 60 controls showed sensitivity and specificity of 98% and 71%, respectively [33]. The sensitivity and specificity of the methylation panel of SOX17, HOXA9, AJAP1, PTGDR, UNCX, and MARCH11 in NSCLC patients were 96.7% and 60%, respectively [34]. Based on the study by Liu et al., the methylation frequency PCDHGB6, HOXA9, MGMT, and miR-126 reached 85.2% sensitivity and 81.5% specificity [34,35].

2.1.2. miRNAs

MicroRNAs (miRNAs) are a large family of short single-stranded RNAs that regulate gene expression. MicroRNAs play a pivotal role in biological processes such as cell proliferation, differentiation, and survival [44]. In cancer, miRNAs may act as either tumor suppressors or oncogenes [45,46]. Profiling miRNA in plasma, serum, and sputum are promising noninvasive biomarkers for the early detection of NSCLC (Table 2). Numerous studies have reported the great potential of miRNAs in lung cancer diagnosis. A recent systematic review and meta-analysis by Wang et al. declared that miRNA-21 has potential clinical value in the diagnosis and prognosis of lung cancer [47]. Lu et al. investigated 723 human microRNAs in 106 plasmas in healthy individuals and patients with NSCLC or SCLC. They identified the diagnostic features of six miRNAs (miR-17, miR-190b, miR-19a, miR-19b, miR-26b, and miR-375) that have value for discriminating lung cancer patients from healthy individuals [48]. A panel of three miRNAs (miR-125a-5p, miR-25, and miR-126) in the study of Wang et al. distinguished early stage lung cancer patients from control subjects [49]. Pan et al. studied the expression of miR-33a-5p and miR-128-3p lung cancer tissues and cell lines. They found that the expression of these miRNAs was downregulated in tissues, cell lines, and the whole blood of early stage lung cancer patients [50]. A multicenter study analyzed bloodborne RNA signatures in 3102 patients and found that a 15-miRNA signature might help distinguish patients diagnosed with lung cancer from cancer-free individuals [51]. Khandelwal et al. explored the role of miR-590-5p in NSCLC patients in plasma samples, and they found that miRNA functions as a tumor suppressor in NSCLC and downregulated in NSCLC patients compared to healthy controls [52].

2.2. Proteomic Early Detection

Proteomics is a powerful approach to studying numerous proteins that generates information about biomarker identification, molecular interactions, and signaling pathways [60]. Mass spectrometry (MS) is an analytical tool in biomarker discovery that can be used to measure the mass-to-charge ratio of particles. MS-based proteomics has been applied to biomarker discovery for several decades. In oncology, MS approaches provided a decisive result for the early detection of breast, colorectal, prostate, and ovarian cancer [61,62,63,64]. Proteomic biomarkers could be used from several sources, including plasma, serum, tears, sputum, saliva, and urine [65,66,67]. We can also find proteomic biomarkers in exhaled breath condensate, bronchoalveolar lavage fluid, and pleural effusion for lung cancer [68] (Table 3). Zhang et al. profiled urine proteome in several cancer types and healthy control. In this study, five biomarkers (FTL: ferritin light chain; MAPK1IP1L: mitogen-activated protein kinase 1 interacting protein 1-like; FGB: fibrinogen beta chain; RAB33B: RAB33B, member RAS oncogene family; RAB15: RAB15, member RAS oncogene family) could distinguish lung cancer patients from healthy individuals [69].

3. Actionable Markers for Treatment

3.1. Genomics Biomarkers

In the United States, the National Cancer Institute (NCI) and National Human Genome Research Institute launched a genomic cancer program, The Cancer Genome Atlas (TCGA), in 2006. This project characterized 20,000 primary cancer and matched standard samples of 33 major cancer types in the genomic, epigenomic, transcriptomics, clinical, and proteomics information level [77]. Over several years, TCGA data has been available in several databases, such as cBioPortal for Cancer Genomics and The Cancer Proteome Atlas Portal (TCPA) [78,79]. The Catalogue of Somatic Mutations in Cancer (COSMIC) and the International Cancer Genome Consortium (ICGC) provide a comprehensive genomic data visualization, integration, and analysis source [80,81]. Integrative analysis from the Pan-Cancer Analysis of Whole Genomes (PCAWG) Consortium of the International Cancer Genome Consortium (ICGC) and TCGA revealed that cancer genomes contained four to five driver mutations [82]. Recent results from the PCAWG Consortium shed light on different aspects of cancers, such as mutational processes, tumor evolution, and diverse transcriptional consequences [82]. The combination of omics data from various databases such as Gene Expression Omnibus and TCGA is one of the promising approaches to finding potential biomarkers for understanding cancer development and progression [83].
Genetic variations such as single-nucleotide polymorphisms (SNPs) and deletion variants (indels) are associated with genetic susceptibility to cancer [84,85]. Pharmacogenetics studies explore genetic variations and focus on individual genetic variability and its correlation with drug efficacy, toxicity, and overall cancer treatment outcomes [86,87].
Several driver gene alterations have been detected in NSCLC patients, including Epidermal growth factor receptor (EGFR), Kirsten rat sarcoma viral oncogene homolog (K-Ras), tumor protein p53 (TP53), MET, B-Raf proto-oncogene, serine/threonine kinase (BRAF), ERBB2, the anaplastic lymphoma kinase (ALK), ROS proto-oncogene 1 receptor tyrosine kinase (ROS), RET, and NTRK rearrangement were detected [88,89,90,91].
Epidermal growth factor receptor (EGFR) controls critical signaling pathways, including RAS/MAPK/ERK, PI3K/AKT, and STATS, and is responsible for cellular proliferation, differentiation, and survival. EGFR is an oncogenic driver in NSCLC, breast, and glioblastoma. EGFR has been linked to types of cancer, including metastatic colorectal cancer (CRC), pancreatic cancer, and head and neck squamous cell carcinoma (HNSCC). Its gene is located on the short arm of chromosome 7 at position 11.2 (7p11.2) [92]. EGFR is overexpressed and/or mutated in cancers such as glioblastoma, head and neck, pancreatic, breast, and metastatic colorectal cancer [93,94,95,96,97]. EGFR is overexpressed in 40–80% of NSCLC patients [98,99]. Abnormal EGFR expression has been associated with mutations in the EGFR, with approximately 10–15% in Caucasians with adenocarcinoma and 50% in lung tumors in East Asian populations [100]. Common mutations of EGFR are summarized in Figure 2. Several tyrosine kinase inhibitors (TKIs) targeting EGFR have been developed, including Erlotinib (Tarceva), Afatinib (Gilotrif), Gefitinib (Iressa), Osimertinib (Tagrisso), and Dacomitinib (Vizimpro), which improved progression-free survival (PFS), time-to-progression, and overall survival compared to standard chemotherapy in NSCLC patients. Gefitinib, Erlotinib, and Afatinib are common options for lung cancer patients with exon 21 Leu858Arg mutation or EGFR-exon 19 deletions. Unfortunately, patients with EGFR-mutant NSCLC develop disease progression within 10–14 months [101]. Different mechanisms underlie intrinsic or acquired resistance to EGFR-targeted inhibitors. T790M mutation in EGFR is responsible for approximately 50% of cases of acquired resistance [102]. Other important resistance mechanisms involve AKT mutations, loss of PTEN, activation of alternative signaling, and BCL2-like 11/BIM deletion polymorphism [103].
The ALK gene consists of 30 exons and is located on chromosome 2p23. ALK is a tyrosine kinase receptor of the insulin receptor superfamily that plays an essential role in embryonic and neural development and human immune responses. The role of ALK in human tumorigenesis was discovered in anaplastic large cell lymphoma (ALCL) in 1994. The presence of EML4-ALK fusion was shown in breast, colorectal, and non-small-cell lung cancers. KIF5B, RET, ROS1, and ALK fusion has been detected in lung cancer. ALK could have different fusion partners, including NPM, in non-Hodgkin’s lymphoma, VCL in Renal cell carcinoma, FN1 in gastrointestinal leiomyomas, and TRK-fused gene (TFG) in anaplastic large-cell lymphoma.
The V-RAF murine sarcoma viral oncogene homolog B (BRAF) encodes serine/threonine kinase. BRAF is a constituent of the MAPK pathway that regulates various cellular processes, including cell growth, proliferation, and survival. The effect of BRAF mutations in lung cancer patients was first reported in 2011 [104]. Mutations in BRAF are seen in 2–5 lung adenocarcinomas [105].
ROS1 (ROS proto-oncogene 1) is located on chromosome 6p22 and encodes 2347 amino acid residues. ROS1 is the subfamily of receptor tyrosine kinase, which regulates cell proliferation, survival, and migration [106]. FIG (GOPC)-ROS1 was the first ROS1 fusion gene reported in the human glioblastoma cell line U118MG in the early 1980s. Genetic rearrangements of ROS1 have been shown in different malignancies, including gastric adenocarcinoma, colorectal cancer, ovarian cancer, inflammatory myofibroblastic tumor, and angiosarcoma [106,107,108,109,110]. The first ROS1 fusion gene in NSCLC was identified in 2007. Previous studies have shown that ROS1 fusions account for 1–2% of all cases of NSCLC [88]. NSCLC patients with ROS1 fusion have several distinct clinical characteristics: they are typically female, younger (<50 years of age), and never or light smokers [111,112,113].

3.1.1. The PI3K Pathway in NSCLC

The phosphatidylinositol 3-kinase (PI3K)/Akt/mechanistic target of the rapamycin (mTOR) pathway has a pivotal role in regulating signal transduction and biological processes such as cell growth and survival, apoptosis, angiogenesis, tumor invasion, and metastasis. PI3Ks, AKT, and mTOR are the core components of the PI3K–AKT–mTOR signaling pathway, and their hyperactivation is observed in many cancers [114,115]. Therefore, alterations in this pathway lead to initiating and enhancing cancer progression. Consequently, this pathway is considered a target for novel anticancer therapies.
In NSCLC, the PI3K/Akt/mTOR pathway has been heavily implicated in carcinogenesis and disease advancement in NSCLC. Mutations in receptor tyrosine kinases (RTKs) and KRAS have been linked to the activation of the PI3K/Akt/mTOR axis in NSCLC [116,117]. Alterations in each component of this pathway may lead to lung cancer development and progression [118]. Mutations or increases in the number of PIK3CA of PIK3CA are frequently found in patients with NSCLC. PIK3CA mutations are one of the most common gene changes in human cancers. Mutations or increases in the number of PIK3CA of PIK3CA are frequently found in patients with NSCLC.
A large study by Scheffler et al. analyzed 1144 consecutive NSCLC patients’ tumor tissue for PIK3CA mutations. This study showed that 3.7% of patients have PIK3CA mutations in exons nine and 20 [118,119]. Yamamoto et al. investigated PIK3CA mutations in exons 9 and 20 in 86 NSCLC cell lines and 356 resected NSCLC tumors and revealed that PIK3CA mutations in NSCLC cell lines were 4.7% and tumors 1.6%. Increased PIK3CA copy number was detected in 9.3% of NSCLC cell lines and 17.1% of NSCLC tumors [120].
Recurrent alterations were observed in PIK3CA that were considered to contribute Osimertinib resistance [121]. A systematic review and meta-analysis study was conducted with a total of 13 studies involving 3908 NSCLC patients. The results indicated that PIK3CA mutations are associated with OS, PFS, and cancer-specific survival (CSS) [122]. PIK3CA mutation could be considered an independent prognostic factor for reduced PFS of EGFR-TKIs treatment and worse OS in NSCLC patients [122,123].
Several PI3K/Akt/mTOR signaling pathway inhibitors are currently developing as promising targets for new anticancer drugs in preclinical investigations and clinical trials. Pilaralisib is a highly selective inhibitor of the class I PI3Ks and successfully inhibits tumor growth in vivo. In phase I trials, Pilaralisib was assessed as a tolerable monotherapy in advanced solid tumors patients [109]. Alpelisib is a p110 alpha isoform-specific PI3K inhibitor that has been FDA-approved for HER2-positive advanced or metastatic breast cancer with PIK3CA-mutated. Alpelisib has been experimentally used in solid tumors with a favorable safety profile [109,124].
However, there is limited data on using PIK3CA inhibitors such as Copanlisib, Idelalisib, Umbralisib, Duvelisib, and Alpelisib in NSCLC patients. There is a need to elucidate the underlying molecular biology to detect the relevant biomarkers of toxicities and resistance to PI3K therapies. Several NGS-based panels related to cancer were approved by the FDA. Oncomine Dx Target Test detects variations in 23 genes that are biomarkers for selecting NSCLC patients for four targeted therapies Dabrafenib, Crizotinib, Gefitinib, and Pralsetinib [125]. The Praxis Extended RAS Panel detects several mutations in exons 2, 3, and 4 of KRAS and NRAS to implicate metastatic colorectal cancer treatment decisions [126]. MSK-IMPACT was created by the Memorial Sloan Kettering Cancer Center (MSK) to detect genetic aberrations in the 468-gene in solid tumors [127]. The FoundationOne CDx diagnostic test detects genetic mutations in 324 genes in several FDA-approved therapies for solid tumors [128]. Ongoing clinical trials are summarized in Table 4.
Of note, Crizotinib is an ALK, MET, and ROS1 kinases inhibitor. The phase I study of Crizotinib in 50 patients who were positive for ROS1 rearrangement proved the antitumor activity of this drug in advanced NSCLC [129]. Consequently, in 2016, Crizotinib was approved by the U.S. Food and Drug Administration (FDA) for the treatment of advanced NSCLC patients with ROS1-rearranged. In a phase II study of Crizotinib in East Asian patients, 127 East Asian patients were enrolled with ROS1-positive. The ORR by IRR was 71.7% (95% CI, 63.0% to 79.3) for NSCLC patients with ROS1-positive and had lower rates of brain metastases compared to ALK rearrangements (ROS1 19.4%, ALK 39.1%; p = 0.033) [130]. Another study identified 33 ROS1 rearrangements in 579 patients with stage IV NSCLC. The median PFS time for ROS1-positive was 11 months. ROS1 fusion partners were CD74, SLC34A2, and ZCCHC8 [131]. In Chinese patients with ROS1-positive advanced NSCLC treated with Crizotinib, the objective response rate (ORR) and the disease control rate (DCR) were 71.4% and 94.3%, respectively. In addition, the median PFS was 11.0 months (95% confidence interval (CI), 7.8–14.2), and median OS was 41.0 months (95% CI, 22.5–59.5) [132].

3.1.2. Current Status of Novel Biomarkers for Response to Immunotherapy

Immune checkpoint inhibitors (ICIs) are considered one of the most important treatments for various advanced malignancies and have become the standard of care in lung cancer [133,134]. Among ICI immunotherapy agents, the antiprogrammed death 1 ligand (PD-L1) (atezolizumab and durvalumab), the cytotoxic T-lymphocyte antigen 4 (anti-CTLA-4) (ipilimumab), and the programmed cell death receptor-1 (PD-1) (nivolumab and pembrolizumab) are currently used in therapy in patients with lung cancer. Treatment outcomes of lung cancer patients have improved considerably with these drugs; however, these drugs can cause immune-related adverse events (irAEs) that threaten the life of patients. Moreover, only a relatively small number of lung cancer patients benefit from immunotherapy, and the rapid tumor progression after the treatment with ICIs defined is as a hyperprogressive disease (HPD). Hence, an urgent need for biomarkers that predict responses and prognosis after treatment with ICIs is warranted.
Previous large clinical trials reported that the higher expression of PD-L1 can predict the response to PD1/PD-L1 inhibitors in different types of cancer, such as non-small-cell lung cancer (NSCLC), urothelial cancer, and melanoma [132,135,136,137]. Immunohistochemical (IHC) detection of PD-L1 is a current predictive biomarker, but it is an imperfect biomarker for responses to ICI therapy.
In the KEYNOTE-001 trial, patients with advanced NSCLC enrolled in a phase I study and received Pembrolizumab. The PD-L1 expression in the tumor samples using the IHC analysis and results declared that at least 50% of the PD-L1 expression in the tumor cells correlated with improved efficacy of Pembrolizumab in patients [138]. In the open-label, phase II randomized controlled POPLAR trial, improvement of NSCLC patients treated with Atezolizumab was correlated with PD-L1 expression [139]. Data from the KEYNOTE-024 study reported that advanced NSCLC patients with PD-L1 expression on at least 50% of tumor cells treated with Pembrolizumab were associated with significantly longer PFS and OS [136]. However, there are some challenging obstacles to consistency in PD-L1 testing, including sampling methods, different techniques used to assess PDL1 expression, dynamic changes in PD-L1 expression, and spatial and temporal heterogeneity [122]. In addition, the single biomarker may not be indicative of patient selection. Therefore, several predictive and prognostic biomarkers have been explored during the past decade to guide patient choice.

3.1.3. Tumor Mutation Burden (TMB) and Circulating Tumor DNA (ctDNA)

The tumor mutation burden (TMB) refers to the total number of somatic mutations in tumor cells. The first association between a high TMB and treatment response was reported in patients with melanoma treated with cytotoxic T-lymphocyte-associated protein 4 (CTLA-4) [140]. Recently, various clinical trials showed that TMB is an important biomarker for predicting the efficacy of immunotherapy across diverse tumor types, including lung, urothelial, and breast cancers [141,142]. New studies have shown that blood-based TMB (bTMB) from circulating tumor DNA (ctDNA) has predictive power for the clinical benefit of immunotherapy in non-NSCLC patients receiving immunotherapy [142,143]. The role of bTMB profiling and sequencing of small amounts of cell-free DNA (cfDNA) in NSCLC patients treated with ICIs has been explored in several studies. Gandara et al. tested a novel assay to measure bTMB using samples collected prospectively from the POPLAR (NCT01903993) and the OAK (NCT02008227) clinical trials. Their results indicated that high bTMB in patients receiving Atezolizumab monotherapy in NSCLC was associated with longer PFS [143]. Wang et al. explored a 150-cancer genes panel (CGP) named NCC-GP150 in blood and tissue samples from non-small-cell lung cancer treated with anti-PD-1 and anti-PD-L1 therapy. Data from this analysis demonstrated that bTMB was a potential biomarker for patients with advanced NSCLC receiving ICIs [144]. Iijima et al. studied early response in 14 NSCLC patients treated with Nivolumab. The high level of tumor burden in circulating tumor DNA (ctDNA) could predict a durable response to Nivolumab [145]. Raja et al. studied variant allele frequencies (VAF) of somatic mutations in 73 genes ctDNA of NSCLC and urothelial cancer patients. Patients with the reduction in ctDNA VAF had a more significant decrease in tumor volume, with longer PFS and OS. [146]. In the study of Guibert et al., targeted sequencing of plasma ctDNA was analyzed in 97 progressive NSCLC patients. This study revealed that if a ctDNA allele fraction (A.F.) decreases, the PFS of patients increases. B-F1RST is the first prospective trial that evaluated bTMB in advanced NSCLC treated with Atezolizumab as a first-line (1L) monotherapy. The final result of this trial confirmed that bTMB could be considered a predictive biomarker for PFS and OS in NSCLC patients [146,147].
Atezolizumab monotherapy resulted in longer overall survival in NSCLC patients with high PD-L1 expression [148]. In the phase II B-F1RST (NCT02848651) trial, Kim et al. evaluated bTMB in locally advanced or metastatic stage IIIB–IVB NSCLC patients treated with first-line Atezolizumab monotherapy. In patients with bTMB ≥ 16, OS were associated with longer than patients with bTMB < 16 [142].
In meta-analyses of Meng et al., predictive value of TMB in NSCLC patients treated with ICIs were analyzed in six randomized controlled trials (3662 patients) and 31 datasets (3437 patients). Longer PFS, OS, and higher ORR were observed in high-TMB patients. Furthermore, immunotherapy was associated with improved PFS, OS, and ORR in patients with high-TMB patients [149].
In the systematic review and meta-analysis of Kim et al., they explored the data of 5712 patients from 26 studies. The result of this study showed that, in patients who were treated with ICIs, the high-TMB groups showed better OS and PFS compared to the low-TMB groups. Moreover, patients with high TMB benefit more from ICI treatment than chemotherapy alone. In NSCLC patients, higher TMB was significantly associated with better PFS [150].
In the study of Thompson et al., ctDNA NGS was identified 50 drivers and 12 resistance mutations in advanced NSCLC. Detection of targetable driver and resistance mutations in ctDNA led to the identification of the resistance mechanisms in patients and to finding additional therapeutic options [151].

3.2. Proteomics

Protein markers can monitor treatment and/or predict drug resistance in NSCLC patients. Xu et al. utilized a SILAC-based quantitative proteomic approach to elucidate the Paclitaxel (PTX) resistance. They found that PDCD4 mediated PTX sensitivity in cancer cell lines, and during adjuvant therapy of lung cancer patients with PTX, the overall survival of patients correlated with high levels of PDCD4 [152]. Programmed cell death 4 (PDCD4) is downregulated in different cancer types and is considered a tumor suppressor. It has been proved that Pdcd4 has various functions inhibiting cell growth, suppressing tumor promotion, metastasis, and inducing apoptosis. For instance, the loss of PDCD4 expression is correlated with tumor progression and prognosis and survival in lung, breast, ovarian, and colorectal cancers. In lung cancer cells, Vihreva et al. demonstrated that the transcription of PDCD4 is negatively regulated by mTOR signaling.
Sandfeld-Paulsen et al. studied exosomal membrane proteins by using the extracellular vesicle array in 276 NSCLC patients. They found increasing concentration levels of NY-ESO-1, EGFR, PLAP, EpCam, and Alix have been associated with inferior OS [153]. In the same year, they published another study isolating exosomes from the blood of 581 patients (431 with lung cancer and 150 controls). They found that markers CD151, CD171, and tetraspanin eight could significantly determine NSCLC patients versus healthy controls [154]. In the PROSE study, Gregorc et al. demonstrated that a serum protein test could predict survival in non-small-cell lung cancer treated with Erlotinib and single-agent chemotherapy [155]. In the study by Salmon et al., L.C. patients were treated with Erlotinib in combination with Bevacizumab, and matrix-assisted laser desorption ionization time-of-flight mass spectrometry (MALDI-TOF MS) could classify patients who benefit from treatment with Erlotinib [155,156]. VeriStrat is a plasma- or serum-based test utilizing MALDI MS methods to monitor EGFR TKIs treatment of various tumor types, including colorectal and head and neck cancer [157]. Multiple studies demonstrated that the blood-based proteomic VeriStrat test could help predict EGFR TKI treatment response. In lung cancer, Taguchi et al. analyzed serum proteins by MALDI MS in NSCLC patients treated with EGFR TKIs. Proteomic evaluation of this study classified patients for good or poor outcomes after treatment with Gefitinib or Erlotinib [157,158]. Plasma samples of 111 NSCLC patients were analyzed by MALDI-TOF MS and ranked by the VeriStrat test [159]. Patients in the VeriStrat “good” classification had longer PFS and OS. Carbone et al. analyzed a plasma sample of 441 NSCLC patients in the NCIC Clinical Trials Group BR.21 trial. According to the results, the VeriStrat test predicts the objective response to Erlotinib and is a prognostic for both OS and PFS [160]. Fidler et al. studies the serum samples of NSCLC patients who received chemotherapy or Erlotinib. They measured 102 biomarkers by Luminex immunoassays and classified patients with the VeriStrat test. This study showed 27 and 16 biomarkers associated with OS and PFS, respectively [161]. In the study of Grossi et al., 481 NSCLC patients were treated with first-line platinum-based chemotherapy. In the group receiving VeriStrat, the PFS and OS were longer [161,162].

4. AI Machine Learning-Driven Discovery of Biomarkers for NSCLC

Robust clinical evaluation requires innovations, including technology platforms and artificial intelligence (AI). AI mimics the problem-solving and decision-making capabilities of the human mind [163]. Machine learning (ML) and its subfield deep learning are the main subsets of AI that can help biologists analyze vast amounts of data, leverage patterns in the data, and transform big data into actionable biomarkers [164]. AI revolutionized various areas in medicine, such as ophthalmology, radiology, and oncology [165,166,167]. Several medical algorithms detect CT-based lesions and are approved by the FDA, such as Arterys Oncology DL, Arterys MICA, and QuantX™, that improve radiologist performance [168,169]. AI affects different aspects of lung cancer, such as medical imaging for screening, early detection, characterizing lung cancers, treatment selection, and monitoring treatment response [170,171]. However, further improvement is required in big data interpretation and management, model fairness and interpretability, and generalizability.
Artificial intelligence has many applications in lung cancer [172]. Zhang et al. found that a combination of five urinary biomarkers discriminated lung cancer patients from healthy ones and differentiated lung cancer from other cancers [69]. Analysis of microbial reads of TCGA of 33 cancer types with the stochastic gradient-boosting machine (GBM) learning models successfully discriminated against cancer-free individuals, patients with cancer, and patients with multiple types of cancer [173]. In the study of Tirzïte et al., lung cancer patients’ exhaled breath samples were explored with logistic regression analysis (LRA). This model can discriminate lung cancer patients from noncancer patients [174]. It is trained on a deep convolutional neural network (inception v3) that could predict most STK11, EGFR, FAT1, SETBP1, KRAS, and TP53 from pathology images [175]. The PD-L1 expression level is a crucial biomarker for identifying responders and nonresponders to NSCLC patients treated with anti-PD-1/PD-L1 treatments. However, the accurate estimation of PD-L1 expression is challenging. Wang et al. proposed a deep learning model that predicted mutated EGFR from the wild EGFR patients and distinguished PD-L1-positive from PD-L1-negative patients through C.T. images [176]. Choi et al. developed an AI-powered tumor proportion score (TPS) analyzer to evaluate whole-slide images of 802 NSCLC. Using this model, they could detect PD-L1 expression in tumors and calculate TPS. Moreover, AI-assisted TPS reading predicted OS and PFS upon ICI treatment [177]. Cheng et al. developed deep learning (DL)-based AI model to analyze the expression of PD-L1 in 1288 lung cancer patients. They used three different AI models (M1, M2, and M3) assessed in both PDL1 (22C3) and PD-L1 (SP263) assays. Their models improved the evaluation of PD-L1 expression, and diagnostic results were consistent with the pathologist’s [177,178]. In the study of Wu et al., TPS of PD-L1 expression was also assessed in whole slide images (WSIs) of the 22c3 assay by the DL model. Their results also confirmed that AI-assisted diagnosis tests improve diagnostic repeatability and are a promising tool for improving the efficiency of clinical pathologists [179].

5. Conclusions and Future Directions

Cancer is a heterogeneous disease caused mainly by genomic aberrations. These aberrations may alter proteome, transcriptome, and metabolome pathways, which in turn become valuable biomarkers. Genomics and proteomics heterogeneity is the primary barrier that may be responsible for developing drug resistance. Using new techniques, such as artificial intelligence or single cell RNA sequencing, to solve old problems could solve old problems such as discovering biomarkers for diagnostic, prognostic, and predictive biomarkers to help guide clinicians to choose their medicine more precisely. These data layers integrate advanced computational methods to overview tumorigenesis comprehensively. Rethinking with new methods the resistance to checkpoint inhibitors (e.g., PD-1, PD-L1 or CTLA-4) or targeted therapies (e.g., PI3K/mROT pathways inhibitors) could help in selecting only patients who would benefit from these therapies in the future.
Solid high-throughput clinical outcomes are pivotal for clinical biomarkers discovery. Therefore, international collaboration with a large sample size of a wide variety of cancers and sharing of treatment data is imperative to develop more accurate biomarkers predicting the therapy response of lung cancer patients. Early detection and a proper treatment regimen prediction can ultimately aid clinicians in improving those patients’ survival rates.

Author Contributions

M.H.A. collected and wrote the drafting of the article. A.M.A., N.S. and R.R. performed the critical revision and verified the whole concept. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors whose names are listed certify that they have no affiliations in any organization or entity with any financial interest and nonfinancial interest in the subject matter or materials discussed in this manuscript.

References

  1. Kotwal, A.A.; Schonberg, M.A. Cancer Screening in the Elderly: A Review of Breast, Colorectal, Lung, and Prostate Cancer Screening. Cancer J. 2017, 23, 246–253. [Google Scholar] [CrossRef] [PubMed]
  2. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  3. Walters, S.; Maringe, C.; Coleman, M.P.; Peake, M.D.; Butler, J.; Young, N.; Bergström, S.; Hanna, L.; Jakobsen, E.; Kölbeck, K.; et al. Lung Cancer Survival and Stage at Diagnosis in Australia, Canada, Denmark, Norway, Sweden and the UK: A Population-Based Study, 2004–2007. Thorax 2013, 68, 551–564. [Google Scholar] [CrossRef] [Green Version]
  4. Flehinger, B.J.; Melamed, M.R.; Heelan, R.T.; McGinnis, C.M.; Zaman, M.B.; Martini, N. Accuracy of Chest Film Screening by Technologists in the New York Early Lung Cancer Detection Program. Am. J. Roentgenol. 1978, 131, 593–597. [Google Scholar] [CrossRef] [PubMed]
  5. Tockman, M.S. Survival and Mortality from Lung Cancer in a Screened Population. Chest 1986, 89, 324S–325S. [Google Scholar] [CrossRef]
  6. Henschke, C.I.; McCauley, D.I.; Yankelevitz, D.F.; Naidich, D.P.; McGuinness, G.; Miettinen, O.S.; Libby, D.M.; Pasmantier, M.W.; Koizumi, J.; Altorki, N.K.; et al. Early Lung Cancer Action Project: Overall Design and Findings from Baseline Screening. Lancet 1999, 354, 99–105. [Google Scholar] [CrossRef]
  7. The National Lung Screening Trial Research Team. Reduced Lung-Cancer Mortality with Low-Dose Computed Tomographic Screening. N. Engl. J. Med. 2011, 365, 395–409. [Google Scholar] [CrossRef] [Green Version]
  8. Wigle, D.A.; Jurisica, I.; Radulovich, N.; Pintilie, M.; Rossant, J.; Liu, N.; Lu, C.; Woodgett, J.; Seiden, I.; Johnston, M.; et al. Molecular Profiling of Non-Small Cell Lung Cancer and Correlation with Disease-Free Survival. Cancer Res. 2002, 62, 3005–3008. [Google Scholar]
  9. The Cancer Genome Atlas Research Network. Comprehensive Molecular Profiling of Lung Adenocarcinoma. Nature 2014, 511, 543–550. [Google Scholar] [CrossRef] [Green Version]
  10. Wu, F.; Fan, J.; He, Y.; Xiong, A.; Yu, J.; Li, Y.; Zhang, Y.; Zhao, W.; Zhou, F.; Li, W.; et al. Single-Cell Profiling of Tumor Heterogeneity and the Microenvironment in Advanced Non-Small Cell Lung Cancer. Nat. Commun. 2021, 12, 2540. [Google Scholar] [CrossRef]
  11. Rosell, R.; Moran, T.; Queralt, C.; Porta, R.; Cardenal, F.; Camps, C.; Majem, M.; Lopez-Vivanco, G.; Isla, D.; Provencio, M.; et al. Screening for Epidermal Growth Factor Receptor Mutations in Lung Cancer. N. Engl. J. Med. 2009, 361, 958–967. [Google Scholar] [CrossRef]
  12. Yoh, K.; Matsumoto, S.; Tsuchihara, K.; Kohno, T.; Ishii, G.; Tsuta, K.; Nishio, M.; Yamamoto, N.; Murakami, H.; Satouchi, M.; et al. The Lung Cancer Genomic Screening Project for Individualized Medicine in Japan (LC-SCRUM-Japan): Screening for RET and ROS1 Fusions in Advanced EGFR Mutation-Negative Nonsquamous Lung Cancer and Development of Molecular Targeted Therapy. J. Clin. Oncol. 2014, 32, 8055. [Google Scholar] [CrossRef]
  13. Kohno, T.; Nakaoku, T.; Tsuta, K.; Tsuchihara, K.; Matsumoto, S.; Yoh, K.; Goto, K. Beyond ALK-RET, ROS1 and Other Oncogene Fusions in Lung Cancer. Transl. Lung Cancer Res. 2015, 4, 156–164. [Google Scholar]
  14. Passaro, A.; Attili, I.; Rappa, A.; Vacirca, D.; Ranghiero, A.; Fumagalli, C.; Guarize, J.; Spaggiari, L.; de Marinis, F.; Barberis, M.; et al. Genomic Characterization of Concurrent Alterations in Non-Small Cell Lung Cancer (NSCLC) Harboring Actionable Mutations. Cancers 2021, 13, 2172. [Google Scholar] [CrossRef]
  15. Macklin, A.; Khan, S.; Kislinger, T. Recent Advances in Mass Spectrometry Based Clinical Proteomics: Applications to Cancer Research. Clin. Proteom. 2020, 17, 17. [Google Scholar] [CrossRef]
  16. Yu, L.; Shen, J.; Mannoor, K.; Guarnera, M.; Jiang, F. Identification of ENO1 as a Potential Sputum Biomarker for Early-Stage Lung Cancer by Shotgun Proteomics. Clin. Lung Cancer 2014, 15, 372–378.e1. [Google Scholar] [CrossRef] [Green Version]
  17. Baran, K.; Brzeziańska-Lasota, E. Proteomic Biomarkers of Non-Small Cell Lung Cancer Patients. Adv. Respir. Med. 2021, 89, 419–426. [Google Scholar] [CrossRef]
  18. Duma, N.; Santana-Davila, R.; Molina, J.R. Non-Small Cell Lung Cancer: Epidemiology, Screening, Diagnosis, and Treatment. Mayo Clin. Proc. 2019, 94, 1623–1640. [Google Scholar] [CrossRef]
  19. Sobhani, N.; Sirico, M.; Generali, D.; Zanconati, F.; Scaggiante, B. Circulating Cell-Free Nucleic Acids as Prognostic and Therapy Predictive Tools for Metastatic Castrate-Resistant Prostate Cancer. World J. Clin. Oncol. 2020, 11, 450–463. [Google Scholar] [CrossRef]
  20. Sobhani, N.; Generali, D.; Zanconati, F.; Bortul, M.; Scaggiante, B. Cell-Free DNA Integrity for the Monitoring of Breast Cancer: Future Perspectives? World J. Clin. Oncol. 2018, 9, 26–32. [Google Scholar] [CrossRef]
  21. Mondelo-Macía, P.; García-González, J.; León-Mateos, L.; Castillo-García, A.; López-López, R.; Muinelo-Romay, L.; Díaz-Peña, R. Current Status and Future Perspectives of Liquid Biopsy in Small Cell Lung Cancer. Biomedicines 2021, 9, 48. [Google Scholar] [CrossRef] [PubMed]
  22. Koch, A.; Joosten, S.C.; Feng, Z.; de Ruijter, T.C.; Draht, M.X.; Melotte, V.; Smits, K.M.; Veeck, J.; Herman, J.G.; Van Neste, L.; et al. Author Correction: Analysis of DNA Methylation in Cancer: Location Revisited. Nat. Rev. Clin. Oncol. 2018, 15, 467. [Google Scholar] [CrossRef] [PubMed]
  23. Gouil, Q.; Keniry, A. Latest Techniques to Study DNA Methylation. Essays Biochem. 2019, 63, 639–648. [Google Scholar] [PubMed]
  24. Maier, S.; Olek, A. Diabetes: A Candidate Disease for Efficient DNA Methylation Profiling. J. Nutr. 2002, 132, 2440S–2443S. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Richardson, B. DNA Methylation and Autoimmune Disease. Clin. Immunol. 2003, 109, 72–79. [Google Scholar] [CrossRef]
  26. Celarain, N.; Tomas-Roig, J. Aberrant DNA Methylation Profile Exacerbates Inflammation and Neurodegeneration in Multiple Sclerosis Patients. J. Neuroinflamm. 2020, 17, 21. [Google Scholar] [CrossRef]
  27. Paska, A.V.; Hudler, P. Aberrant Methylation Patterns in Cancer: A Clinical View. Biochem. Med. 2015, 25, 161–176. [Google Scholar] [CrossRef]
  28. Tavares, N.T.; Gumauskaitė, S.; Lobo, J.; Jerónimo, C.; Henrique, R. DNA Methylation Biomarkers for Prediction of Response to Platinum-Based Chemotherapy: Where Do We Stand? Cancers 2022, 14, 2918. [Google Scholar] [CrossRef]
  29. Liang, R.; Li, X.; Li, W.; Zhu, X.; Li, C. DNA Methylation in Lung Cancer Patients: Opening a “Window of Life” under Precision Medicine. Biomed. Pharmacother. 2021, 144, 112202. [Google Scholar] [CrossRef]
  30. Wang, Y.; Yu, Z.; Wang, T.; Zhang, J.; Hong, L.; Chen, L. Identification of Epigenetic Aberrant Promoter Methylation of RASSF1A in Serum DNA and Its Clinicopathological Significance in Lung Cancer. Lung Cancer 2007, 56, 289–294. [Google Scholar] [CrossRef]
  31. Schotten, L.M.; Darwiche, K.; Seweryn, M.; Yildiz, V.; Kneuertz, P.J.; Eberhardt, W.E.E.; Eisenmann, S.; Welter, S.; Sisson, B.E.; Pietrzak, M.; et al. DNA Methylation of PTGER4 in Peripheral Blood Plasma Helps to Distinguish between Lung Cancer, Benign Pulmonary Nodules and Chronic Obstructive Pulmonary Disease Patients. Eur. J. Cancer 2021, 147, 142–150. [Google Scholar] [CrossRef]
  32. Raos, D.; Ulamec, M.; Bojanac, A.K.; Bulic-Jakus, F.; Jezek, D.; Sincic, N. Epigenetically Inactivated RASSF1A as a Tumor Biomarker. Bosn. J. Basic Med. Sci. 2020, 21, 386–397. [Google Scholar] [CrossRef]
  33. Hulbert, A.; Jusue-Torres, I.; Stark, A.; Chen, C.; Rodgers, K.; Lee, B.; Griffin, C.; Yang, A.; Huang, P.; Wrangle, J.; et al. Early Detection of Lung Cancer Using DNA Promoter Hypermethylation in Plasma and Sputum. Clin. Cancer Res. 2017, 23, 1998–2005. [Google Scholar] [CrossRef] [Green Version]
  34. Ooki, A.; Maleki, Z.; Tsay, J.-C.J.; Goparaju, C.; Brait, M.; Turaga, N.; Nam, H.-S.; Rom, W.N.; Pass, H.I.; Sidransky, D.; et al. A Panel of Novel Detection and Prognostic Methylated DNA Markers in Primary Non–Small Cell Lung Cancer and Serum DNA. Clin. Cancer Res. 2017, 23, 7141–7152. [Google Scholar] [CrossRef] [Green Version]
  35. Liu, F.; Zhang, H.; Lu, S.; Wu, Z.; Zhou, L.; Cheng, Z.; Bai, Y.; Zhao, J.; Zhang, Q.; Mao, H. Quantitative Assessment of Gene Promoter Methylation in Non-small Cell Lung Cancer Using Methylation-sensitive High-resolution Melting. Oncol. Lett. 2018, 15, 7639–7648. [Google Scholar] [CrossRef] [Green Version]
  36. Wang, B.-H.; Li, Y.-Y.; Han, J.-Z.; Zhou, L.-Y.; Lv, Y.-Q.; Zhang, H.-L.; Zhao, L. Gene Methylation as a Powerful Biomarker for Detection and Screening of Non-Small Cell Lung Cancer in Blood. Oncotarget 2017, 8, 31692–31704. [Google Scholar] [CrossRef] [Green Version]
  37. Zhang, Y.; Wang, R.; Song, H.; Huang, G.; Yi, J.; Zheng, Y.; Wang, J.; Chen, L. Methylation of Multiple Genes as a Candidate Biomarker in Non-Small Cell Lung Cancer. Cancer Lett. 2011, 303, 21–28. [Google Scholar] [CrossRef]
  38. Ponomaryova, A.A.; Rykova, E.Y.; Cherdyntseva, N.V.; Skvortsova, T.E.; Dobrodeev, A.Y.; Zav’yalov, A.A.; Bryzgalov, L.O.; Tuzikov, S.A.; Vlassov, V.V.; Laktionov, P.P. Potentialities of Aberrantly Methylated Circulating DNA for Diagnostics and Post-Treatment Follow-up of Lung Cancer Patients. Lung Cancer 2013, 81, 397–403. [Google Scholar] [CrossRef]
  39. Kneip, C.; Schmidt, B.; Seegebarth, A.; Weickmann, S.; Fleischhacker, M.; Liebenberg, V.; Field, J.K.; Dietrich, D. SHOX2 DNA Methylation is a Biomarker for the Diagnosis of Lung Cancer in Plasma. J. Thorac. Oncol. 2011, 6, 1632–1638. [Google Scholar] [CrossRef] [Green Version]
  40. Powrózek, T.; Krawczyk, P.; Nicoś, M.; Kuźnar-Kamińska, B.; Batura-Gabryel, H.; Milanowski, J. Methylation of the DCLK1 Promoter Region in Circulating Free DNA and Its Prognostic Value in Lung Cancer Patients. Clin. Transl. Oncol. 2016, 18, 398–404. [Google Scholar] [CrossRef]
  41. Powrózek, T.; Krawczyk, P.; Kucharczyk, T.; Milanowski, J. Septin 9 Promoter Region Methylation in Free Circulating DNA—Potential Role in Noninvasive Diagnosis of Lung Cancer: Preliminary Report. Med. Oncol. 2014, 31, 917. [Google Scholar] [CrossRef] [Green Version]
  42. Diaz-Lagares, A.; Mendez-Gonzalez, J.; Hervas, D.; Saigi, M.; Pajares, M.J.; Garcia, D.; Crujerias, A.B.; Pio, R.; Montuenga, L.M.; Zulueta, J.; et al. A Novel Epigenetic Signature for Early Diagnosis in Lung Cancer. Clin. Cancer Res. 2016, 22, 3361–3371. [Google Scholar] [CrossRef]
  43. Lee, S.M.; Park, J.Y.; Kim, D.S. Methylation of TMEFF2 Gene in Tissue and Serum DNA from Patients with Non-Small Cell Lung Cancer. Mol. Cells 2012, 34, 171–176. [Google Scholar] [CrossRef]
  44. Ardekani, A.M.; Naeini, M.M. The Role of MicroRNAs in Human Diseases. Avicenna J. Med. Biotechnol. 2010, 2, 161–179. [Google Scholar]
  45. Peng, Y.; Croce, C.M. The Role of MicroRNAs in Human Cancer. Signal Transduct. Target. Ther. 2016, 1, 15004. [Google Scholar] [CrossRef] [Green Version]
  46. Sobhani, N.; Chahwan, R.; Roudi, R.; Morris, R.; Volinia, S.; Chai, D.; D’Angelo, A.; Generali, D. Predictive and Prognostic Value of Non-Coding RNA in Breast Cancer. Cancers 2022, 14, 2952. [Google Scholar] [CrossRef]
  47. Wang, W.; Li, X.; Liu, C.; Zhang, X.; Wu, Y.; Diao, M.; Tan, S.; Huang, S.; Cheng, Y.; You, T. MicroRNA-21 as a Diagnostic and Prognostic Biomarker of Lung Cancer: A Systematic Review and Meta-Analysis. Biosci. Rep. 2022, 42, BSR20211653. [Google Scholar] [CrossRef]
  48. Lu, S.; Kong, H.; Hou, Y.; Ge, D.; Huang, W.; Ou, J.; Yang, D.; Zhang, L.; Wu, G.; Song, Y.; et al. Two Plasma microRNA Panels for Diagnosis and Subtype Discrimination of Lung Cancer. Lung Cancer 2018, 123, 44–51. [Google Scholar] [CrossRef]
  49. Wang, P.; Yang, D.; Zhang, H.; Wei, X.; Ma, T.; Cheng, Z.; Hong, Q.; Hu, J.; Zhuo, H.; Song, Y.; et al. Early Detection of Lung Cancer in Serum by a Panel of MicroRNA Biomarkers. Clin. Lung Cancer 2015, 16, 313–319.e1. [Google Scholar] [CrossRef]
  50. Pan, J.; Zhou, C.; Zhao, X.; He, J.; Tian, H.; Shen, W.; Han, Y.; Chen, J.; Fang, S.; Meng, X.; et al. A Two-miRNA Signature (miR-33a-5p and miR-128-3p) in Whole Blood as Potential Biomarker for Early Diagnosis of Lung Cancer. Sci. Rep. 2018, 8, 16699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Fehlmann, T.; Kahraman, M.; Ludwig, N.; Backes, C.; Galata, V.; Keller, V.; Geffers, L.; Mercaldo, N.; Hornung, D.; Weis, T.; et al. Evaluating the Use of Circulating MicroRNA Profiles for Lung Cancer Detection in Symptomatic Patients. JAMA Oncol. 2020, 6, 714–723. [Google Scholar] [CrossRef] [PubMed]
  52. Khandelwal, A.; Seam, R.K.; Gupta, M.; Rana, M.K.; Prakash, H.; Vasquez, K.M.; Jain, A. Circulating Micro RNA-590-5p Functions as a Liquid Biopsy Marker in Non-small Cell Lung Cancer. Cancer Sci. 2020, 111, 826–839. [Google Scholar] [CrossRef] [PubMed]
  53. Sozzi, G.; Boeri, M.; Rossi, M.; Verri, C.; Suatoni, P.; Bravi, F.; Roz, L.; Conte, D.; Grassi, M.; Sverzellati, N.; et al. Clinical Utility of a Plasma-Based miRNA Signature Classifier Within Computed Tomography Lung Cancer Screening: A Correlative MILD Trial Study. J. Clin. Oncol. 2014, 32, 768–773. [Google Scholar] [CrossRef]
  54. Li, N.; Ma, J.; Guarnera, M.A.; Fang, H.; Cai, L.; Jiang, F. Digital PCR Quantification of miRNAs in Sputum for Diagnosis of Lung Cancer. J. Cancer Res. Clin. Oncol. 2014, 140, 145–150. [Google Scholar] [CrossRef] [Green Version]
  55. Roa, W.H.; Kim, J.O.; Razzak, R.; Du, H.; Guo, L.; Singh, R.; Gazala, S.; Ghosh, S.; Wong, E.; Joy, A.A.; et al. Sputum MicroRNA Profiling: A Novel Approach for the Early Detection of Non-Small Cell Lung Cancer. Clin. Investig. Med. 2012, 35, 271. [Google Scholar] [CrossRef] [Green Version]
  56. Arab, A.; Karimipoor, M.; Irani, S.; Kiani, A.; Zeinali, S.; Tafsiri, E.; Sheikhy, K. Potential Circulating miRNA Signature for Early Detection of NSCLC. Cancer Genet. 2017, 216–217, 150–158. [Google Scholar] [CrossRef]
  57. Li, W.; Wang, Y.; Zhang, Q.; Tang, L.; Liu, X.; Dai, Y.; Xiao, L.; Huang, S.; Chen, L.; Guo, Z.; et al. MicroRNA-486 as a Biomarker for Early Diagnosis and Recurrence of Non-Small Cell Lung Cancer. PLoS ONE 2015, 10, e0134220. [Google Scholar] [CrossRef] [Green Version]
  58. Leng, Q.; Lin, Y.; Jiang, F.; Lee, C.-J.; Zhan, M.; Fang, H.; Wang, Y.; Jiang, F. A Plasma miRNA Signature for Lung Cancer Early Detection. Oncotarget 2017, 8, 111902–111911. [Google Scholar] [CrossRef] [Green Version]
  59. Foss, K.M.; Sima, C.; Ugolini, D.; Neri, M.; Allen, K.E.; Weiss, G.J. miR-1254 and miR-574-5p: Serum-Based microRNA Biomarkers for Early-Stage Non-Small Cell Lung Cancer. J. Thorac. Oncol. 2011, 6, 482–488. [Google Scholar] [CrossRef] [Green Version]
  60. Tyers, M.; Mann, M. From Genomics to Proteomics. Nature 2003, 422, 193–197. [Google Scholar] [CrossRef]
  61. Yu, J.; Zhai, X.; Li, X.; Zhong, C.; Guo, C.; Yang, F.; Yuan, Y.; Zheng, S. Identification of MST1 as a Potential Early Detection Biomarker for Colorectal Cancer through a Proteomic Approach. Sci. Rep. 2017, 7, 14265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Petricoin, E.F.; Ardekani, A.M.; Hitt, B.A.; Levine, P.J.; Fusaro, V.A.; Steinberg, S.M.; Mills, G.B.; Simone, C.; Fishman, D.A.; Kohn, E.C.; et al. Use of Proteomic Patterns in Serum to Identify Ovarian Cancer. Lancet 2002, 359, 572–577. [Google Scholar] [CrossRef]
  63. Petricoin, E.F., 3rd; Ornstein, D.K.; Paweletz, C.P.; Ardekani, A.; Hackett, P.S.; Hitt, B.A.; Velassco, A.; Trucco, C.; Wiegand, L.; Wood, K.; et al. Serum Proteomic Patterns for Detection of Prostate Cancer. J. Natl. Cancer Inst. 2002, 94, 1576–1578. [Google Scholar] [CrossRef]
  64. Beretov, J.; Wasinger, V.C.; Millar, E.K.A.; Schwartz, P.; Graham, P.H.; Li, Y. Proteomic Analysis of Urine to Identify Breast Cancer Biomarker Candidates Using a Label-Free LC-MS/MS Approach. PLoS ONE 2015, 10, e0141876. [Google Scholar] [CrossRef]
  65. Zhou, L.; Beuerman, R.W. The Power of Tears: How Tear Proteomics Research Could Revolutionize the Clinic. Expert Rev. Proteom. 2017, 14, 189–191. [Google Scholar] [CrossRef] [Green Version]
  66. Erozenci, L.A.; Böttger, F.; Bijnsdorp, I.V.; Jimenez, C.R. Urinary Exosomal Proteins as (pan-)cancer Biomarkers: Insights from the Proteome. FEBS Lett. 2019, 593, 1580–1597. [Google Scholar] [CrossRef] [Green Version]
  67. Geyer, P.E.; Voytik, E.; Treit, P.V.; Doll, S.; Kleinhempel, A.; Niu, L.; Müller, J.B.; Buchholtz, M.; Bader, J.M.; Teupser, D.; et al. Plasma Proteome Profiling to Detect and Avoid Sample-related Biases in Biomarker Studies. EMBO Mol. Med. 2019, 11, e10427. [Google Scholar] [CrossRef]
  68. Youssef, O.; Sarhadi, V.K.; Armengol, G.; Piirilä, P.; Knuuttila, A.; Knuutila, S. Exhaled Breath Condensate as a Source of Biomarkers for Lung Carcinomas. A Focus on Genetic and Epigenetic Markers-A Mini-Review. Genes Chromosomes Cancer 2016, 55, 905–914. [Google Scholar] [CrossRef]
  69. Zhang, C.; Leng, W.; Sun, C.; Lu, T.; Chen, Z.; Men, X.; Wang, Y.; Wang, G.; Zhen, B.; Qin, J. Urine Proteome Profiling Predicts Lung Cancer from Control Cases and Other Tumors. EBioMedicine 2018, 30, 120–128. [Google Scholar] [CrossRef] [Green Version]
  70. Hsu, C.-H.; Hsu, C.-W.; Hsueh, C.; Wang, C.-L.; Wu, Y.-C.; Wu, C.-C.; Liu, C.-C.; Yu, J.-S.; Chang, Y.-S.; Yu, C.-J. Identification and Characterization of Potential Biomarkers by Quantitative Tissue Proteomics of Primary Lung Adenocarcinoma. Mol. Cell. Proteom. 2016, 15, 2396–2410. [Google Scholar] [CrossRef] [Green Version]
  71. Ortea, I.; Rodríguez-Ariza, A.; Chicano-Gálvez, E.; Arenas Vacas, M.S.; Jurado Gámez, B. Discovery of Potential Protein Biomarkers of Lung Adenocarcinoma in Bronchoalveolar Lavage Fluid by SWATH MS Data-Independent Acquisition and Targeted Data Extraction. J. Proteom. 2016, 138, 106–114. [Google Scholar] [CrossRef]
  72. Carvalho, A.S.; Cuco, C.M.; Lavareda, C.; Miguel, F.; Ventura, M.; Almeida, S.; Pinto, P.; de Abreu, T.T.; Rodrigues, L.V.; Seixas, S.; et al. Bronchoalveolar Lavage Proteomics in Patients with Suspected Lung Cancer. Sci. Rep. 2017, 7, srep42190. [Google Scholar] [CrossRef] [Green Version]
  73. Jin, Y.; Wang, J.; Ye, X.; Su, Y.; Yu, G.; Yang, Q.; Liu, W.; Yu, W.; Cai, J.; Chen, X.; et al. Identification of GlcNAcylated Alpha-1-Antichymotrypsin as an Early Biomarker in Human Non-Small-Cell Lung Cancer by Quantitative Proteomic Analysis with Two Lectins. Br. J. Cancer 2016, 114, 532–544. [Google Scholar] [CrossRef]
  74. Boccellino, M.; Pinto, F.; Ieluzzi, V.; Giovane, A.; Quagliuolo, L.; Fariello, C.; Coppola, M.; Carlucci, A.; Santini, M.; Ferati, K.; et al. Proteomics Analysis of Human Serum of Patients with Non-small-cell Lung Cancer Reveals Proteins as Diagnostic Biomarker Candidates. J. Cell. Physiol. 2019, 234, 23798–23806. [Google Scholar] [CrossRef] [PubMed]
  75. Zhou, M.; Kong, Y.; Wang, X.; Li, W.; Chen, S.; Wang, L.; Wang, C.; Zhang, Q. LC-MS/MS-Based Quantitative Proteomics Analysis of Different Stages of Non-Small-Cell Lung Cancer. Biomed. Res. Int. 2021, 2021, 5561569. [Google Scholar] [CrossRef] [PubMed]
  76. Codreanu, S.G.; Hoeksema, M.D.; Slebos, R.J.C.; Zimmerman, L.J.; Rahman, S.M.J.; Li, M.; Chen, S.-C.; Chen, H.; Eisenberg, R.; Liebler, D.C.; et al. Identification of Proteomic Features To Distinguish Benign Pulmonary Nodules from Lung Adenocarcinoma. J. Proteome Res. 2017, 16, 3266–3276. [Google Scholar] [CrossRef] [PubMed]
  77. Tomczak, K.; Czerwińska, P.; Wiznerowicz, M. Review The Cancer Genome Atlas (TCGA): An Immeasurable Source of Knowledge. Współczesna Onkol. 2015, 1A, 68–77. [Google Scholar] [CrossRef]
  78. Gao, J.; Aksoy, B.A.; Dogrusoz, U.; Dresdner, G.; Gross, B.; Sumer, S.O.; Sun, Y.; Jacobsen, A.; Sinha, R.; Larsson, E.; et al. Integrative Analysis of Complex Cancer Genomics and Clinical Profiles Using the cBioPortal. Sci. Signal. 2013, 6, pl1. [Google Scholar] [CrossRef] [Green Version]
  79. Li, J.; Lu, Y.; Akbani, R.; Ju, Z.; Roebuck, P.L.; Liu, W.; Yang, J.-Y.; Broom, B.M.; Verhaak, R.G.W.; Kane, D.W.; et al. TCPA: A Resource for Cancer Functional Proteomics Data. Nat. Methods 2013, 10, 1046–1047. [Google Scholar] [CrossRef] [Green Version]
  80. Forbes, S.A.; Bhamra, G.; Bamford, S.; Dawson, E.; Kok, C.; Clements, J.; Menzies, A.; Teague, J.W.; Futreal, P.A.; Stratton, M.R. The Catalogue of Somatic Mutations in Cancer (COSMIC). Curr. Protoc. Hum. Genet. 2008, 57, D941–D947. [Google Scholar] [CrossRef] [Green Version]
  81. Zhang, J.; Baran, J.; Cros, A.; Guberman, J.M.; Haider, S.; Hsu, J.; Liang, Y.; Rivkin, E.; Wang, J.; Whitty, B.; et al. International Cancer Genome Consortium Data Portal—A One-Stop Shop for Cancer Genomics Data. Database 2011, 2011, bar026. [Google Scholar] [CrossRef] [Green Version]
  82. ICGC/TCGA Pan-Cancer Analysis of Whole Genomes Consortium Pan-Cancer Analysis of Whole Genomes. Nature 2020, 578, 82–93. [CrossRef] [Green Version]
  83. Roudi, R.; Beikzadeh, B.; Roviello, G.; D’angelo, A.; Hadizadeh, M. Identification of Hub Genes, Modules and Biological Pathways Associated with Lung Adenocarcinoma: A System Biology Approach. Gene Rep. 2022, 27, 101638. [Google Scholar] [CrossRef]
  84. Brennan, P.; Hainaut, P.; Boffetta, P. Genetics of Lung-Cancer Susceptibility. Lancet Oncol. 2011, 12, 399–408. [Google Scholar] [CrossRef]
  85. Abbasian, M.H.; Abbasi, B.; Ansarinejad, N.; Ardekani, A.M.; Samizadeh, E.; Moghaddam, K.G.; Hashemi, M.R. Association of Interleukin-1 Gene Polymorphism with Risk of Gastric and Colorectal Cancers in an Iranian Population. Iran. J. Immunol. 2018, 15, 321–328. [Google Scholar]
  86. Abbasian, M.H.; Ansarinejad, N.; Abbasi, B.; Iravani, M.; Ramim, T.; Hamedi, F.; Ardekani, A.M. The Role of and Polymorphisms in Fluoropyrimidine-Based Cancer Chemotherapy in an Iranian Population. Avicenna J. Med. Biotechnol. 2020, 12, 157–164. [Google Scholar]
  87. Santarpia, M.; Rolfo, C.; Peters, G.J.; Leon, L.G.; Giovannetti, E. On the Pharmacogenetics of Non-Small Cell Lung Cancer Treatment. Expert Opin. Drug Metab. Toxicol. 2016, 12, 307–317. [Google Scholar] [CrossRef] [Green Version]
  88. Takeuchi, K.; Soda, M.; Togashi, Y.; Suzuki, R.; Sakata, S.; Hatano, S.; Asaka, R.; Hamanaka, W.; Ninomiya, H.; Uehara, H.; et al. RET, ROS1 and ALK Fusions in Lung Cancer. Nat. Med. 2012, 18, 378–381. [Google Scholar] [CrossRef]
  89. Rotow, J.; Bivona, T.G. Understanding and Targeting Resistance Mechanisms in NSCLC. Nat. Rev. Cancer 2017, 17, 637–658. [Google Scholar] [CrossRef]
  90. Fathi, Z.; Mousavi, S.A.J.; Roudi, R.; Ghazi, F. Distribution of KRAS, DDR2, and TP53 Gene Mutations in Lung Cancer: An Analysis of Iranian Patients. PLoS ONE 2018, 13, e0200633. [Google Scholar] [CrossRef]
  91. Roudi, R.; Haji, G.; Madjd, Z.; Shariftabrizi, A.; Mehrazma, M. Evaluation of Anaplastic Lymphoma Kinase Expression in Nonsmall Cell Lung Cancer; a Tissue Microarray Analysis. J. Cancer Res. Ther. 2016, 12, 1065–1069. [Google Scholar]
  92. Khalil, F.K.; Altiok, S. Advances in EGFR as a Predictive Marker in Lung Adenocarcinoma. Cancer Control 2015, 22, 193–199. [Google Scholar] [CrossRef] [Green Version]
  93. Fitzgerald, T.L.; Lertpiriyapong, K.; Cocco, L.; Martelli, A.M.; Libra, M.; Candido, S.; Montalto, G.; Cervello, M.; Steelman, L.; Abrams, S.L.; et al. Roles of EGFR and KRAS and Their Downstream Signaling Pathways in Pancreatic Cancer and Pancreatic Cancer Stem Cells. Adv. Biol. Regul. 2015, 59, 65–81. [Google Scholar] [CrossRef]
  94. Eskilsson, E.; Røsland, G.V.; Solecki, G.; Wang, Q.; Harter, P.N.; Graziani, G.; Verhaak, R.G.W.; Winkler, F.; Bjerkvig, R.; Miletic, H. EGFR Heterogeneity and Implications for Therapeutic Intervention in Glioblastoma. Neuro-Oncol. 2018, 20, 743–752. [Google Scholar] [CrossRef] [Green Version]
  95. Byeon, H.K.; Ku, M.; Yang, J. Beyond EGFR Inhibition: Multilateral Combat Strategies to Stop the Progression of Head and Neck Cancer. Exp. Mol. Med. 2019, 51, 1–14. [Google Scholar] [CrossRef] [Green Version]
  96. Khan, K.; Valeri, N.; Dearman, C.; Rao, S.; Watkins, D.; Starling, N.; Chau, I.; Cunningham, D. Targeting EGFR Pathway in Metastatic Colorectal Cancer- Tumour Heterogeniety and Convergent Evolution. Crit. Rev. Oncol. Hematol. 2019, 143, 153–163. [Google Scholar] [CrossRef]
  97. Maennling, A.E.; Tur, M.K.; Niebert, M.; Klockenbring, T.; Zeppernick, F.; Gattenlöhner, S.; Meinhold-Heerlein, I.; Hussain, A.F. Molecular Targeting Therapy against EGFR Family in Breast Cancer: Progress and Future Potentials. Cancers 2019, 11, 1826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Grandis, J.R.; Sok, J.C. Signaling through the Epidermal Growth Factor Receptor during the Development of Malignancy. Pharmacol. Ther. 2004, 102, 37–46. [Google Scholar] [CrossRef] [PubMed]
  99. Merrick, D.T.; Kittelson, J.; Winterhalder, R.; Kotantoulas, G.; Ingeberg, S.; Keith, R.L.; Kennedy, T.C.; Miller, Y.E.; Franklin, W.A.; Hirsch, F.R. Analysis of c-ErbB1/Epidermal Growth Factor Receptor and c-ErbB2/HER-2 Expression in Bronchial Dysplasia: Evaluation of Potential Targets for Chemoprevention of Lung Cancer. Clin. Cancer Res. 2006, 12, 2281–2288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Chan, B.A.; Hughes, B.G.M. Targeted Therapy for Non-Small Cell Lung Cancer: Current Standards and the Promise of the Future. Transl. Lung Cancer Res. 2015, 4, 36–54. [Google Scholar]
  101. Wu, S.-G.; Shih, J.-Y. Management of Acquired Resistance to EGFR TKI–targeted Therapy in Advanced Non-Small Cell Lung Cancer. Mol. Cancer 2018, 17, 38. [Google Scholar] [CrossRef] [Green Version]
  102. Stewart, E.L.; Mascaux, C.; Shakashita, S.; Panchal, D.; Wang, D.; Li, M.; Pham, N.-A.; Leighl, N.; Liu, G.; Shepherd, F.A.; et al. Abstract 1184: Modeling Mechanisms of Resistance of Epidermal Growth Factor Receptor (EGFR) Mutations to Targeted Drugs through Patient-Derived Xenografts (PDX) of Non-Small Cell Lung Cancer (NSCLC). Cancer Res. 2014, 74, 1184. [Google Scholar] [CrossRef]
  103. Morgillo, F.; Della Corte, C.M.; Fasano, M.; Ciardiello, F. Mechanisms of Resistance to EGFR-Targeted Drugs: Lung Cancer. ESMO Open 2016, 1, e000060. [Google Scholar] [CrossRef] [Green Version]
  104. Marchetti, A.; Felicioni, L.; Malatesta, S.; Sciarrotta, M.G.; Guetti, L.; Chella, A.; Viola, P.; Pullara, C.; Mucilli, F.; Buttitta, F. Clinical Features and Outcome of Patients with Non–Small-Cell Lung Cancer Harboring BRAF Mutations. J. Clin. Oncol. 2011, 29, 3574–3579. [Google Scholar] [CrossRef]
  105. Kim, H.C.; Kang, Y.R.; Ji, W.; Kim, Y.J.; Yoon, S.; Lee, J.C.; Choi, C.-M. Frequency and Clinical Features of BRAF Mutations among Patients with Stage III/IV Lung Adenocarcinoma without EGFR/ALK Aberrations. OncoTargets Ther. 2019, 12, 6045–6052. [Google Scholar] [CrossRef] [Green Version]
  106. Lee, J.; Lee, S.E.; Kang, S.Y.; Do, I.-G.; Lee, S.; Ha, S.Y.; Cho, J.; Kang, W.K.; Jang, J.; Ou, S.-H.I.; et al. Identification of ROS1 rearrangement in Gastric Adenocarcinoma. Cancer 2013, 119, 1627–1635. [Google Scholar] [CrossRef]
  107. Birch, A.H.; Arcand, S.L.; Oros, K.K.; Rahimi, K.; Watters, A.K.; Provencher, D.; Greenwood, C.M.; Mes-Masson, A.-M.; Tonin, P.N. Chromosome 3 Anomalies Investigated by Genome Wide SNP Analysis of Benign, Low Malignant Potential and Low Grade Ovarian Serous Tumours. PLoS ONE 2011, 6, e28250. [Google Scholar] [CrossRef]
  108. Aisner, D.L.; Nguyen, T.T.; Paskulin, D.D.; Le, A.T.; Haney, J.; Schulte, N.; Chionh, F.; Hardingham, J.; Mariadason, J.; Tebbutt, N.; et al. ROS1 and ALK Fusions in Colorectal Cancer, with Evidence of Intratumoral Heterogeneity for Molecular Drivers. Mol. Cancer Res. 2014, 12, 111–118. [Google Scholar] [CrossRef] [Green Version]
  109. Shapiro, G.I.; Rodon, J.; Bedell, C.; Kwak, E.L.; Baselga, J.; Braña, I.; Pandya, S.S.; Scheffold, C.; Douglas Laird, A.; Nguyen, L.T.; et al. Phase I Safety, Pharmacokinetic, and Pharmacodynamic Study of SAR245408 (XL147), an Oral Pan-Class I PI3K Inhibitor, in Patients with Advanced Solid Tumors. Clin. Cancer Res. 2014, 20, 233–245. [Google Scholar] [CrossRef] [Green Version]
  110. Marks, E.I.; Pamarthy, S.; Dizon, D.; Birnbaum, A.; Yakirevich, E.; Safran, H.; Carneiro, B.A. ROS1-GOPC/FIG: A Novel Gene Fusion in Hepatic Angiosarcoma. Oncotarget 2019, 10, 245–251. [Google Scholar] [CrossRef] [Green Version]
  111. Bergethon, K.; Shaw, A.T.; Ou, S.-H.I.; Katayama, R.; Lovly, C.M.; McDonald, N.T.; Massion, P.P.; Siwak-Tapp, C.; Gonzalez, A.; Fang, R.; et al. ROS1 Rearrangements Define a Unique Molecular Class of Lung Cancers. J. Clin. Oncol. 2012, 30, 863–870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Yoshida, A.; Kohno, T.; Tsuta, K.; Wakai, S.; Arai, Y.; Shimada, Y.; Asamura, H.; Furuta, K.; Shibata, T.; Tsuda, H. ROS1-Rearranged Lung Cancer. Am. J. Surg. Pathol. 2013, 37, 554–562. [Google Scholar] [CrossRef] [PubMed]
  113. Yu, Z.-Q.; Wang, M.; Zhou, W.; Mao, M.-X.; Chen, Y.-Y.; Li, N.; Peng, X.-C.; Cai, J.; Cai, Z.-Q. ROS1-Positive Non-Small Cell Lung Cancer (NSCLC): Biology, Diagnostics, Therapeutics and Resistance. J. Drug Target. 2022, 30, 845–857. [Google Scholar] [CrossRef] [PubMed]
  114. Polivka, J.; Janku, F. Molecular Targets for Cancer Therapy in the PI3K/AKT/mTOR Pathway. Pharmacol. Ther. 2014, 142, 164–175. [Google Scholar] [CrossRef]
  115. Yu, L.; Wei, J.; Liu, P. Attacking the PI3K/Akt/mTOR Signaling Pathway for Targeted Therapeutic Treatment in Human Cancer. In Seminars in Cancer Biology; Academic Press: Cambridge, MA, USA, 2021. [Google Scholar]
  116. Kempf, E.; Rousseau, B.; Besse, B.; Paz-Ares, L. KRAS oncogene in Lung Cancer: Focus on Molecularly Driven Clinical Trials. Eur. Respir. Rev. 2016, 25, 71–76. [Google Scholar] [CrossRef] [Green Version]
  117. Bockorny, B.; Rusan, M.; Chen, W.; Liao, R.G.; Li, Y.; Piccioni, F.; Wang, J.; Tan, L.; Thorner, A.R.; Li, T.; et al. RAS–MAPK Reactivation Facilitates Acquired Resistance in FGFR1-Amplified Lung Cancer and Underlies a Rationale for Upfront FGFR–MEK Blockade. Mol. Cancer Ther. 2018, 17, 1526–1539. [Google Scholar] [CrossRef] [Green Version]
  118. Iksen; Pothongsrisit, S.; Pongrakhananon, V. Targeting the PI3K/AKT/mTOR Signaling Pathway in Lung Cancer: An Update Regarding Potential Drugs and Natural Products. Molecules 2021, 26, 4100. [Google Scholar] [CrossRef]
  119. Scheffler, M.; Bos, M.; Gardizi, M.; König, K.; Michels, S.; Fassunke, J.; Heydt, C.; Künstlinger, H.; Ihle, M.; Ueckeroth, F.; et al. PIK3CA Mutations in Non-Small Cell Lung Cancer (NSCLC): Genetic Heterogeneity, Prognostic Impact and Incidence of Prior Malignancies. Oncotarget 2015, 6, 1315–1326. [Google Scholar] [CrossRef] [Green Version]
  120. Yamamoto, H.; Shigematsu, H.; Nomura, M.; Lockwood, W.W.; Sato, M.; Okumura, N.; Soh, J.; Suzuki, M.; Wistuba, I.I.; Fong, K.M.; et al. PIK3CA Mutations and Copy Number Gains in Human Lung Cancers. Cancer Res. 2008, 68, 6913–6921. [Google Scholar] [CrossRef] [Green Version]
  121. Le, X.; Puri, S.; Negrao, M.V.; Nilsson, M.B.; Robichaux, J.; Boyle, T.; Hicks, J.K.; Lovinger, K.L.; Roarty, E.; Rinsurongkawong, W.; et al. Landscape of EGFR-Dependent and -Independent Resistance Mechanisms to Osimertinib and Continuation Therapy Beyond Progression in EGFR-Mutant NSCLC. Clin. Cancer Res. 2018, 24, 6195–6203. [Google Scholar] [CrossRef] [Green Version]
  122. Wang, M.; Wang, S.; Trapani, J.A.; Neeson, P.J. Challenges of PD-L1 Testing in Non-Small Cell Lung Cancer and beyond. J. Thorac. Dis. 2020, 12, 4541–4548. [Google Scholar] [CrossRef]
  123. Song, Z.; Yu, X.; Zhang, Y. Mutation and Prognostic Analyses of PIK 3 CA in Patients with Completely Resected Lung Adenocarcinoma. Cancer Med. 2016, 5, 2694–2700. [Google Scholar] [CrossRef]
  124. Ando, Y.; Iwasa, S.; Takahashi, S.; Saka, H.; Kakizume, T.; Natsume, K.; Suenaga, N.; Quadt, C.; Yamada, Y. Phase I Study of Alpelisib (BYL719), an α-Specific PI3K Inhibitor, in Japanese Patients with Advanced Solid Tumors. Cancer Sci. 2019, 110, 1021–1031. [Google Scholar] [CrossRef] [Green Version]
  125. Yu, T.M.; Morrison, C.; Gold, E.J.; Tradonsky, A.; Layton, A.J. Multiple Biomarker Testing Tissue Consumption and Completion Rates With Single-Gene Tests and Investigational Use of Oncomine Dx Target Test for Advanced Non–Small-Cell Lung Cancer: A Single-Center Analysis. Clin. Lung Cancer 2019, 20, 20–29.e8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Udar, N.; Iyer, A.; Porter, M.; Haigis, R.; Smith, S.; Dhillon, S.; Meier, K.; Ward, D.; Lu, J.; Wenz, P.; et al. Development and Analytical Validation of a DNA Dual-Strand Approach for the US Food and Drug Administration–Approved Next-Generation Sequencing–Based Praxis Extended RAS Panel for Metastatic Colorectal Cancer Samples. J. Mol. Diagn. 2020, 22, 159–178. [Google Scholar] [CrossRef] [PubMed]
  127. Cheng, D.T.; Mitchell, T.N.; Zehir, A.; Shah, R.H.; Benayed, R.; Syed, A.; Chandramohan, R.; Liu, Z.Y.; Won, H.H.; Scott, S.N.; et al. Memorial Sloan Kettering-Integrated Mutation Profiling of Actionable Cancer Targets (MSK-IMPACT). J. Mol. Diagn. 2015, 17, 251–264. [Google Scholar] [CrossRef] [PubMed]
  128. Takeda, M.; Takahama, T.; Sakai, K.; Shimizu, S.; Watanabe, S.; Kawakami, H.; Tanaka, K.; Sato, C.; Hayashi, H.; Nonagase, Y.; et al. Clinical Application of the FoundationOne CDx Assay to Therapeutic Decision-Making for Patients with Advanced Solid Tumors. Oncologist 2021, 26, e588–e596. [Google Scholar] [CrossRef] [PubMed]
  129. Shaw, A.T.; Ou, S.-H.I.; Bang, Y.-J.; Camidge, D.R.; Solomon, B.J.; Salgia, R.; Riely, G.J.; Varella-Garcia, M.; Shapiro, G.I.; Costa, D.B.; et al. Crizotinib in ROS1-Rearranged Non–Small-Cell Lung Cancer. N. Engl. J. Med. 2014, 371, 1963–1971. [Google Scholar] [CrossRef] [Green Version]
  130. Wu, Y.-L.; Yang, J.C.-H.; Kim, D.-W.; Lu, S.; Zhou, J.; Seto, T.; Yang, J.-J.; Yamamoto, N.; Ahn, M.-J.; Takahashi, T.; et al. Phase II Study of Crizotinib in East Asian Patients with ROS1-Positive Advanced Non–Small-Cell Lung Cancer. J. Clin. Oncol. 2018, 36, 1405–1411. [Google Scholar] [CrossRef]
  131. Patil, T.; Smith, D.E.; Bunn, P.A.; Aisner, D.L.; Le, A.T.; Hancock, M.; Purcell, W.T.; Bowles, D.W.; Camidge, D.R.; Doebele, R.C. The Incidence of Brain Metastases in Stage IV ROS1-Rearranged Non–Small Cell Lung Cancer and Rate of Central Nervous System Progression on Crizotinib. J. Thorac. Oncol. 2018, 13, 1717–1726. [Google Scholar] [CrossRef] [Green Version]
  132. Liu, C.; Yu, H.; Chang, J.; Chen, H.; Li, Y.; Zhao, W.; Zhao, K.; Zhu, Z.; Sun, S.; Fan, M.; et al. Crizotinib in Chinese Patients with ROS1-Rearranged Advanced Non-Small-Cell Lung Cancer in Routine Clinical Practice. Target. Oncol. 2019, 14, 315–323. [Google Scholar] [CrossRef] [Green Version]
  133. Tartarone, A.; Roviello, G.; Lerose, R.; Roudi, R.; Aieta, M.; Zoppoli, P. Anti-PD-1 versus Anti-PD-L1 Therapy in Patients with Pretreated Advanced Non-Small-Cell Lung Cancer: A Meta-Analysis. Future Oncol. 2019, 15, 2423–2433. [Google Scholar] [CrossRef]
  134. Petrelli, F.; Ferrara, R.; Signorelli, D.; Ghidini, A.; Proto, C.; Roudi, R.; Sabet, M.N.; Facelli, S.; Garassino, M.C.; Luciani, A.; et al. Immune Checkpoint Inhibitors and Chemotherapy in First-Line NSCLC: A Meta-Analysis. Immunotherapy 2021, 13, 621–631. [Google Scholar] [CrossRef]
  135. Taube, J.M.; Klein, A.; Brahmer, J.R.; Xu, H.; Pan, X.; Kim, J.H.; Chen, L.; Pardoll, D.M.; Topalian, S.L.; Anders, R.A. Association of PD-1, PD-1 Ligands, and Other Features of the Tumor Immune Microenvironment with Response to Anti–PD-1 Therapy. Clin. Cancer Res. 2014, 20, 5064–5074. [Google Scholar] [CrossRef] [Green Version]
  136. Reck, M.; Rodríguez-Abreu, D.; Robinson, A.G.; Hui, R.; Csőszi, T.; Fülöp, A.; Gottfried, M.; Peled, N.; Tafreshi, A.; Cuffe, S.; et al. Pembrolizumab versus Chemotherapy for PD-L1–Positive Non–Small-Cell Lung Cancer. N. Engl. J. Med. 2016, 375, 1823–1833. [Google Scholar] [CrossRef] [Green Version]
  137. Balar, A.V.; Castellano, D.; O’Donnell, P.H.; Grivas, P.; Vuky, J.; Powles, T.; Plimack, E.R.; Hahn, N.M.; de Wit, R.; Pang, L.; et al. First-Line Pembrolizumab in Cisplatin-Ineligible Patients with Locally Advanced and Unresectable or Metastatic Urothelial Cancer (KEYNOTE-052): A Multicentre, Single-Arm, Phase 2 Study. Lancet Oncol. 2017, 18, 1483–1492. [Google Scholar] [CrossRef]
  138. Garon, E.B.; Rizvi, N.A.; Hui, R.; Leighl, N.; Balmanoukian, A.S.; Eder, J.P.; Patnaik, A.; Aggarwal, C.; Gubens, M.; Horn, L.; et al. Pembrolizumab for the Treatment of Non–Small-Cell Lung Cancer. N. Engl. J. Med. 2015, 372, 2018–2028. [Google Scholar] [CrossRef]
  139. Fehrenbacher, L.; Spira, A.; Ballinger, M.; Kowanetz, M.; Vansteenkiste, J.; Mazieres, J.; Park, K.; Smith, D.; Artal-Cortes, A.; Lewanski, C.; et al. Atezolizumab versus Docetaxel for Patients with Previously Treated Non-Small-Cell Lung Cancer (POPLAR): A Multicentre, Open-Label, Phase 2 Randomised Controlled Trial. Lancet 2016, 387, 1837–1846. [Google Scholar] [CrossRef]
  140. Snyder, A.; Makarov, V.; Merghoub, T.; Yuan, J.; Zaretsky, J.M.; Desrichard, A.; Walsh, L.A.; Postow, M.A.; Wong, P.; Ho, T.S.; et al. Genetic Basis for Clinical Response to CTLA-4 Blockade in Melanoma. N. Engl. J. Med. 2014, 371, 2189–2199. [Google Scholar] [CrossRef] [Green Version]
  141. Thomas, A.; Routh, E.D.; Pullikuth, A.; Jin, G.; Su, J.; Chou, J.W.; Hoadley, K.A.; Print, C.; Knowlton, N.; Black, M.A.; et al. Tumor Mutational Burden is a Determinant of Immune-Mediated Survival in Breast Cancer. OncoImmunology 2018, 7, e1490854. [Google Scholar] [CrossRef]
  142. Kim, E.S.; Velcheti, V.; Mekhail, T.; Yun, C.; Shagan, S.M.; Hu, S.; Chae, Y.K.; Leal, T.A.; Dowell, J.E.; Tsai, M.L.; et al. Blood-Based Tumor Mutational Burden as a Biomarker for Atezolizumab in Non-Small Cell Lung Cancer: The Phase 2 B-F1RST Trial. Nat. Med. 2022, 28, 939–945. [Google Scholar] [CrossRef]
  143. Gandara, D.R.; Paul, S.M.; Kowanetz, M.; Schleifman, E.; Zou, W.; Li, Y.; Rittmeyer, A.; Fehrenbacher, L.; Otto, G.; Malboeuf, C.; et al. Blood-Based Tumor Mutational Burden as a Predictor of Clinical Benefit in Non-Small-Cell Lung Cancer Patients Treated with Atezolizumab. Nat. Med. 2018, 24, 1441–1448. [Google Scholar] [CrossRef]
  144. Wang, Z.; Duan, J.; Cai, S.; Han, M.; Dong, H.; Zhao, J.; Zhu, B.; Wang, S.; Zhuo, M.; Sun, J.; et al. Assessment of Blood Tumor Mutational Burden as a Potential Biomarker for Immunotherapy in Patients With Non–Small Cell Lung Cancer With Use of a Next-Generation Sequencing Cancer Gene Panel. JAMA Oncol. 2019, 5, 696–702. [Google Scholar] [CrossRef]
  145. Iijima, Y.; Hirotsu, Y.; Amemiya, K.; Ooka, Y.; Mochizuki, H.; Oyama, T.; Nakagomi, T.; Uchida, Y.; Kobayashi, Y.; Tsutsui, T.; et al. Very Early Response of Circulating Tumour–derived DNA in Plasma Predicts Efficacy of Nivolumab Treatment in Patients with Non–small Cell Lung Cancer. Eur. J. Cancer 2017, 86, 349–357. [Google Scholar] [CrossRef]
  146. Raja, R.; Kuziora, M.; Brohawn, P.Z.; Higgs, B.W.; Gupta, A.; Dennis, P.A.; Ranade, K. Early Reduction in ctDNA Predicts Survival in Patients with Lung and Bladder Cancer Treated with Durvalumab. Clin. Cancer Res. 2018, 24, 6212–6222. [Google Scholar] [CrossRef] [Green Version]
  147. Socinski, M.; Velcheti, V.; Mekhail, T.; Chae, Y.K.; Leal, T.A.; Dowell, J.E.; Tsai, M.L.; Dakhil, C.S.; Stella, P.; Shen, V.; et al. Final Efficacy Results from B-F1RST, a Prospective Phase II Trial Evaluating Blood-Based Tumour Mutational Burden (bTMB) as a Predictive Biomarker for Atezolizumab (atezo) in 1L Non-Small Cell Lung Cancer (NSCLC). Ann. Oncol. 2019, 30, v919–v920. [Google Scholar] [CrossRef]
  148. Herbst, R.S.; Giaccone, G.; de Marinis, F.; Reinmuth, N.; Vergnenegre, A.; Barrios, C.H.; Morise, M.; Felip, E.; Andric, Z.; Geater, S.; et al. Atezolizumab for First-Line Treatment of PD-L1-Selected Patients with NSCLC. N. Engl. J. Med. 2020, 383, 1328–1339. [Google Scholar] [CrossRef]
  149. Meng, G.; Liu, X.; Ma, T.; Lv, D.; Sun, G. Predictive Value of Tumor Mutational Burden for Immunotherapy in Non-Small Cell Lung Cancer: A Systematic Review and Meta-Analysis. PLoS ONE 2022, 17, e0263629. [Google Scholar] [CrossRef]
  150. Kim, J.Y.; Kronbichler, A.; Eisenhut, M.; Hong, S.H.; van der Vliet, H.J.; Kang, J.; Shin, J.I.; Gamerith, G. Tumor Mutational Burden and Efficacy of Immune Checkpoint Inhibitors: A Systematic Review and Meta-Analysis. Cancers 2019, 11, 1798. [Google Scholar] [CrossRef] [Green Version]
  151. Thompson, J.C.; Yee, S.S.; Troxel, A.B.; Savitch, S.L.; Fan, R.; Balli, D.; Lieberman, D.B.; Morrissette, J.D.; Evans, T.L.; Bauml, J.; et al. Detection of Therapeutically Targetable Driver and Resistance Mutations in Lung Cancer Patients by Next-Generation Sequencing of Cell-Free Circulating Tumor DNA. Clin. Cancer Res. 2016, 22, 5772–5782. [Google Scholar] [CrossRef] [Green Version]
  152. Xu, H.; Dephoure, N.; Sun, H.; Zhang, H.; Fan, F.; Liu, J.; Ning, X.; Dai, S.; Liu, B.; Gao, M.; et al. Proteomic Profiling of Paclitaxel Treated Cells Identifies a Novel Mechanism of Drug Resistance Mediated by PDCD4. J. Proteome Res. 2015, 14, 2480–2491. [Google Scholar] [CrossRef] [PubMed]
  153. Sandfeld-Paulsen, B.; Aggerholm-Pedersen, N.; Baek, R.; Jakobsen, K.R.; Meldgaard, P.; Folkersen, B.H.; Rasmussen, T.R.; Varming, K.; Jørgensen, M.M.; Sorensen, B.S. Exosomal Proteins as Prognostic Biomarkers in Non-Small Cell Lung Cancer. Mol. Oncol. 2016, 10, 1595–1602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Sandfeld-Paulsen, B.; Jakobsen, K.R.; Bæk, R.; Folkersen, B.H.; Rasmussen, T.R.; Meldgaard, P.; Varming, K.; Jørgensen, M.M.; Sorensen, B.S. Exosomal Proteins as Diagnostic Biomarkers in Lung Cancer. J. Thorac. Oncol. 2016, 11, 1701–1710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Gregorc, V.; Novello, S.; Lazzari, C.; Barni, S.; Aieta, M.; Mencoboni, M.; Grossi, F.; De Pas, T.; de Marinis, F.; Bearz, A.; et al. Predictive Value of a Proteomic Signature in Patients with Non-Small-Cell Lung Cancer Treated with Second-Line Erlotinib or Chemotherapy (PROSE): A Biomarker-Stratified, Randomised Phase 3 Trial. Lancet Oncol. 2014, 15, 713–721. [Google Scholar] [CrossRef]
  156. Salmon, S.; Chen, H.; Chen, S.; Herbst, R.; Tsao, A.; Tran, H.; Sandler, A.; Billheimer, D.; Shyr, Y.; Lee, J.-W.; et al. Classification by Mass Spectrometry Can Accurately and Reliably Predict Outcome in Patients with Non-Small Cell Lung Cancer Treated with Erlotinib-Containing Regimen. J. Thorac. Oncol. 2009, 4, 689–696. [Google Scholar] [CrossRef] [Green Version]
  157. Chung, C.H.; Seeley, E.H.; Roder, H.; Grigorieva, J.; Tsypin, M.; Roder, J.; Burtness, B.A.; Argiris, A.; Forastiere, A.A.; Gilbert, J.; et al. Detection of Tumor Epidermal Growth Factor Receptor Pathway Dependence by Serum Mass Spectrometry in Cancer Patients. Cancer Epidemiol. Biomark. Prev. 2010, 19, 358–365. [Google Scholar] [CrossRef] [Green Version]
  158. Taguchi, F.; Solomon, B.; Gregorc, V.; Roder, H.; Gray, R.; Kasahara, K.; Nishio, M.; Brahmer, J.; Spreafico, A.; Ludovini, V.; et al. Mass Spectrometry to Classify Non–Small-Cell Lung Cancer Patients for Clinical Outcome After Treatment With Epidermal Growth Factor Receptor Tyrosine Kinase Inhibitors: A Multicohort Cross-Institutional Study. JNCI J. Natl. Cancer Inst. 2007, 99, 838–846. [Google Scholar] [CrossRef]
  159. Lazzari, C.; Spreafico, A.; Bachi, A.; Roder, H.; Floriani, I.; Garavaglia, D.; Cattaneo, A.; Grigorieva, J.; Viganò, M.G.; Sorlini, C.; et al. Changes in Plasma Mass-Spectral Profile in Course of Treatment of Non-Small Cell Lung Cancer Patients with Epidermal Growth Factor Receptor Tyrosine Kinase Inhibitors. J. Thorac. Oncol. 2012, 7, 40–48. [Google Scholar] [CrossRef] [Green Version]
  160. Carbone, D.P.; Ding, K.; Roder, H.; Grigorieva, J.; Roder, J.; Tsao, M.-S.; Seymour, L.; Shepherd, F.A. Prognostic and Predictive Role of the VeriStrat Plasma Test in Patients with Advanced Non–Small-Cell Lung Cancer Treated with Erlotinib or Placebo in the NCIC Clinical Trials Group BR.21 Trial. J. Thorac. Oncol. 2012, 7, 1653–1660. [Google Scholar] [CrossRef] [Green Version]
  161. Fidler, M.J.; Fhied, C.L.; Roder, J.; Basu, S.; Sayidine, S.; Fughhi, I.; Pool, M.; Batus, M.; Bonomi, P.; Borgia, J.A. The Serum-Based VeriStrat® Test Is Associated with Proinflammatory Reactants and Clinical Outcome in Non-Small Cell Lung Cancer Patients. BMC Cancer 2018, 18, 310. [Google Scholar] [CrossRef]
  162. Grossi, F.; Genova, C.; Rijavec, E.; Barletta, G.; Biello, F.; Bello, M.G.D.; Meyer, K.; Roder, J.; Roder, H.; Grigorieva, J. Prognostic Role of the VeriStrat Test in First Line Patients with Non-Small Cell Lung Cancer Treated with Platinum-Based Chemotherapy. Lung Cancer 2018, 117, 64–69. [Google Scholar] [CrossRef] [Green Version]
  163. Mintz, Y.; Brodie, R. Introduction to Artificial Intelligence in Medicine. Minim. Invasive Ther. Allied Technol. 2019, 28, 73–81. [Google Scholar] [CrossRef]
  164. Libbrecht, M.W.; Noble, W.S. Machine Learning Applications in Genetics and Genomics. Nat. Rev. Genet. 2015, 16, 321–332. [Google Scholar] [CrossRef] [Green Version]
  165. Brown, J.M.; Campbell, J.P.; Beers, A.; Chang, K.; Ostmo, S.; Chan, R.V.P.; Dy, J.; Erdogmus, D.; Ioannidis, S.; Kalpathy-Cramer, J.; et al. Automated Diagnosis of Plus Disease in Retinopathy of Prematurity Using Deep Convolutional Neural Networks. JAMA Ophthalmol. 2018, 136, 803–810. [Google Scholar] [CrossRef]
  166. Baltruschat, I.M.; Nickisch, H.; Grass, M.; Knopp, T.; Saalbach, A. Comparison of Deep Learning Approaches for Multi-Label Chest X-Ray Classification. Sci. Rep. 2019, 9, 6381. [Google Scholar] [CrossRef] [Green Version]
  167. Bibault, J.-E.; Burgun, A.; Fournier, L.; Dekker, A.; Lambin, P. Artificial Intelligence in Oncology. Artif. Intell. Med. 2021, 361–381. [Google Scholar] [CrossRef] [Green Version]
  168. Hwang, T.J.; Kesselheim, A.S.; Vokinger, K.N. Lifecycle Regulation of Artificial Intelligence- and Machine Learning-Based Software Devices in Medicine. JAMA 2019, 322, 2285. [Google Scholar] [CrossRef]
  169. Jiang, Y.; Edwards, A.V.; Newstead, G.M. Artificial Intelligence Applied to Breast MRI for Improved Diagnosis. Radiology 2021, 298, 38–46. [Google Scholar] [CrossRef]
  170. Aerts, H.J.W.L.; Aerts, H.J.W.; Velazquez, E.R.; Leijenaar, R.T.H.; Parmar, C.; Grossmann, P.; Carvalho, S.; Bussink, J.; Monshouwer, R.; Haibe-Kains, B.; et al. Decoding Tumour Phenotype by Noninvasive Imaging Using a Quantitative Radiomics Approach. Nat. Commun. 2014, 5, 4006. [Google Scholar] [CrossRef] [Green Version]
  171. Sun, R.; Limkin, E.J.; Vakalopoulou, M.; Dercle, L.; Champiat, S.; Han, S.R.; Verlingue, L.; Brandao, D.; Lancia, A.; Ammari, S.; et al. A Radiomics Approach to Assess Tumour-Infiltrating CD8 Cells and Response to Anti-PD-1 or Anti-PD-L1 Immunotherapy: An Imaging Biomarker, Retrospective Multicohort Study. Lancet Oncol. 2018, 19, 1180–1191. [Google Scholar] [CrossRef]
  172. Chiu, H.-Y.; Chao, H.-S.; Chen, Y.-M. Application of Artificial Intelligence in Lung Cancer. Cancers 2022, 14, 1370. [Google Scholar] [CrossRef]
  173. Poore, G.D.; Kopylova, E.; Zhu, Q.; Carpenter, C.; Fraraccio, S.; Wandro, S.; Kosciolek, T.; Janssen, S.; Metcalf, J.; Song, S.J.; et al. Microbiome Analyses of Blood and Tissues Suggest Cancer Diagnostic Approach. Nature 2020, 579, 567–574. [Google Scholar] [CrossRef]
  174. Tirzïte, M.; Bukovskis, M.; Strazda, G.; Jurka, N.; Taivans, I. Detection of Lung Cancer with Electronic Nose and Logistic Regression Analysis. J. Breath Res. 2018, 13, 016006. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Coudray, N.; Ocampo, P.S.; Sakellaropoulos, T.; Narula, N.; Snuderl, M.; Fenyö, D.; Moreira, A.L.; Razavian, N.; Tsirigos, A. Classification and Mutation Prediction from Non–small Cell Lung Cancer Histopathology Images Using Deep Learning. Nat. Med. 2018, 24, 1559–1567. [Google Scholar] [CrossRef]
  176. Wang, C.; Ma, J.; Shao, J.; Zhang, S.; Liu, Z.; Yu, Y.; Li, W. Predicting EGFR and PD-L1 Status in NSCLC Patients Using Multitask AI System Based on CT Images. Front. Immunol. 2022, 13, 813072. [Google Scholar] [CrossRef]
  177. Choi, S.; Cho, S.I.; Ma, M.; Park, S.; Pereira, S.; Aum, B.J.; Shin, S.; Paeng, K.; Yoo, D.; Jung, W.; et al. Artificial Intelligence–powered Programmed Death Ligand 1 Analyser Reduces Interobserver Variation in Tumour Proportion Score for Non–small Cell Lung Cancer with Better Prediction of Immunotherapy Response. Eur. J. Cancer 2022, 170, 17–26. [Google Scholar] [CrossRef]
  178. Cheng, G.; Zhang, F.; Xing, Y.; Hu, X.; Zhang, H.; Chen, S.; Li, M.; Peng, C.; Ding, G.; Zhang, D.; et al. Artificial Intelligence-Assisted Score Analysis for Predicting the Expression of the Immunotherapy Biomarker PD-L1 in Lung Cancer. Front. Immunol. 2022, 13, 893198. [Google Scholar] [CrossRef] [PubMed]
  179. Wu, J.; Liu, C.; Liu, X.; Sun, W.; Li, L.; Gao, N.; Zhang, Y.; Yang, X.; Zhang, J.; Wang, H.; et al. Artificial Intelligence-Assisted System for Precision Diagnosis of PD-L1 Expression in Non-Small Cell Lung Cancer. Mod. Pathol. 2022, 35, 403–411. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic overview of the NSCLC biomarker discovery. (A) Different samples. Nasal epithelial scraping, exhaled breath, sputum, bronchoscopy, blood, and urine are used for NSCLC biomarker discovery. (B) NSCLC biomarker applications. The biomarkers can be helpful for screening, early detection, personalized medicine, and monitoring treatment response. (C) Different approaches for biomarker discovery. Biomarkers can be discovered through genomic, transcriptomic, proteomic, and metabolomics changes in NSCLC.
Figure 1. Schematic overview of the NSCLC biomarker discovery. (A) Different samples. Nasal epithelial scraping, exhaled breath, sputum, bronchoscopy, blood, and urine are used for NSCLC biomarker discovery. (B) NSCLC biomarker applications. The biomarkers can be helpful for screening, early detection, personalized medicine, and monitoring treatment response. (C) Different approaches for biomarker discovery. Biomarkers can be discovered through genomic, transcriptomic, proteomic, and metabolomics changes in NSCLC.
Cancers 14 05144 g001
Figure 2. Genetic alteration in NSCLC. (A) Frequency of driver genes in NSCLC. (B) Frequency of driver mutations in EGFR. Fusion partners of ROS1 (C) and ALK (D) in NSCLC.
Figure 2. Genetic alteration in NSCLC. (A) Frequency of driver genes in NSCLC. (B) Frequency of driver mutations in EGFR. Fusion partners of ROS1 (C) and ALK (D) in NSCLC.
Cancers 14 05144 g002
Table 1. Methylation-based biomarkers for early detection of NSCLC.
Table 1. Methylation-based biomarkers for early detection of NSCLC.
Biomarker (S)MethodSpecimenPopulationSensitivity and Specificity (%)Reference
MGMT, p16, RASSF1A, DAPK, and RAR-βMeta-analysisBlood37 case-control studiesNA[36]
APC, CDH13, KLK10,
DLEC1, RASSF1A,
EFEMP1, SFRP1, RARβ and p16INK4A
MSPTissues78 paired NSCLC specimens and adjacent normal tissues
110 stage I/II NSCLC and 50 plasmas cancer-free
83.64 and 74.0[37]
RARB2, RASSF1AQuantitative
methylation-specific PCR
Cell-Free DNA circulating in the blood
(cirDNA)
32 healthy donors and 60 patients with lung cancer87 and 75 [38]
SHOX2Quantitative real-time polymerase chain reactionPlasma371 samples from patients with lung cancer and controls60 and 90 [39]
DCLK1qMSP-PCRPlasma65 patients with lung cancer and 95 healthy donorsNA[39,40]
SEPT9Real-time PCR with the use of
specific SEPT9 promoter methylation probe
Plasma70 lung cancer patients and 100 healthy individuals44.3 and 92.3[41]
CDO1,
BCAT1,
TRIM58,
ZNF177
PyrosequencingParaffin-
embedded tissues
Bronchial aspirates and bronchoalveolar lavages
237 stage I NSCLC and 25 nontumoral matched lung tissuesNA[42]
TMEFF2Methylation-specific PCRSerum316 NSCLC, 50 NC9.2 and 100 [43]
Table 2. MicroRNA biomarkers for early detection of NSCLC.
Table 2. MicroRNA biomarkers for early detection of NSCLC.
BiomarkersSpecimenPopulationResultSensitivity and Specificity (%)Reference
13 miRNAPlasma939 participants, including 69
patients with lung
cancer and 870 healthy control subjects
Screening87 and 81[53]
miR-31 and miR-210Sputum35 patients with lung cancer and 40 healthy control subjectsScreening65.71 and 85.00[53,54]
miR-125a-5p, miR-25, and miR-126Serum24 early stage lung cancer
patients and 24 healthy control subjects
Early Detection87.5 and 87.5[49]
miR-21, miR-143, miR-
155, miR-210, miR-372
Sputum24 NSCLC cases and 6 negative controlsEarly Detection83.3 and
100
[55]
miR-141PlasmaNSCLC patients
(n: 72) and N.C. (n: 50)
Early Detection82.7 and 98[56]
miRNA (miR)-486 and miR-150Peripheral
Blood
Early Diagnosis and
Recurrence
90.9 and 81.8 for miR-486
and 81.8 for
miR-150
[57]
miRs-126, 145, 210, and 205-5p Plasma64 individuals comprising 34 lung cancer
patients and 30 healthy control smokers
Early Detection91.5 and 96.2[58]
I-miR-1254 and hsamiR-574-5pSerum22 individuals (11 healthy control subjects and 11
patients with early stage
NSCLC).
Early Detection82 and 77[58,59]
Table 3. Protein biomarkers for early detection of NSCLC.
Table 3. Protein biomarkers for early detection of NSCLC.
BiomarkerMethodSpecimenPopulationSensitivity and Specificity (%)Reference
FTL, FGB, RAB33B, RAB15 LC-MS/MS Urine Lung cancers from healthy control subjects90 and 90[69]
ERO1L, NARS,
PABPC4, RCC1, RPS25, TARS
(iTRAQ) labeling combined with 2D-LCMS/M.S. Tumor and Lung Tissues ADC tumors without
L.N. metastasis and adjacent normal tissues
NA[70]
44 proteins showed a fold-change > 3.75 (L.C.–MS/MS) Bronchoalveolar Lavage Fluid (BALF) Adenocarcinoma vs. healthy control subjects NA[71]
133 biomarkers LC-MS Bronchoalveolar Lavage (BAL) Lung cancer versus nonlung cancer NA[71,72]
GlcNAcylated AACT iTRAQ labeling and LC-MS/MS. Serum NSCLC patients, benign lung diseases subjects, and healthy individuals 90.8 and 76.9 [73]
α2 macroglobulin,
αmicroglobulin/bikunin, and SERPINA1
MRM Serum NSCLC lung adenocarcinoma cancer and healthy control subjects NA[74]
Elongation factor 1-
alpha 2, proteasome
subunit alpha type, and spermatogenesis-associated protein
LC-MS/MSSerumLung cancer and healthy control subjectsNA[75]
ALOX5, ALOX5AP,
SLC2A3, CEACAM6,
ITGAX, CRABP2, LAD1
LC-MS, PMR-MS, and immunohistochemistryTissues and Normal Bronchial Biopsies Adenocarcinoma
samples and benign nodules
NA [76]
LC-MS/MS: liquid chromatography–tandem mass spectrometry, PMR-MS: parallel reaction monitoring mass spectrometry, iTRAQ: isobaric tags for relative and absolute quantification MRM: multiple reaction monitoring, NA: not applicable.
Table 4. Ongoing clinical trials are investigating treatments targeting the AKT–PI3K–mTOR pathway in NSCLC.
Table 4. Ongoing clinical trials are investigating treatments targeting the AKT–PI3K–mTOR pathway in NSCLC.
NCT NumberClinical PhaseTypes of PatientsPurposePrimary End PointsIntervention/s
NCT04467801II60 metastatic/advanced NSCLCTreatmentProgression-free survivalIpatasertib
NCT04184921NA350 advanced lung cancer patientsNAProgression-free survivalOsimertinib
NCT03543683NA330 metastatic NSCLCNA1-year median progression-free survivalOsimertinib
NCT03532698NA100 stage IIIB and IV NSCLCNAObjective response rate (ORR)Osimertinib
NCT03845270II46 stage III or IV NSCLCTreatmentOverall responsePertuzumab + Trastuzumab + Docetaxel
NCT01306045II471 advanced NSCLC, SCLC, and thymic malignanciesTreatmentEstimate the response rate and feasibility of the use of tumor molecular profiling and targeted therapies in the treatment of NSCLC, SCLC, and thymic malignanciesAZD6244
MK-2206
Lapatinib
Erlotinib
Sunitinib
NCT02664935II423 NSCLC stage III or stage IVTreatmentObjective response (OR), progression-free survival time (PFS), and durable clinical benefit (DCB)AZD4547
Vistusertib
Palbociclib
Crizotinib
Selumetinib
Docetaxel
AZD5363
Osimertinib
Durvalumab
Sitravatinib
AZD6738
NCT02117167II999 metastatic relapse or stage IVTreatment Progression-free survivalAZD2014
AZD4547
AZD5363
AZD8931
Selumetinib
Vandetanib
Pemetrexed
Durvalumab
Savolitinib
Olaparib
NCT04591431II384 recurrent/metastatic breast, gastrointestinal cancer, non-small-cell lung cancer, or othersTreatment Overall response rate (ORR)Erlotinib
Trastuzumab
Trastuzumab emtansine
Pertuzumab
Lapatinib
Everolimus
Vemurafenib
Cobimetinib
Alectinib
Brigatinib
Palbociclib
Ponatinib
Vismogedib
Itacitinib
Ipatasertib
Entrectinib
Atezolizumab
Nivolumab
Ipilimumab
Pemigatinib
NCT04467801II60 metastatic/advanced NSCLCTreatmentProgression Free SurvivalIpatasertib
NCT04184921NA350 advanced lung cancerNAProgression-free survivalOsimertinib
NCT03543683NA330 metastatic NSCLCNA1-year median progression-free survival (PFS)Osimertinib
NCT03532698NA100 metastatic NSCLCNAObjective response rate (ORR)Osimertinib
NCT03845270II46 stage III and metastatic Treatment Overall responsePertuzumab + Trastuzumab + Docetaxel
NCT01306045II AZD6244
MK-2206
Lapatinib
Erlotinib
Sunitinib
NCT02664935II AZD4547
Vistusertib
Palbociclib
Crizotinib
Selumetinib
Docetaxel
AZD5363
Osimertinib
Durvalumab
Sitravatinib
AZD6738
NCT02117167II AZD2014
AZD4547
AZD5363
AZD8931
Selumetinib
Vandetanib
Pemetrexed
Durvalumab
Savolitinib
Olaparib
NCT04591431II Erlotinib
Trastuzumab
Trastuzumab emtansine
Pertuzumab
Lapatinib
Everolimus
Vemurafenib
Cobimetinib
Alectinib
Brigatinib
Palbociclib
Ponatinib
Vismogedib
Itacitinib
Ipatasertib
Entrectinib
Atezolizumab
Nivolumab
Ipilimumab
Pemigatinib
NCT01737502I and II47 lung cancer (squamous, Ras-mutated adenocarcinoma, or small-cell lung cancer)TreatmentMaximum tolerated dose of Auranofin, number and severity of all adverse events, and progression-free survivalAuranofin
Sirolimus
NCT05445791III Metformin Hydrochloride
NCT02664935II AZD4547
Vistusertib
Palbociclib
Crizotinib
Selumetinib
Docetaxel
AZD5363
Osimertinib
Durvalumab
Sitravatinib
AZD6738
NCT02117167II AZD2014
AZD4547
AZD5363
AZD8931
Selumetinib
Vandetanib
Pemetrexed
Durvalumab
Savolitinib
Olaparib
NCT04591431II Erlotinib
Trastuzumab
Pertuzumab
Lapatinib
Everolimus
Vemurafenib
Cobimetinib
Alectinib
Brigatinib
Palbociclib
Ponatinib
Vismogedib
Itacitinib
Ipatasertib
Entrectinib
Atezolizumab
Nivolumab
Ipilimumab
Pemigatinib
NCT05144698II22 advanced metastatic, recurrent, and unresectable solid tumors Treatment Safety of RAPA-201 Cell TherapyRAPA-201 Rapamycin-Resistant T Cells
Chemotherapy Prior to RAPA-201 Therapy
AZD2014; Novel mTOR inhibitor; AZD4547; FGFR inhibitor; AZD5363; Akt inhibitor; AZD8931; Novel EGFR/HER2/HER3 signaling inhibitor; NA; Not available.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Abbasian, M.H.; Ardekani, A.M.; Sobhani, N.; Roudi, R. The Role of Genomics and Proteomics in Lung Cancer Early Detection and Treatment. Cancers 2022, 14, 5144. https://doi.org/10.3390/cancers14205144

AMA Style

Abbasian MH, Ardekani AM, Sobhani N, Roudi R. The Role of Genomics and Proteomics in Lung Cancer Early Detection and Treatment. Cancers. 2022; 14(20):5144. https://doi.org/10.3390/cancers14205144

Chicago/Turabian Style

Abbasian, Mohammad Hadi, Ali M. Ardekani, Navid Sobhani, and Raheleh Roudi. 2022. "The Role of Genomics and Proteomics in Lung Cancer Early Detection and Treatment" Cancers 14, no. 20: 5144. https://doi.org/10.3390/cancers14205144

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