Next Article in Journal
Noninvasive Physical Plasma as Innovative and Tissue-Preserving Therapy for Women Positive for Cervical Intraepithelial Neoplasia
Next Article in Special Issue
Resistance to TKIs in EGFR-Mutated Non-Small Cell Lung Cancer: From Mechanisms to New Therapeutic Strategies
Previous Article in Journal
CD47-SIRPα Checkpoint Disruption in Metastases Requires Tumor-Targeting Antibody for Molecular and Engineered Macrophage Therapies
Previous Article in Special Issue
The Role of TP53 Mutations in EGFR-Mutated Non-Small-Cell Lung Cancer: Clinical Significance and Implications for Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Acquired Mechanisms of Resistance to Osimertinib—The Next Challenge

by
Alejandro Ríos-Hoyo
1,
Laura Moliner
1,† and
Edurne Arriola
1,2,*
1
Department of Medical Oncology, Hospital del Mar-CIBERONC (Centro de Investigación Biomédica en Red de Oncología), 08003 Barcelona, Spain
2
Cancer Research Program, IMIM (Institut Hospital del Mar d’Investigacions Mèdiques), 08003 Barcelona, Spain
*
Author to whom correspondence should be addressed.
Current Affiliation: The Christie NHS Foundation Trust, Manchester M20 4BX, UK.
Cancers 2022, 14(8), 1931; https://doi.org/10.3390/cancers14081931
Submission received: 31 March 2022 / Revised: 3 April 2022 / Accepted: 5 April 2022 / Published: 12 April 2022
(This article belongs to the Special Issue Mechanisms of Resistance in EGFR-Mutated Non-Small Cell Lung Cancer)

Abstract

:

Simple Summary

Osimertinib has revolutionized the treatment of EGFR-mutated tumors. Its current applications include the first-line setting, second-line setting, as well as the adjuvant setting. Although it represents a milestone in the context of targeted therapy, inevitably all tumors develop an acquired resistance, some mechanisms involve EGFR, others do so through alternative pathways leading to a bypass in osimertinib inhibition. It is key to understand these acquired mechanisms of resistance, both in the clinical setting, as well as in preclinical models, in order to develop and contribute to the identification of possible therapeutic strategies to overcome this acquired resistance.

Abstract

EGFR-mutated tumors represent a significant percentage of non-small cell lung cancer. Despite the increasing use of osimertinib, a treatment that has demonstrated an outstanding clinical benefit with a tolerable toxicity profile, EGFR tumors eventually acquire mechanisms of resistance. In the last years, multiple mechanisms of resistance have been identified; however, after progressing on osimertinib, treatment options remain bleak. In this review, we cover the most frequent alterations and potential therapeutic strategies to overcome them.

1. Introduction

Non-small cell lung cancer (NSCLC) accounts for approximately 80–90% of lung cancers; adenocarcinoma represents the most frequent histologic subtype of NSCLC. Different molecular oncogene addiction drivers have been identified in NSCLC, particularly in adenocarcinomas, including EGFR mutations, ALK and ROS1 rearrangements, MET mutations and amplifications, BRAF and KRAS mutations, RET fusions, NTRK fusions, and HER2 mutations, among others [1,2]. EGFR mutations were first described in 2004. The most frequent mutations are exon 21 point L858R mutation and in-frame deletions of exon 19, accounting for 40% and 50% of EGFR mutations, respectively, and conferring sensitivity to EGFR inhibitors. Other described mutations include G718X, S768I, and L861Q, as well as an in-frame insertion in exon 20 [3,4,5]. In NSCLC, EGFR mutations display a heterogeneous representation depending on ethnicity and region, accounting for 40–60% in South-East Asian patients, 9–67% in patients from the Americas, 22–27% in patients from the Indian subcontinent, around 12% in African Americans, and 10–20% in Caucasian patients [4,6,7]. First-generation EGFR inhibitors include erlotinib, gefitinib, and icotinib, which reversibly bind to EGFR and inhibit the binding of ATP to the tyrosine kinase domain. The second-generation EGFR inhibitors, afatinib and dacomitinib, bind covalently to EGFR, thus irreversibly inhibiting its activity [8]. After an initial benefit with the use of EGFR inhibitors, acquired resistance invariably develops; approximately 60% of the patients develop the T790M mutation, and other mechanisms of resistance include EGFR amplification concurrent with T790M, HER2 amplification, MET amplification, SCLC transformation, and others [9]. Third-generation EGFR inhibitors include osimertinib, rociletinib, and olmutinib; however, only the first is approved in current clinical practice. Osimertinib binds covalently to the C797 residue in the ATP-binding site of mutant EGFR [10].
Osimertinib was designed to overcome the effects of the T790M EGFR-resistant mutation while targeting the initial EGFR activating mutation [11]. It was first tested in the second-line setting, showing an increase in median progression-free survival (mPFS) compared to platinum plus pemetrexed based chemotherapy in those patients whose T790M-positive tumors had progressed to an EGFR-TKI therapy [12]. Subsequently, the results of the FLAURA clinical trial, demonstrating improvements in PFS and OS compared to first-generation EGFR TKIs, led to osimertinib being approved as frontline treatment [13,14].
Despite all these achievements, significant challenges remain ahead. Patients with EGFR mutated tumors seem to benefit less, if anything, from immunotherapy compared to other patients with advanced NSCLC and no targetable alterations. Clinical trials combining osimertinib and immunotherapy showed intolerable toxicity, and, on the other hand, most clinical trials combining chemo- and immunotherapy lacked a significant number of patients with EGFR mutant NSCLC to draw definitive conclusions on efficacy [15].
It is now more essential than ever that we understand the mechanisms of resistance of NSCLC EGFR mutant tumors. In this review, we present and analyze the most frequent alterations associated with resistance to osimertinib (Figure 1), potential therapeutic strategies in this scenario (Table 1), in vitro studies assessing mechanisms of resistance to osimertinib (Table 2), and potential therapeutic strategies.
To identify published articles or abstracts that described acquired mechanisms of resistance against first- or second-line osimertinib in clinical practice, as well as in vivo and in vitro models, we systematically searched PubMed (https://pubmed.ncbi.nlm.nih.gov/, accessed on 1 July 2021) for relevant studies. Our search criteria included the following terms: “osimertinib”, “resistance”, “non-small cell lung cancer”, “clinical trials”, “cell lines”, as well as keywords related to acquired mechanisms of resistance, including but not limited to “C797X”, “small-cell transformation”, “squamous cell transformation”, “MET amplification”, “HER2 amplification”, “HER2 mutation”, “RET fusions”, “BRAF mutations”, “BRAF amplification”, and “RAS mutations”.

2. EGFR-Dependent Mechanisms of Resistance

2.1. C797X Mutations

Osimertinib selectively blocks mutated-EGFR by irreversibly binding to its C797 residue. As expected, the most common EGFR-dependent mechanisms of resistance are based on mutations in this spot, usually a substitution to serine, leading to the C797S mutation [63].
In vitro studies evaluating the EGFR C797S mutations have mostly been achieved through the induction of genetic alterations in cell lines and not through exposure to osimertinib. For instance, NIH3T3 cells (an immortalized mouse embryonic fibroblast cell line) were used and transduced with lentiviruses bearing genes encoding the del19/T790M/C797S and L858R/T790M/C797S mutations in order to study the growth inhibitor effects of the compound CH7233163 [25]. Another mechanism to induce the C797S mutation was used in Ba/F3 cells (a murine, IL-3 dependent, hematopoietic cell line) harboring del19/L858R +/− T790M through a site-directed mutagenesis kit, developing constructs which were later transferred to the retroviral vector JP1540 or lentiviral vector JP1698 in cells infected with the viruses. after exposure to the EGFR TKI WZ4002 for two weeks, resistant clones were selected, identifying the previously transduced C797S [55,64].
The incidence of C797X in the clinical setting differs depending on the treatment setting of osimertinib. In the first-line scenario, 6 of 91 patients (7%) acquired a C797X mutation in plasma samples from the FLAURA clinical trial [65]. Analysis of ctDNA from plasma samples from 73 patients who progressed to osimertinib in the second line in the AURA3 trial showed the emergence of C797X mutations in up to 15% of the patients, 91% of them corresponding to C797S [66]. In the second-line setting, Oxnard et al. found C797S in 22% of 41 patients from tissue biopsies at progression [67]. Interestingly, the T790M mutations consistently remained detectable in all patients with a C797X mutation.
Several therapeutic strategies have been proposed to tackle the emergence of these mutations; however, currently, chemotherapy remains the standard treatment after progression to osimertinib.
Preclinical data suggest that, in the absence of T790M, cells remain sensitive to reversible binding EGFR TKIs (gefitinib, erlotinib), whose binding does not rely on C797 [68]. However, the only clinical information regarding this strategy is based on isolated clinical cases. One clinical case treated a patient after progression to first-line osimertinib with erlotinib, achieving a response that lasted 5 months [17].
In the second-line setting, the position in which T790M is located, whether in the same (cis) or in a different (trans) position related to C797S, has a significant impact. In vitro studies have reported that trans-T790M/C797S cells were resistant to third-generation EGFR TKIs but sensitive to the combination of first- and third-generation TKIs, whereas cis-T790M/C797S cells were refractory to all EGFR TKIs tested, as well as their combinations [68].
Case reports explored this clinical approach: a patient with a triple mutation tumor (del19/trans-T790M/C797S) received osimertinib and erlotinib after progression to osimertinib, reaching a PFS of 3 months [69]. In another reported case, a patient with a similar clinical scenario received osimertinib, erlotinib, and bevacizumab, which resulted in an 8-months PFS [18].
In the cis-T790M/C797S population, in vitro studies suggest that brigatinib (an ALK and ROS1 inhibitor with EGFR inhibition activity) could overcome this resistance mechanism with the combination of brigatinib plus an anti-EGFR antibody [70]. A retrospective cohort of patients who progressed to osimertinib with cis-T790M/C797S, treated with a combination of brigatinib and cetuximab, obtained an ORR (overall response rate) of 60%, a DCR (disease control rate) of 100%, and an mPFS (median progression-free survival) of 14 months [19].
Preliminary data was presented recently about patritumab deruxtecan (HER3-DXd), an antibody–drug conjugate (ADC) against HER3 [20]. Forty-four patients with EGFR mutant NSCLC who had received osimertinib and had a median of four prior lines of therapy obtained an ORR of 39% (with complete responses in 2% of the patients) and an mPFS of 8.2 months. The efficacy of patritumab deruxtecan was maintained across various resistance mechanisms and different levels of HER3 expression. Interestingly, in the subgroup of patients with a known EGFR resistance mutation, ORR rose to 50%. This promising drug will be further evaluated in the HERTHENA-Lung01 trial [71] and also in combination with osimertinib [72].
Preventive strategies are also being tested in patients with EGFR-mutated untreated NSCLC. One phase I/II study is exploring whether the combination of osimertinib and gefitinib could delay the emergence of acquired resistance mechanisms. To date, the combination has shown an acceptable toxicity profile [73].

2.2. Less Common EGFR-Dependent Alterations

Less frequent mutations different to C797X have also been described. After progression to osimertinib in the FLAURA study [65], 2% of patients acquired the L718Q and 1% acquired the S768L mutations, respectively. In the AURA3 study [66], L792X emerged in 3% of the patients who progressed; G796X, L718Q, and exon 20 insertions were described at a frequency of 1% each. In a cohort of 93 Asian patients who received osimertinib as second-line treatment, G796X, G719A, L718Q, and L792X emerged in 2%, 2%, 8%, and 12% of cases, respectively [74]. Most of these alterations occur concurrently with other mutations, which suggests the appearance of clonal heterogeneity at progression. L718Q is a notable exception, as it usually emerges isolated.
There is limited evidence regarding the functional effect of these less frequent mutations. G796X mutations occur close to C797 residue and thus, affect the interaction between osimertinib and its binding residue [75]. L792X, L718Q, and G719X mutations also decrease the interaction between osimertinib and its binding residue [76,77].
A case report by Fang et al. described a patient whose EGFR L858R/L781Q tumor responded partially to a third-line of treatment with afatinib before progressing and identifying a KRAS mutation [78]. For L792X mutations, in vitro studies using different EGFR inhibitors and combinations have failed to show any sign of response [77]. To our knowledge, other strategies to address these rare mutations are not published.
Another EGFR-dependent mechanism of resistance is the amplification of EGFR, with an incidence of 33% and 31% in the first and later lines, respectively. In vitro studies have confirmed EGFR amplification as an independent resistance mechanism to osimertinib, even though, in most cases (up to 50%), it co-occurs with other EGFR resistance mutations [62,74,79]. One branch of the ORCHARD trial is currently evaluating the combination of osimertinib and necitumumab in this population [80].

2.3. Fourth-Generation EGFR TKIs as a Strategy to Overcome Resistance to Osimertinib

Given that the C797S mutation is the most frequently acquired resistance mechanism to osimertinib, fourth-generation EGFR TKIs are currently being developed. We describe some of the most relevant in the following section.

2.3.1. EAI045

EAI045 is the first allosteric reversible, non-ATP competitive inhibitor targeting L858R/T790M/C979S EGFR co-occurring mutations [21]. In vitro and in vivo studies confirmed that EAI045 and cetuximab were effective in inhibiting L858R/T790M/C979S cells, as well as in mice models with L858R/T790M, leading to an inhibition of the downstream signaling proteins. It is noteworthy that EAI045 was not effective in inhibiting exon19del/T790M mutation models [22]. EAI045 has also shown activity inhibiting the exon 21 L861Q mutation, as reported in preclinical studies [23]. No clinical trials were found studying this molecule.

2.3.2. JBJ-04-125-02

JBJ-04-125-02 is a reversible, non-ATP competitive allosteric inhibitor that was effective in preclinical models harboring the L858R/T790M/C979S mutations in combination with osimertinib, leading to increased apoptosis and an inhibition of cellular growth. JBJ-04-125-02 lacks binding affinity against exon19del; therefore, it is ineffective in this situation [24]. At the moment, we could not find any clinical trials evaluating this drug.

2.3.3. CH7233163

CH7233163 is a non-covalent ATP-competitive inhibitor, which has shown in vitro and in vivo activity overcoming the EGFR del19/T790M/C797S, L858R/T790M/C797S, del19/T790M, L858R/T790M, del19, and L858R mutations [25]. However, no clinical trials testing this molecule are ongoing at the moment. Nonetheless, CH7233163 appears to be an attractive drug to be tested in this setting.

2.3.4. BLU-945

BLU-945 is a potent, selective EGFR inhibitor with activity against del19/T790M/C797S, L858R/T790M/C797S, del19/T790M, and L858R/T790M in in vitro and in vivo assays [26,27]. The phase I clinical trial SYMPHONY is currently evaluating the use of BLU-945 in the second-line for patients whose tumors acquired C797S [81].

2.3.5. Other Drugs

Early evidence is available for other inhibitors. TQB3804 is a drug that effectively inhibited del19/T790M/C797S, L858R/T790M/C797S, del19/T790M, and L858R/T790M in in vitro and in vivo assays [82]. TRE-069 has shown preclinical data as an EGFR del19/T790M/C797S inhibitor [83].

3. Histologic and Phenotypic Transformation

3.1. Small Cell Transformation

The underlying mechanisms that cause a number of EGFR mutant tumors to transform into small cell lung cancer (SCLC) are unknown; limited data due to the lack of tissue availability makes it difficult to understand this phenomenon [84]. Its incidence has been reported between 6% and 15% in the second-line setting, being less common in the first-line setting, with up to 4% of cases reported [79,85,86].
Tissue biopsy at the time of progression to osimertinib remains of paramount importance in order to identify this transformation; however, in several cases, elevated serum levels of neuron-specific enolase can be found and suggest this phenomenon [87].
Consistently, most of the transformed tumors maintain the original activating EGFR mutation [88]. However, after transformation, EGFR expression levels drop, which explains why these tumors are not sensitive to EGFR-TKI therapy [89]. Common mutations identified in samples following transformation are TP53 (91%), Rb1 (58%), and PIK3CA (27%). Interestingly, in this study, these mutations were not described prior to histologic transformation [28].
Among the population of EGFR-mutated advanced NSCLC, a higher risk population for this transformation into SCLC has been defined by the presence of the concurrent loss of both TP53 and RB1. This population represents approximately 5% of all EGFR mutated tumors, and its incidence of transformation is 18%, in contrast to EGFR mutant tumors without mutation on RB1 and TP53, representing 3% of the cases. These patients have worse outcomes, with median OS in a retrospective cohort of 29.1 months compared to 56.4 months of patients with preserved RB1 and TP53 [90]. A retrospective study identified that pretreatment Rb1 loss was significantly associated with SCLC transformation. The study also identified patients with worse outcomes as those with concurrent pretreatment TP53 and Rb1 loss [91].
A retrospective analysis examined a cohort of 67 patients diagnosed with transformed-SCLC, where up to 30% had received osimertinib. The most common treatment was platinum-etoposide, achieving an ORR of 54%, mPFS of 3.4 months, and mOS of 10.9 months since the transformation [28]. Interestingly, there were no responses among the 17 patients who were treated with nivolumab in monotherapy or combination with ipilimumab. To date, platinum-etoposide chemotherapy is the only treatment strategy with confirmed clinical efficacy. An ongoing clinical trial is testing the combination of osimertinib and platinum-etoposide chemotherapy [29]; the role of chemo-immunotherapy combinations remains unknown in this population.

3.2. Squamous Cell Transformation

Schoenfeld et al. identified five cases of transformation into squamous cell carcinoma [79]. The incidence was similar in first-line osimertinib, with 7%, and later-lines, with 9%. All squamous cell-transformed tumors maintained the original EGFR mutation. There was no clear molecular pattern after evolving into squamous tumors, with only one patient gaining a PIK3CA mutation.
Currently, there are no clinical data regarding therapeutic strategies in these patients; however, a histology-based approach is recommended [30].

4. MET Amplification

The MET (mesenchymal-epithelial transition factor) gene is located on chromosome 7q21-q31. It encodes for the MET polypeptide, which is processed into a glycoprotein and serves as a transmembrane receptor tyrosine kinase. MET becomes activated through the binding of the hepatocyte growth factor ligands, subsequently activating different pathways, such as RAS/ERK/MAP kinase, PI3K/AKT, Wnt/β-catenin, and STAT [92,93].
MET gene amplification has been identified as a resistance mechanism against osimertinib in HCC827 cells, which is a hypersensitive EGFR exon 19 mutant NSCLC cell line [94]. Induction of resistance to osimertinib was achieved by exposing cells to gradually increasing concentrations of osimertinib (initially at 10 nM up to 500 nM) for 6 months. MET amplification, as well as hyperactivation, were detected [56] in the osimertinib resistant cells.
In clinical practice, MET amplifications are found between 9% and 24% of tumors that progress to osimertinib as second or later lines [66,67,79,85]. After progression to first-line osimertinib, MET amplification was found between 7% and 15%, representing the second most common resistance mechanism after C797X [65,79].
Multiple clinical trials are currently open to find effective strategies in this setting of unmet need, especially since osimertinib is now the current standard of treatment in the first line. Different combinations of both anti-EGFR and anti-MET drugs show promising results for these patients.
The TATTON phase 1b study assessed the combination of osimertinib and savolitinib in patients with EGFR-mutated and MET-amplified tumors across different clinical scenarios [31]. The cohort B1 consisted of 69 patients who had previously received a third-generation TKI, achieving an ORR of 30%, a DCR of 75%, and an mPFS of 5.4 months. Even though responses were considerably lower than in other cohorts, it represents a significant achievement considering that up to one-third of the patients received over three lines of treatment. Based on these results, as well as on its safety profile, an additional phase 2 trial was initiated [95]. One of the cohorts of the multi-arm clinical trial ORCHARD also examines this combination [19].
The phase I clinical trial CHRYSALIS evaluated the combination of lazertinib (a third-generation EGFR TKI) and amivantamab (a bi-specific antibody that targets both MET and EGFR) in patients with osimertinib-relapsed and chemotherapy-naïve advanced NSCLC [32,96]. In this study, patients with an identified MET or EGFR/MET-based mechanism of resistance had a 50% ORR. Interestingly, the response rate rose to 90% in 10 patients with high expression by immunohistochemistry (defined as a combined EGFR and MET H-score > 400). However, of these high expressors, five of them had an unknown mechanism of resistance. The results of the general cohort showed an ORR of 36%, with an mDoR (median Duration of Response) of 9.6 months and an mPFS of 4.9 months. A cohort of the future CHRYSALIS-2 trial will try to validate these findings.
The INSIGHT 2 trial is evaluating tepotinib (a MET TKI) in monotherapy and its combination with osimertinib in patients who progressed to first-line osimertinib with an acquired MET amplification [33].
Several case reports have used the combination of osimertinib and crizotinib in patients whose tumors had progressed to osimertinib as a second or later line and had acquired a MET amplification. Responses in the published cases differ, with benefits ranging from 2 to 7 months [34,35,97].
Responses were also seen in the clinical trial that evaluated patritumab-deruxtecan; however, only 8% of the tumors had MET alterations in the pretreatment samples. Consequently, efficacy data in this setting is currently insufficient to draw any conclusion [98].

5. HER2 Alterations

HER2 is a tyrosine kinase receptor that belongs to the EGFR family. Among the different HER family proteins, HER2 has the strongest catalytic kinase activity [57,99]. Its phosphorylation leads to the downstream activation of the PI3-Akt, MAPK, and ERK MET/MAPK pathways [1,2,3]. Aberrations in HER2 can be found in NSCLC, including amplifications and mutations, both leading to HER2 activation [6].
In vitro studies have developed the HER2 exon 16 skipping (HER2D16) in HEK293 cells (human embryonic kidney cell line) [100] and H1975 cells (T790M/L858R) through plasmid transfection. In HEK293, HER2D16 was able to induce the phosphorylation of ERK, confirming the signaling activity of this mutation. H1975 cells with HER2D16 were confirmed to be resistant to osimertinib, pointing out that the HER2 interaction with EGFR was able to form heterodimers, thus maintaining the phosphorylation of EGFR, as well as AKT and ERK [57].
In the clinical setting, HER2 aberrations have been identified as acquired mechanisms of resistance to osimertinib. In the first-line therapy, amplification was detected in 2% of the cases and HER2 mutations in 1% [65]. In second-line osimertinib, findings include HER2 amplifications (5%), HER2 amplifications co-occurring with EGFR L792X + C797X + PIK3CA amplification (1%), HER2 amplifications and EGFR G796S + MET amplification (1%), and HER2 amplifications and PIK3CA amplifications (1%) [66]. The most common mutations reported include in-frame exon 20 insertions, as well as exon 16 skipping HER2 deletion [57,101,102]. Preclinical models have shown the efficacy of osimertinib in HER2 amplification [16,101], however, these results have not been translated to a clinical benefit.
In the clinical trial evaluating patritumab-deruxtecan, some patients presented with HER2 alterations (mutations or amplifications) and benefited from the drug in terms of response [20].
Different targeted therapies against HER2 have been tested in NSCLC without co-occurring EGFR mutations, including lapatinib, neratinib (HER2 and EGFR inhibitors) [36,37,103], and trastuzumab-deruxtecan (T-DXd, an ADC against HER2, linked to a topoisomerase I inhibitor). These combinations would represent an attractive strategy to overcome this resistance mechanism.

6. RET Alterations

The RET proto-oncogene encodes a receptor tyrosine kinase (RTK). Its activation can develop as a consequence of gain of function amino acid substitutions and genomic rearrangements, leading to the formation of fusion proteins; RET fusions are usually generated by pericentric and paracentric inversions of chromosome 10 [104,105]. RET alterations frequently coexist with other genomic alterations such as TP53, cell cycle-associated genes, the PI3K pathway, and mitogen-activated protein (MAP) kinase effectors [104]. In the AURA3 study, RET-ERC1 fusions were reported in 1% of the cases as a mechanism of resistance to osimertinib [66]. Other RET fusions include CCD6-RET, NCOA4-RET, and MYH9-RET [85,106,107]. Rearrangements have also been described in the RET region of exon 11 to intron 11, which are sites of fragile DNA secondary structures [108]. Liquid biopsy is useful in detecting tumor heterogeneity, particularly when diverse resistance mechanisms develop in different tumor sites, as reported in the case of an NSCLC with an EGFR exon 19del and T790M, which had a liver progression with an acquired CCDC6-RET fusion. The patient received treatment with selpercatinib and responded to this combination [39].

7. BRAF Alterations

BRAF is a member of the RAF family of serine/threonine kinases. It is a part of the MAP kinase pathway, and its signaling follows downstream from BRAF to MEK 1 and 2 and ERK, which further phosphorylates multiple molecules [109].
In vitro studies have identified mutations in BRAF G469A as a resistance mechanism against osimertinib in PC9 cells (EGFR exon 19 delE746-A750). The cells were exposed to increasing concentrations of osimertinib (ranging from 10 to 500 nM) for nine months. When identifying possible resistance mechanisms, EGFR, HER2, or MET alterations were not identified, however, a BRAF G469 mutation was detected. The study observed that this mutation maintained the activity of the MAP kinases pathway. Combination treatment with osimertinib plus selumetinib or trametinib (both MEK 1/2 inhibitors) was effective in restoring the sensitivity of osimertinib to resistant BRAF G469A mutated cells [42].
In the clinical trial AURA3, BRAF alterations were reported as a mechanism of resistance to osimertinib in 3% of the cases (BRAF V600E mutation). This mutation co-occurred with MET amplification and FGFR3-TACC3 fusion, with concurrent EGFR C797X mutation; other co-occurring alterations included MET amplification with BRAF V600E and CDK6 amplification [66,109,110]. Analysis of the FLAURA study detected BRAF V600E in 3% of the cases as a mechanism of resistance to first-line osimertinib [65]. A combination of osimertinib with BRAF and MEK inhibitors, dabrafenib, and trametinib has been used effectively in the case of BRAF V600E acquired mutation and EGFR exon 19del/T790M, which had progressed to osimertinib, with acceptable tolerance to the treatment [111]. The use of osimertinib in combination with a single BRAF inhibitor, vemurafenib, has been reported as a successful strategy in overcoming BRAF V600E acquired resistance to osimertinib [41].
BRAF fusions represent approximately 2% of the cases of acquired resistance to osimertinib; described fusions include PJA2-BRAF, MKRN1-BRAF, and AGK-BRAF. The latter was shown to develop in the primary tumor but was absent in a metastatic lesion, highlighting the impact of clonal heterogeneity in resistance mechanisms [112,113]. The use of osimertinib in combination with the MEK inhibitor trametinib has been reported as a fifth line therapy in a patient with NSCLC and EGFR exon 19del and an AGK-BRAF fusion, observing a partial response to this treatment [114].

8. KRAS Mutations

KRAS is a member of the membrane-bound family proteins RAS. It possesses inherent GTPase activity. RAS can activate different effector molecules, such as RAF and the MAP kinase pathway, as well as PI3K, ultimately activating mTOR [115].
In vitro studies have detected RAS alterations as a resistance mechanism against first-line osimertinib in PC9 cells, developing an NRAS E63K mutation. As well as NRAS G12V and G12R, KRAS copy number gains were also detected. PC9 cells were chronically treated with escalating concentrations of osimertinib and a single concentration of osimertinib; different models were used, exposing cells to a final concentration of osimertinib of 160 nM and 1500 nM in the different models. The treatment combination of osimertinib with either selumetinib or an Aurora kinase b inhibitor was effective in overcoming resistance to osimertinib [43].
In the clinical setting, KRAS mutations were reported as acquired resistance mechanisms in 1% (KRAS G12D) and 3% (KRAS A1467T, KRAS G12C, and KRAS G12D, 1% each mutation) in the AURA3 and FLAURA trials, respectively [65,66]. Other reported acquired resistance mechanisms include co-existing alterations, such as the loss of T790M and development of C797S, with different KRAS mutations, including G12D, G12S, G61K, and Q61R, as well as an amplification of CDK4/KRAS/MDM2 [67,116,117,118,119]. To the best of our knowledge, no strategies have been developed to overcome this acquired resistance mechanism; however, the concomitant use of osimertinib with novel KRAS G12C inhibitors, such as sotorasib [44] or adagrasib [45], could be an attractive alternative in this particular mutation.

9. PI3K Alterations

PI3K is activated through different upstream pathways involving tyrosine kinases, G coupled proteins, and RAS-related GTPases. Its activation can lead to diverging downstream pathways, such as Akt, TEC family tyrosine kinases, and mTOR [120]. PIK3CA mutations or amplifications were reported as an acquired resistance mechanism in the AURA3 trial, presenting, in coexistence with HER2 amplification, CCND2 and CCNE1 amplifications. PIK3CA E545K was the most frequent mutation detected [66]. In the FLAURA trial, PIK3CA mutations were detected in 6% of the cases (E545K 4%, E453K 1%, H1047R 1%) [65]. An in vitro study observed that the PIK3CA H1047R mutation drives resistance to osimertinib. Co-treatment of cells with osimertinib and alpelisib, a PIK3CA inhibitor, resulted in a downregulation of the AKT signaling pathway [48]. A study involving 605 patients with NSCLC detected up to 14.9% PIK3CA, PTEN, or Akt mutations in patients who had progressed to EGFR inhibitors. Subsequently, six patients were treated with EGFR TKIs and everolimus, an mTOR inhibitor. This combination resulted in a limited antitumoral activity with stable disease in five patients and a progressive disease in one [46,47].

10. Cell Cycle Gene Alterations

Cyclin D-dependent kinases (CDK4 and CDK6) are major oncogenic drivers; their sustained activation leads to cancer cells entering the cell cycle repeatedly by producing G1-S phase transitions and reducing the duration of the G1 phase. Genes associated with CDKs include CCND, CCNE, and CDKN, among others [121].
Cell cycle alterations have been described in in vitro studies as a resistance mechanism to osimertinib in H1975 cells (EGFR L858R and T790M mutations) through an increased expression of CDK4. Osimertinib resistance was achieved by exposing the cells to osimertinib concentrations from 5 nM to 1.5 µM for 22 weeks. Subsequently, H1975 OR (osimertinib resistant) cells were developed. No alterations in MET, KRAS, BRAF, MEK, or PI3K were detected, and neither C797X nor T790M loss was identified. H1975 OR cells had fewer G1 phase and more G2 phase cells than H1975 cells [49]. A combination of palbociclib, a CKD4/6 inhibitor, with osimertinib in an in vitro study was effective in controlling tumor cells proliferation [49]. Similar results were observed in an in vitro model, where a combination of abemaciclib, a CDK 4/6 inhibitor, and osimertinib inhibited the onset of resistance to osimertinib [50].
In the AURA3 trial, cell cycle gene alterations were acquired in 12% of samples, as follows, mutation of CDKN2A E27fs 1%, and amplifications of CCND1 1%, CCND2 1%, CCNE1 7%, and CDK6 7% [66]. In the FLAURA study, the alterations reported were amplifications of the following genes: CCND1 2%, CCND2 1%, CCND3 1%, CCNE1 2%, CDK4 2%, and CDK6 3% [65]. Other studies have reported cell cycle alterations in up to 26.3% of cases after progression to osimertinib [122]. These genomic alterations have been associated with poor outcomes regarding progression-free survival and overall survival in the context of osimertinib treatment [123]. A clinical trial with osimertinib in combination with a CDK 4/6 inhibitor in a population with EGFR mutated NSCLC with or without T790M is evaluating the efficacy of this combination; however, this trial excludes patients who previously received osimertinib [124].

11. AXL Overexpression

AXL is a receptor tyrosine kinase, which belongs to the tumor-associated macrophage family (TAM), including TYRO-3 and MER. AXL ligand is Gas6, which binds to the ectodomain of AXL; its activation leads to cellular growth, proliferation, motility, and invasion, involving different signaling pathways, such as MAP kinase and PI3K/Akt, among others. Furthermore, AXL has been implicated in the process of epithelial-mesenchymal transition [125,126].
In vitro studies have identified a concurrent MET amplification, as well as an AXL upregulation, as a mechanism of acquired resistance in EFGR TKI resistant NSCLC cell lines [127,128]. Furthermore, AXL induced reactivation of HER3, MET, and EGFR was associated with maintaining cell survival and resistance to osimertinib [129]. A study in cell lines evaluated the use of cabozantinib and osimertinib in osimertinib-resistant NSCLC with AXL upregulation, observing a significant tumor suppression [128].
AXL activation has been described as an acquired resistance mechanism to first-line osimertinib in HCC827 cells (EGFR exon 19 delE746-A750), PC9 cells, H1975 cells, and HCC4006 cells (EGFR exon 19 delL747-A750, P ins) [58,59,60]. HCC827 cells were treated with 30 nM of osimertinib for 3 days and then cultured in a drug-free medium for more than 6 months, subsequently generating osimertinib-resistant cells with MET amplification and AXL upregulation identified by Sanger sequencing [127]. In another study, HCC827, HCC4006, PC-9, HCC4011, and H1975 cells were used to develop acquired osimertinib-resistant cell lines through exposure to osimertinib with an escalation method ranging from 10 nmol/L to 2 µmol/L over 6 months, or through a high concentration method, exposing cells to osimertinib at a 2 µmol/L over 6 months. MET amplification and AXL expression were detected; however, T790M, C797S mutations of KRAS, NRAS, BRAF, and TP53 mutations were not detected [128].
AXL overexpression has been reported as an acquired mechanism of resistance to first- and second-generation EGFR TKIs, as well as to second-line osimertinib [126,130]. In NSCLC EGFR mutated patients, baseline AXL overexpression was associated with a decreased response to first-line osimertinib, compared to non-overexpressing tumors [131].
Enapotamab vedotin and ADC specific against AXL has shown activity in an in vivo model of osimertinib resistant NSCLC [51] and is currently being tested in a Phase 1/2 clinical trial including different tumors [52].

12. Insulin-like Growth Factor (IGF)-1 Receptor Activation

The IGF-1 receptor belongs to the insulin receptor family and has roles in cell growth and differentiation. It can be activated by IGF1, IGF2, and insulin [132].
Activation of the IGF-1 receptor as an acquired resistance mechanism to osimertinib was described in in vitro studies using PC9 and H1975 cells. PC9 cells were exposed to gefitinib over 6 months, developing resistance through the T790M mutation. Subsequently, these cells were cultured with stepwise escalation to osimertinib in concentrations ranging from 150 nmol/L to 1 µmol/L over 6 months, thus developing osimertinib resistant cells. H1975 cells were exposed to osimertinib using the high-concentration method, culturing cells with 1µmol/L osimertinib for 3 months. Whole exome sequencing of several genes, including EGFR, MET, KRAS, MEK, BRAF, and PIK3CA, was performed. No alterations were detected; however, the IGF-1 receptor was detected to be overactivated in these cells [53].
Activation of the IGF-1 receptor through IGF-2 overexpression has been identified both in clinical specimens and in cell lines of NSCLC EGFR mutated osimertinib resistant and proposed as a mechanism of acquired resistance. Furthermore, treatment with linsitinib (an IGF-1 receptor inhibitor) and osimertinib restored sensitivity to osimertinib in in vitro studies [53,133]. We could not find any clinical trials evaluating the combination of linsitinib with osimertinib or other TKIs to overcome this resistance mechanism.

13. Epithelial-Mesenchymal Transition (EMT)

EMT is a cellular process involved in different types of cancer, allowing cells to enhance invasive capacity, cancer stem-cell similar properties, as well as resistance to treatments. In this process, epithelial cells lose their cell polarity and cell-to-cell adhesion, including a downregulation of epithelial proteins, such as E-cadherin, and acquire mesenchymal characteristics, including increased migration and invasion properties, and an upregulation of proteins, such as N-cadherin and vimentin. EMT develops through the involvement of different proteins and pathways, including TGF-β, SMAD and MAP kinase pathways, induction of IGF-1 receptor, and Notch signaling, among others [134,135,136].
In vitro studies using H1975/AR cells (gefitinib resistant) were exposed to osimertinib through a stepwise escalation process, later detecting a decreased E-cadherin and increased vimentin expression, as well as the capability to form larger spheroids [61]. In a different study with H1975 OR cells, mRNA expression was evaluated, detecting an upregulation of N-cadherin and vimentin and a downregulation of E-cadherin, which was further confirmed by a Western-Blot analysis [137], supporting EMT as an additional mechanism of resistance to osimertinib.

14. Other Rare Acquired Resistance Mechanisms

Other mechanisms of resistance to osimertinib have been described in in vitro studies, including the Src-AKT pathway and EGFR wild-type amplification in PC9 and H1975 cells. The cells were cultured with osimertinib at an initial concentration of 0.3 µmol/L and incrementally increased to 1 µmol/L. After several months of exposure, osimertinib resistant cells were developed [62].
Other reported resistance mechanisms in the clinical setting to first and second-line osimertinib include fusions of SPTBN1-ALK, EML4-ALK, FGFR3-TACC3, NTRK1-TMP3, and GOPC-ROS1, among others [65,66,138,139].

15. Conclusions

Osimertinib has become the standard of care in the first-line setting for patients with advanced lung cancer and sensitizing EGFR mutations. Despite the long duration of response with osimertinib, eventually, all patients will progress. Up until now, chemotherapy remains the standard treatment in this setting. However, in the advent of new diagnostic genomic platforms, a myriad of diverse mechanisms of resistance to osimertinib have been identified. Different molecular mechanisms can coexist, as well as histologic transformation and molecular derangements. It is therefore of paramount importance to assess these genomic and phenotypic changes through a rebiopsy when feasible and alternatively using liquid biopsies. One of the main challenges when multiple alterations are found is to elucidate which mechanism is driving the resistance to efficiently target that mechanism and improve the chances of therapeutic success. In vitro studies can contribute to this. Another important challenge is that the mechanisms of resistance are diverse, thus, making it complex to design clinical trials encompassing all possible scenarios. Umbrella trials such as ORCHARD are good examples of maximizing the efficiency to address different mechanisms of resistance in one trial.
Another strategy is designing drugs that can potentially overcome different mechanisms of resistance, such as patritumab-deruxtecan or even resistance mechanisms to different drivers, e.g., the phase 1 TROPION-PanTumor01 study. In this trial, datopotamab deruxtecan, an anti-TROP2 ADC, was used to treat patients with NSCLC and targetable genomic alterations (including EGFR mutation and ALK and RET fusions) who had progressed to prior treatment options (82% of patients had received ≥3 lines). The ORR was 35%, and the median DoR was 9.5 months, thus, positioning datopotamab deruxtecan as an attractive option in the future [105]. The outcomes of larger phase 2 or 3 studies will direct the future in the treatment of patients harvesting these complex genomic alterations.
Upfront combinations of osimertinib with chemotherapy or other targeted agents might also result in the delay of the emergence of resistance, but still, we need to pursue curative strategies in this scenario.
To date, there are no available data on mechanisms of acquired resistance to osimertinib in the adjuvant setting; however, the increasing use of osimertinib in this early disease setting will probably lead to changes in the profile of resistance.
The technological advances, along with the creative designs of clinical trials in these infrequent subpopulations, will be key to improving the outcomes of patients with EGFR mutant NSCLC.

Author Contributions

Conceptualization: A.R.-H., L.M. and E.A.; methodology: A.R.-H., L.M. and E.A.; data curation A.R.-H., L.M. and E.A.; writing—original draft preparation: A.R.-H., L.M. and E.A.; writing—review and editing: A.R.-H., L.M. and E.A.; visualization: A.R.-H., L.M. and E.A.; supervision: E.A.; project administration: E.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available in the article.

Conflicts of Interest

E.A. reports personal fees and non-financial support from BMS, personal fees from Astra Zeneca, grants, personal fees, and non-financial support from Roche, personal fees and non-financial support from MSD, personal fees from Lilly, grants and personal fees from Pfizer, and personal fees from Boehringer Ingelheim, outside the submitted work. A.R.-H. and L.M. declare no conflict of interest.

References

  1. Planchard, D.; Popat, S.; Kerr, K.; Novello, S.; Smit, E.; Faivre-Finn, C.; Mok, T.; Reck, M.; Van Schil, P.; Hellmann, M.; et al. Correction to: “Metastatic non-small cell lung cancer: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up”. Ann. Oncol. 2019, 30, 863–870. [Google Scholar] [CrossRef] [PubMed]
  2. Lamberti, G.; Andrini, E.; Sisi, M.; Rizzo, A.; Parisi, C.; Di Federico, A.; Gelsomino, F.; Ardizzoni, A. Beyond EGFR, ALK and ROS1: Current evidence and future perspectives on newly targetable oncogenic drivers in lung adenocarcinoma. Crit. Rev. Oncol. 2020, 156, 103119. [Google Scholar] [CrossRef] [PubMed]
  3. Paez, J.G.; Jänne, P.A.; Lee, J.C.; Tracy, S.; Greulich, H.; Gabriel, S.; Herman, P.; Kaye, F.J.; Lindeman, N.; Boggon, T.J.; et al. EGFR Mutations in Lung Cancer: Correlation with Clinical Response to Gefitinib Therapy. Science 2004, 304, 1497–1500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Hsu, W.-H.; Yang, J.C.-H.; Mok, T.; Loong, H. Overview of current systemic management of EGFR-mutant NSCLC. Ann. Oncol. 2018, 29, i3–i9. [Google Scholar] [CrossRef]
  5. Russo, A.; Franchina, T.; Ricciardi, G.R.R.; Picone, A.; Ferraro, G.; Zanghì, M.; Toscano, G.; Giordano, A.; Adamo, V. A decade of EGFR inhibition in EGFR-mutated non small cell lung cancer (NSCLC): Old successes and future perspectives. Oncotarget 2015, 6, 26814–26825. [Google Scholar] [CrossRef] [Green Version]
  6. Midha, A.; Dearden, S.; McCormack, R. EGFR mutation incidence in non-small-cell lung cancer of adenocarcinoma histology: A systematic review and global map by ethnicity (mutMapII). Am. J. Cancer Res. 2015, 5, 2892–2911. [Google Scholar]
  7. Cote, M.L.; Haddad, R.; Edwards, D.J.; Atikukke, G.; Gadgeel, S.; Soubani, A.O.; Lonardo, F.; Bepler, G.; Schwartz, A.G.; Ethier, S.P. Frequency and Type of Epidermal Growth Factor Receptor Mutations in African Americans with Non-small Cell Lung Cancer. J. Thorac. Oncol. 2011, 6, 627–630. [Google Scholar] [CrossRef] [Green Version]
  8. Karachaliou, N.; Fernandez-Bruno, M.; Bracht, J.W.P.; Rosell, R. EGFR first- and second-generation TKIs—there is still place for them in EGFR-mutant NSCLC patients. Transl. Cancer Res. 2018, 8, S23–S47. [Google Scholar] [CrossRef]
  9. Westover, D.; Zugazagoitia, J.; Cho, B.C.; Lovly, C.M.; Paz-Ares, L. Mechanisms of acquired resistance to first- and second-generation EGFR tyrosine kinase inhibitors. Ann. Oncol. 2018, 29, i10–i19. [Google Scholar] [CrossRef]
  10. Remon, J.; Steuer, C.; Ramalingam, S.; Felip, E. Osimertinib and other third-generation EGFR TKI in EGFR-mutant NSCLC patients. Ann. Oncol. 2018, 29, i20–i27. [Google Scholar] [CrossRef]
  11. Lee, K.-O.; Cha, M.Y.; Kim, M.; Song, J.Y.; Lee, J.-H.; Kim, Y.H.; Lee, Y.-M.; Suh, K.H.; Son, J. Abstract LB-100: Discovery of HM61713 as an orally available and mutant EGFR selective inhibitor. In Proceedings of the Experimental and Molecular Therapeutics; American Association for Cancer Research: Philadelphia, PA, USA, 2014. [Google Scholar]
  12. Mok, T.S.; Wu, Y.-L.; Ahn, M.-J.; Garassino, M.C.; Kim, H.R.; Ramalingam, S.S.; Shepherd, F.A.; He, Y.; Akamatsu, H.; Theelen, W.S.; et al. Osimertinib or Platinum–Pemetrexed in EGFR T790M–Positive Lung Cancer. N. Engl. J. Med. 2017, 376, 629–640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Soria, J.-C.; Ohe, Y.; Vansteenkiste, J.; Reungwetwattana, T.; Chewaskulyong, B.; Lee, K.H.; Dechaphunkul, A.; Imamura, F.; Nogami, N.; Kurata, T.; et al. Osimertinib in Untreated EGFR-Mutated Advanced Non–Small-Cell Lung Cancer. N. Engl. J. Med. 2018, 378, 113–125. [Google Scholar] [CrossRef] [PubMed]
  14. Wu, Y.-L.; Tsuboi, M.; He, J.; John, T.; Grohe, C.; Majem, M.; Goldman, J.W.; Laktionov, K.; Kim, S.-W.; Kato, T.; et al. Osimertinib in Resected EGFR-Mutated Non–Small-Cell Lung Cancer. N. Engl. J. Med. 2020, 383, 1711–1723. [Google Scholar] [CrossRef]
  15. Oxnard, G.R.; Yang, J.C.-H.; Yu, H.; Kim, S.-W.; Saka, H.; Horn, L.; Goto, K.; Ohe, Y.; Mann, H.; Thress, K.S.; et al. TATTON: A multi-arm, phase Ib trial of osimertinib combined with selumetinib, savolitinib, or durvalumab in EGFR-mutant lung cancer. Ann. Oncol. 2020, 31, 507–516. [Google Scholar] [CrossRef] [Green Version]
  16. Leonetti, A.; Sharma, S.; Minari, R.; Perego, P.; Giovannetti, E.; Tiseo, M. Resistance mechanisms to osimertinib in EGFR-mutated non-small cell lung cancer. Br. J. Cancer 2019, 121, 725–737. [Google Scholar] [CrossRef] [PubMed]
  17. Rangachari, D.; To, C.; Shpilsky, J.E.; VanderLaan, P.; Kobayashi, S.S.; Mushajiang, M.; Lau, C.J.; Paweletz, C.P.; Oxnard, G.R.; Jänne, P.A.; et al. EGFR-Mutated Lung Cancers Resistant to Osimertinib through EGFR C797S Respond to First-Generation Reversible EGFR Inhibitors but Eventually Acquire EGFR T790M/C797S in Preclinical Models and Clinical Samples. J. Thorac. Oncol. 2019, 14, 1995–2002. [Google Scholar] [CrossRef]
  18. Zhou, Z.; Zhao, Y.; Shen, S.; Gu, L.; Niu, X.; Xu, Y.; Zhang, T.; Xiang, J.; Mao, X.; Lu, S. Durable Clinical Response of Lung Adenocarcinoma Harboring EGFR 19Del/T790M/in trans-C797S to Combination Therapy of First- and Third-Generation EGFR Tyrosine Kinase Inhibitors. J. Thorac. Oncol. 2019, 14, e157–e159. [Google Scholar] [CrossRef]
  19. Wang, Y.; Yang, N.; Zhang, Y.; Li, L.; Han, R.; Zhu, M.; Feng, M.; Chen, H.; Lizaso, A.; Qin, T.; et al. Effective Treatment of Lung Adenocarcinoma Harboring EGFR-Activating Mutation, T790M, and cis-C797S Triple Mutations by Brigatinib and Cetuximab Combination Therapy. J. Thorac. Oncol. 2020, 15, 1369–1375. [Google Scholar] [CrossRef]
  20. Jänne, P.A.; Baik, C.; Su, W.-C.; Johnson, M.L.; Hayashi, H.; Nishio, M.; Kim, D.-W.; Koczywas, M.; Gold, K.A.; Steuer, C.E.; et al. Efficacy and Safety of Patritumab Deruxtecan (HER3-DXd) in EGFR Inhibitor–Resistant, EGFR-Mutated Non–Small Cell Lung Cancer. Cancer Discov. 2021, 12, 74–89. [Google Scholar] [CrossRef]
  21. Wang, S.; Song, Y.; Liu, D. EAI045: The fourth-generation EGFR inhibitor overcoming T790M and C797S resistance. Cancer Lett. 2016, 385, 51–54. [Google Scholar] [CrossRef]
  22. Jia, Y.; Yun, C.-H.; Park, E.; Ercan, D.; Manuia, M.; Juarez, J.; Xu, C.; Rhee, K.; Chen, T.; Zhang, H.; et al. Overcoming EGFR(T790M) and EGFR(C797S) resistance with mutant-selective allosteric inhibitors. Nature 2016, 534, 129–132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Kannan, S.; Venkatachalam, G.; Lim, H.H.; Surana, U.; Verma, C. Conformational landscape of the epidermal growth factor receptor kinase reveals a mutant specific allosteric pocket. Chem. Sci. 2018, 9, 5212–5222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. To, C.; Jang, J.; Chen, T.; Park, E.; Mushajiang, M.; De Clercq, D.J.; Xu, M.; Wang, S.; Cameron, M.D.; Heppner, D.E.; et al. Single and Dual Targeting of Mutant EGFR with an Allosteric Inhibitor. Cancer Discov. 2019, 9, 926–943. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Kashima, K.; Kawauchi, H.; Tanimura, H.; Tachibana, Y.; Chiba, T.; Torizawa, T.; Sakamoto, H. CH7233163 Overcomes Osimertinib-Resistant EGFR-Del19/T790M/C797S Mutation. Mol. Cancer Ther. 2020, 19, 2288–2297. [Google Scholar] [CrossRef] [PubMed]
  26. Schalm, S.; Dineen, T.; Lim, S.; Park, C.; Hsieh, J.; Woessner, R.; Zhang, Z.; Wilson, K.; Eno, M.; Wilson, D.; et al. 384P BLU-945, a highly potent and selective 4th generation EGFR TKI for the treatment of EGFR T790M/C797S resistant NSCLC. Ann. Oncol. 2020, 31, S1391. [Google Scholar] [CrossRef]
  27. Lim, S.M.; Park, C.W.; Zhang, Z.; Woessner, R.; Dineen, T.; Stevison, F.; Hsieh, J.; Eno, M.; Wilson, D.; Campbell, J.; et al. Abstract 1467: BLU-945, a fourth-generation, potent and highly selective epidermal growth factor receptor (EGFR) tyrosine kinase inhibitor (TKI) with intracranial activity, demonstrates robust in vivo antitumor activity in models of osimertinib-resistant non-small cell lung cancer (NSCLC). Cancer Res. 2021, 81, 1467. [Google Scholar] [CrossRef]
  28. Marcoux, N.; Gettinger, S.N.; O’kane, G.; Arbour, K.C.; Neal, J.W.; Husain, H.; Bonomi, P.H.; Evans, T.L.; Brahmer, J.R.; Muzikansky, A.; et al. EGFR-Mutant Adenocarcinomas That Transform to Small-Cell Lung Cancer and Other Neuroendocrine Carcinomas: Clinical Outcomes. J. Clin. Oncol. 2018, 37, 278–285. [Google Scholar] [CrossRef]
  29. A Study of the Combination of Osimertinib, Platinum and Etoposide for Patients with Metastatic EGFR Mutant Lung Cancers. NCT03567642. Available online: https://clinicaltrials.gov/ct2/show/NCT03567642 (accessed on 31 March 2022).
  30. Piper-Vallillo, A.J.; Sequist, L.V.; Piotrowska, Z. Emerging Treatment Paradigms for EGFR-Mutant Lung Cancers Progressing on Osimertinib: A Review. J. Clin. Oncol. 2020, 38, 2926–2936. [Google Scholar] [CrossRef]
  31. Sequist, L.V.; Han, J.-Y.; Ahn, M.-J.; Cho, B.C.; Yu, H.; Kim, S.-W.; Yang, J.C.-H.; Lee, J.S.; Su, W.-C.; Kowalski, D.; et al. Osimertinib plus savolitinib in patients with EGFR mutation-positive, MET-amplified, non-small-cell lung cancer after progression on EGFR tyrosine kinase inhibitors: Interim results from a multicentre, open-label, phase 1b study. Lancet Oncol. 2020, 21, 373–386. [Google Scholar] [CrossRef]
  32. Bauml, J.; Cho, B.C.; Park, K.; Lee, K.H.; Cho, E.K.; Kim, D.-W.; Kim, S.-W.; Haura, E.B.; Sabari, J.K.; Sanborn, R.E.; et al. Amivantamab in combination with lazertinib for the treatment of osimertinib-relapsed, chemotherapy-naïve EGFR mutant (EGFRm) non-small cell lung cancer (NSCLC) and potential biomarkers for response. J. Clin. Oncol. 2021, 39, 9006. [Google Scholar] [CrossRef]
  33. A Study of Tepotinib Plus Osimertinib in Osimertinib Relapsed MET Amplified NSCLC (INSIGHT 2). NCT03940703. Available online: https://www.clinicaltrials.gov/ct2/show/NCT03940703 (accessed on 31 March 2022).
  34. Giroux-Leprieur, E.; Dumenil, C.; Chinet, T. Combination of Crizotinib and Osimertinib or Erlotinib Might Overcome MET-Mediated Resistance to EGFR Tyrosine Kinase Inhibitor in EGFR-Mutated Adenocarcinoma. J. Thorac. Oncol. 2018, 13, e232–e234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Zhu, V.W.; Schrock, A.B.; Ali, S.M.; Ou, S.-H.I. Differential response to a combination of full-dose osimertinib and crizotinib in a patient with EGFR-mutant non-small cell lung cancer and emergent MET amplification. Lung Cancer Targets Ther. 2019, 10, 21–26. [Google Scholar] [CrossRef] [Green Version]
  36. Ross, H.J.; Blumenschein, G.R.; Aisner, J.; Damjanov, N.; Dowlati, A.; Garst, J.; Rigas, J.R.; Smylie, M.; Hassani, H.; Allen, K.E.; et al. Randomized Phase II Multicenter Trial of Two Schedules of Lapatinib as First- or Second-Line Monotherapy in Patients with Advanced or Metastatic Non–Small Cell Lung Cancer. Clin. Cancer Res. 2010, 16, 1938–1949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Hyman, D.M.; Piha-Paul, S.A.; Won, H.; Rodon, J.; Saura, C.; Shapiro, G.I.; Juric, D.; Quinn, D.I.; Moreno, V.; Doger, B.; et al. HER kinase inhibition in patients with HER2- and HER3-mutant cancers. Nature 2018, 554, 189–194. [Google Scholar] [CrossRef] [PubMed]
  38. Li, B.T.; Smit, E.F.; Goto, Y.; Nakagawa, K.; Udagawa, H.; Mazières, J.; Nagasaka, M.; Bazhenova, L.; Saltos, A.N.; Felip, E.; et al. Trastuzumab Deruxtecan in HER2-Mutant Non–Small-Cell Lung Cancer. N. Engl. J. Med. 2022, 386, 241–251. [Google Scholar] [CrossRef] [PubMed]
  39. Rehman, M.; Kim, C.; Reuss, J.E.; Kiedrowski, L.A.; Garg, R.J.; Liu, S.V. Divergent RET- and BRAF-Mediated Resistance to Osimertinib in EGFR-Mutant NSCLC: A Case Report. JCO Precis. Oncol. 2021, 939–942. [Google Scholar] [CrossRef] [PubMed]
  40. Ribeiro, M.F.S.A.; Knebel, F.H.; Bettoni, F.; Saddi, R.; Sacardo, K.P.; Canedo, F.S.N.A.; Alessi, J.V.M.; Shimada, A.K.; Marin, J.F.G.; Camargo, A.A.; et al. Impressive response to dabrafenib, trametinib, and osimertinib in a metastatic EGFR-mutant/BRAF V600E lung adenocarcinoma patient. NPJ Precis. Oncol. 2021, 5, 1–7. [Google Scholar] [CrossRef]
  41. Xie, Z.; Gu, Y.; Xie, X.; Lin, X.; Ouyang, M.; Qin, Y.; Zhang, J.; Lizaso, A.; Chen, S.; Zhou, C. Lung Adenocarcinoma Harboring Concomitant EGFR Mutations and BRAF V600E Responds to a Combination of Osimertinib and Vemurafenib to Overcome Osimertinib Resistance. Clin. Lung Cancer 2020, 22, e390–e394. [Google Scholar] [CrossRef]
  42. La Monica, S.; Minari, R.; Cretella, D.; Bonelli, M.; Fumarola, C.; Cavazzoni, A.; Galetti, M.; Digiacomo, G.; Riccardi, F.; Petronini, P.G.; et al. Acquired BRAF G469A Mutation as a Resistance Mechanism to First-Line Osimertinib Treatment in NSCLC Cell Lines Harboring an EGFR Exon 19 Deletion. Target. Oncol. 2019, 14, 619–626. [Google Scholar] [CrossRef]
  43. Eberlein, C.A.; Stetson, D.; Markovets, A.A.; Al-Kadhimi, K.J.; Lai, Z.; Fisher, P.R.; Meador, C.B.; Spitzler, P.; Ichihara, E.; Ross, S.J.; et al. Acquired Resistance to the Mutant-Selective EGFR Inhibitor AZD9291 Is Associated with Increased Dependence on RAS Signaling in Preclinical Models. Cancer Res. 2015, 75, 2489–2500. [Google Scholar] [CrossRef] [Green Version]
  44. Skoulidis, F.; Li, B.T.; Dy, G.K.; Price, T.J.; Falchook, G.S.; Wolf, J.; Italiano, A.; Schuler, M.; Borghaei, H.; Barlesi, F.; et al. Sotorasib for Lung Cancers with KRAS p.G12C Mutation. N. Engl. J. Med. 2021, 384, 2371–2381. [Google Scholar] [CrossRef] [PubMed]
  45. Ou, S.-H.I.; Jänne, P.A.; Leal, T.A.; Rybkin, I.I.; Sabari, J.K.; Barve, M.A.; Bazhenova, L.A.; Johnson, M.L.; Velastegui, K.L.; Cilliers, C.; et al. First-in-Human Phase I/IB Dose-Finding Study of Adagrasib (MRTX849) in Patients with Advanced KRASG12C Solid Tumors (KRYSTAL-1). J. Clin. Oncol. 2022. [Google Scholar] [CrossRef] [PubMed]
  46. Fang, W.; Huang, Y.; Gu, W.; Gan, J.; Wang, W.; Zhang, S.; Wang, K.; Zhan, J.; Yang, Y.; Huang, Y.; et al. PI3K-AKT-mTOR pathway alterations in advanced NSCLC patients after progression on EGFR-TKI and clinical response to EGFR-TKI plus everolimus combination therapy. Transl. Lung Cancer Res. 2020, 9, 1258–1267. [Google Scholar] [CrossRef] [PubMed]
  47. Fang, W.; Huang, Y.; Gan, J.; Yang, Y.; Wu, Y.; Huang, J.; Xu, Z.; Wang, W.; Zhang, L. Abstract 327: The impact of PIK3CA/PTEN/AKT1 genes in advanced NSCLC patients with acquired EGFR-TKI resistance and clinical response to EGFR-TKI plus everolimus combination therapy. In Proceedings of the Experimental and Molecular Therapeutics; American Association for Cancer Research: Philadelphia, PA, USA, 2019; p. 327. [Google Scholar]
  48. Vaclova, T.; Grazini, U.; Ward, L.; O’Neill, D.; Markovets, A.; Huang, X.; Chmielecki, J.; Hartmaier, R.; Thress, K.S.; Smith, P.D.; et al. Clinical impact of subclonal EGFR T790M mutations in advanced-stage EGFR-mutant non-small-cell lung cancers. Nat. Commun. 2021, 12, 1780. [Google Scholar] [CrossRef]
  49. Qin, Q.; Li, X.; Liang, X.; Zeng, L.; Wang, J.; Sun, L.; Zhong, D. CDK4/6 inhibitor palbociclib overcomes acquired resistance to third-generation EGFR inhibitor osimertinib in non-small cell lung cancer (NSCLC). Thorac. Cancer 2020, 11, 2389–2397. [Google Scholar] [CrossRef]
  50. La Monica, S.; Fumarola, C.; Cretella, D.; Bonelli, M.; Minari, R.; Cavazzoni, A.; Digiacomo, G.; Galetti, M.; Volta, F.; Mancini, M.; et al. Efficacy of the CDK4/6 Dual Inhibitor Abemaciclib in EGFR-Mutated NSCLC Cell Lines with Different Resistance Mechanisms to Osimertinib. Cancers 2020, 13, 6. [Google Scholar] [CrossRef]
  51. Koopman, L.A.; Terp, M.G.; Zom, G.G.; Janmaat, M.L.; Jacobsen, K.; Heuvel, E.G.-V.D.; Brandhorst, M.; Forssmann, U.; De Bree, F.; Pencheva, N.; et al. Enapotamab vedotin, an AXL-specific antibody-drug conjugate, shows preclinical antitumor activity in non-small cell lung cancer. JCI Insight 2019, 4, e128199. [Google Scholar] [CrossRef]
  52. Enapotamab Vedotin (HuMax-AXL-ADC) Safety Study in Patients with Solid Tumors. NCT02988817. Available online: https://clinicaltrials.gov/ct2/show/NCT02988817 (accessed on 31 March 2022).
  53. Hayakawa, D.; Takahashi, F.; Mitsuishi, Y.; Tajima, K.; Hidayat, M.; Winardi, W.; Ihara, H.; Kanamori, K.; Matsumoto, N.; Asao, T.; et al. Activation of insulin-like growth factor-1 receptor confers acquired resistance to osimertinib in non-small cell lung cancer withEGFRT790M mutation. Thorac. Cancer 2019, 11, 140–149. [Google Scholar] [CrossRef] [Green Version]
  54. Garon, E.; Johnson, M.; Lisberg, A.; Spira, A.; Yamamoto, N.; Heist, R.; Sands, J.; Yoh, K.; Meric-Bernstam, F.; Kitazono, S.; et al. LBA49 Efficacy of datopotamab deruxtecan (Dato-DXd) in patients (pts) with advanced/metastatic (adv/met) non-small cell lung cancer (NSCLC) and actionable genomic alterations (AGAs): Preliminary results from the phase I TROPION-PanTumor01 study. Ann. Oncol. 2021, 32, S1326–S1327. [Google Scholar] [CrossRef]
  55. Ercan, D.; Choi, H.G.; Yun, C.-H.; Capelletti, M.; Xie, T.; Eck, M.J.; Gray, N.S.; Jänne, P.A. EGFR Mutations and Resistance to Irreversible Pyrimidine-Based EGFR Inhibitors. Clin. Cancer Res. 2015, 21, 3913–3923. [Google Scholar] [CrossRef] [Green Version]
  56. Shi, P.; Oh, Y.-T.; Zhang, G.; Yao, W.; Yue, P.; Li, Y.; Kanteti, R.; Riehm, J.; Salgia, R.; Owonikoko, T.K.; et al. Met gene amplification and protein hyperactivation is a mechanism of resistance to both first and third generation EGFR inhibitors in lung cancer treatment. Cancer Lett. 2016, 380, 494–504. [Google Scholar] [CrossRef] [PubMed]
  57. Hsu, C.-C.; Liao, B.-C.; Liao, W.-Y.; Markovets, A.; Stetson, D.; Thress, K.; Yang, J.C.-H. Exon 16–Skipping HER2 as a Novel Mechanism of Osimertinib Resistance in EGFR L858R/T790M–Positive Non–Small Cell Lung Cancer. J. Thorac. Oncol. 2019, 15, 50–61. [Google Scholar] [CrossRef] [PubMed]
  58. Han, H.-S.; Lim, S.-N.; An, J.Y.; Lee, K.M.; Choe, K.H.; Kim, S.T.; Son, S.-M.; Choi, S.-Y.; Lee, H.-C.; Lee, O.-J. Detection of EGFR Mutation Status in Lung Adenocarcinoma Specimens with Different Proportions of Tumor Cells Using Two Methods of Differential Sensitivity. J. Thorac. Oncol. 2012, 7, 355–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Yu, Y.; Abudula, M.; Li, C.; Chen, Z.; Zhang, Y.; Chen, Y. Icotinib-resistant HCC827 cells produce exosomes with mRNA MET oncogenes and mediate the migration and invasion of NSCLC. Respir. Res. 2019, 20, 217. [Google Scholar] [CrossRef] [PubMed]
  60. Truini, A.; Starrett, J.H.; Stewart, T.F.; Ashtekar, K.D.; Walther, Z.; Wurtz, A.; Lu, D.; Park, J.H.; Deveaux, M.; Song, X.; et al. The EGFR Exon 19 Mutant L747-A750>P Exhibits Distinct Sensitivity to Tyrosine Kinase Inhibitors in Lung Adenocarcinoma. Clin. Cancer Res. 2019, 25, 6382–6391. [Google Scholar] [CrossRef]
  61. Weng, C.-H.; Chen, L.-Y.; Lin, Y.-C.; Shih, J.-Y.; Lin, Y.-C.; Tseng, R.-Y.; Chiu, A.-C.; Yeh, Y.-H.; Liu, C.; Lin, Y.-T.; et al. Epithelial-mesenchymal transition (EMT) beyond EGFR mutations per se is a common mechanism for acquired resistance to EGFR TKI. Oncogene 2018, 38, 455–468. [Google Scholar] [CrossRef]
  62. Nukaga, S.; Yasuda, H.; Tsuchihara, K.; Hamamoto, J.; Masuzawa, K.; Kawada, I.; Naoki, K.; Matsumoto, S.; Mimaki, S.; Ikemura, S.; et al. Amplification of EGFR Wild-Type Alleles in Non–Small Cell Lung Cancer Cells Confers Acquired Resistance to Mutation-Selective EGFR Tyrosine Kinase Inhibitors. Cancer Res. 2017, 77, 2078–2089. [Google Scholar] [CrossRef] [Green Version]
  63. Yu, Z.; Boggon, T.J.; Kobayashi, S.; Jin, C.; Ma, P.C.; Dowlati, A.; Kern, J.; Tenen, D.; Halmos, B. Resistance to an Irreversible Epidermal Growth Factor Receptor (EGFR) Inhibitor in EGFR-Mutant Lung Cancer Reveals Novel Treatment Strategies. Cancer Res. 2007, 67, 10417–10427. [Google Scholar] [CrossRef] [Green Version]
  64. Warmuth, M.; Kim, S.; Gu, X.-J.; Xia, G.; Adrián, F. Ba/F3 cells and their use in kinase drug discovery. Curr. Opin. Oncol. 2007, 19, 55–60. [Google Scholar] [CrossRef]
  65. Ramalingam, S.S.; Cheng, Y.; Zhou, C.; Ohe, Y.; Imamura, F.; Cho, B.C.; Lin, M.C.; Majem, M.; Shah, R.; Rukazenkov, Y.; et al. Mechanisms of acquired resistance to first-line osimertinib: Preliminary data from the phase III FLAURA study. Ann. Oncol. 2018, 29, viii740. [Google Scholar] [CrossRef]
  66. Papadimitrakopoulou, V.A.; Wu, Y.L.; Han, J.Y.; Ahn, M.J.; Ramalingam, S.S.; John, T.; Okamoto, I.; Yang, J.H.; Bulusu, K.C.; Laus, G.J.A.O.O.; et al. Analysis of resistance mechanisms to osimertinib in patients with EGFR T790M advanced NSCLC from the AURA3 study. Ann. Oncol. 2018, 29, viii741. [Google Scholar] [CrossRef]
  67. Oxnard, G.R.; Hu, E.; Mileham, K.F.; Husain, H.; Costa, D.; Tracy, P.; Feeney, N.; Sholl, L.M.; Dahlberg, S.E.; Redig, A.J.; et al. Assessment of Resistance Mechanisms and Clinical Implications in Patients with EGFRT790M–Positive Lung Cancer and Acquired Resistance to Osimertinib. JAMA Oncol. 2018, 4, 1527–1534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Niederst, M.J.; Hu, H.; Mulvey, H.E.; Lockerman, E.L.; Garcia, A.R.; Piotrowska, Z.; Sequist, L.V.; Engelman, J.A. The Allelic Context of the C797S Mutation Acquired upon Treatment with Third-Generation EGFR Inhibitors Impacts Sensitivity to Subsequent Treatment Strategies. Clin. Cancer Res. 2015, 21, 3924–3933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Wang, Z.; Yang, J.-J.; Huang, J.; Ye, J.-Y.; Zhang, X.-C.; Tu, H.-Y.; Han-Zhang, H.; Wu, Y.-L. Lung Adenocarcinoma Harboring EGFR T790M and In Trans C797S Responds to Combination Therapy of First- and Third-Generation EGFR TKIs and Shifts Allelic Configuration at Resistance. J. Thorac. Oncol. 2017, 12, 1723–1727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Uchibori, K.; Inase, N.; Araki, M.; Kamada, M.; Sato, S.; Okuno, Y.; Fujita, N.; Katayama, R. Brigatinib combined with anti-EGFR antibody overcomes osimertinib resistance in EGFR-mutated non-small-cell lung cancer. Nat. Commun. 2017, 8, 14768. [Google Scholar] [CrossRef] [Green Version]
  71. HERTHENA-Lung01: Patritumab Deruxtecan in Subjects with Metastatic or Locally Advanced EGFR-Mutated Non-Small Cell Lung Cancer. NCT04619004. Available online: https://clinicaltrials.gov/ct2/show/NCT04619004 (accessed on 31 March 2022).
  72. Patritumab Deruxtecan in Combination with Osimertinib in Subjects with Locally Advanced or Metastatic EGFR-Mutated Non-Small Cell Lung Cancer. NCT04676477. Available online: https://clinicaltrials.gov/ct2/show/NCT04676477 (accessed on 31 March 2022).
  73. Rotow, J.K.; Costa, D.B.; Paweletz, C.P.; Awad, M.M.; Marcoux, P.; Rangachari, D.; Barbie, D.A.; Sands, J.; Cheng, M.L.; Johnson, B.E.; et al. Concurrent osimertinib plus gefitinib for first-line treatment of EGFR-mutated non-small cell lung cancer (NSCLC). J. Clin. Oncol. 2020, 38, 9507. [Google Scholar] [CrossRef]
  74. Yang, Z.; Yang, N.; Ou, Q.; Xiang, Y.; Jiang, T.; Wu, X.; Bao, H.; Tong, X.; Wang, X.; Shao, Y.W.; et al. Investigating Novel Resistance Mechanisms to Third-Generation EGFR Tyrosine Kinase Inhibitor Osimertinib in Non–Small Cell Lung Cancer Patients. Clin. Cancer Res. 2018, 24, 3097–3107. [Google Scholar] [CrossRef] [Green Version]
  75. Zheng, D.; Hu, M.; Bai, Y.; Zhu, X.; Lu, X.; Wu, C.; Wang, J.; Liu, L.; Wang, Z.; Ni, J.; et al. EGFR G796D mutation mediates resistance to osimertinib. Oncotarget 2017, 8, 49671–49679. [Google Scholar] [CrossRef] [Green Version]
  76. Fairclough, S.R.; Kiedrowski, L.A.; Lin, J.J.; Zelichov, O.; Tarcic, G.; Stinchcombe, T.E.; Odegaard, J.I.; Lanman, R.B.; Shaw, A.T.; Nagy, R.J. Identification of osimertinib-resistant EGFR L792 mutations by cfDNA sequencing: Oncogenic activity assessment and prevalence in large cfDNA cohort. Exp. Hematol. Oncol. 2019, 8, 1–6. [Google Scholar] [CrossRef]
  77. Bersanelli, M.; Minari, R.; Bordi, P.; Gnetti, L.; Bozzetti, C.; Squadrilli, A.; Lagrasta, C.A.M.; Bottarelli, L.; Osipova, G.; Capelletto, E.; et al. L718Q Mutation as New Mechanism of Acquired Resistance to AZD9291 in EGFR-Mutated NSCLC. J. Thorac. Oncol. 2016, 11, e121–e123. [Google Scholar] [CrossRef] [Green Version]
  78. Yang, X.; Huang, C.; Chen, R.; Zhao, J. Resolving Resistance to Osimertinib Therapy with Afatinib in an NSCLC Patient with EGFR L718Q Mutation. Clin. Lung Cancer 2020, 21, e258–e260. [Google Scholar] [CrossRef] [PubMed]
  79. Schoenfeld, A.J.; Chan, J.; Kubota, D.; Sato, H.; Rizvi, H.; Daneshbod, Y.; Chang, J.C.; Paik, P.K.; Offin, M.; Arcila, M.E.; et al. Tumor Analyses Reveal Squamous Transformation and Off-Target Alterations As Early Resistance Mechanisms to First-line Osimertinib in EGFR-Mutant Lung Cancer. Clin. Cancer Res. 2020, 26, 2654–2663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Phase 2 Platform Study in Patients with Advanced Non-Small Lung Cancer Who Progressed on First-Line Osimertinib Therapy (ORCHARD). NCT03944772. Available online: https://www.clinicaltrials.gov/ct2/show/NCT03944772 (accessed on 31 March 2022).
  81. (SYMPHONY) Phase 1/2 Study Targeting EGFR Resistance Mechanisms in NSCLC. NCT04862780. Available online: https://clinicaltrials.gov/ct2/show/NCT04862780 (accessed on 31 March 2022).
  82. Liu, X.; Zhang, X.; Yang, L.; Tian, X.; Dong, T.; Ding, C.Z.; Hu, L.; Wu, L.; Zhao, L.; Mao, J.; et al. Abstract 1320: Preclinical evaluation of TQB3804, a potent EGFR C797S inhibitor. In Proceedings of the Experimental and Molecular Therapeutics, Atlanta, GA, USA, 29 March–3 April 2019; American Association for Cancer Research: Atlanta, GA, USA, 2019; p. 1320. [Google Scholar]
  83. Choe, H.; Jeon, B.U.; Jung, M.E.; Jeon, M.-K.; Shin, I.; Cho, B.C.; Choi, G.; Chae, C.H.; Lee, K. Structure-Activity Relationship Study of 2,4-Dianilinopyrimidine Containing Methanesulfonamide (TRE-069) as Potent and Selective Epidermal Growth Factor Receptor T790M/C797S Mutant Inhibitor for Anticancer Treatment. Bull. Korean Chem. Soc. 2017, 38, 1353–1357. [Google Scholar] [CrossRef]
  84. Quintanal-Villalonga, Á.; Chan, J.M.; Yu, H.A.; Pe’Er, D.; Sawyers, C.L.; Sen, T.; Rudin, C.M. Lineage plasticity in cancer: A shared pathway of therapeutic resistance. Nat. Rev. Clin. Oncol. 2020, 17, 360–371. [Google Scholar] [CrossRef]
  85. Piotrowska, Z.; Isozaki, H.; Lennerz, J.K.; Gainor, J.F.; Lennes, I.T.; Zhu, V.W.; Marcoux, N.; Banwait, M.K.; Digumarthy, S.R.; Su, W.; et al. Landscape of Acquired Resistance to Osimertinib in EGFR-Mutant NSCLC and Clinical Validation of Combined EGFR and RET Inhibition with Osimertinib and BLU-667 for Acquired RET Fusion. Cancer Discov. 2018, 8, 1529–1539. [Google Scholar] [CrossRef] [Green Version]
  86. Lee, J.; Kim, H.S.; Lee, B.; Kim, H.K.; Sun, J.; Ahn, J.S.; Ahn, M.; Park, K.; Lee, S. Genomic landscape of acquired resistance to third-generation EGFR tyrosine kinase inhibitors in EGFR T790M-mutant non–small cell lung cancer. Cancer 2020, 126, 2704–2712. [Google Scholar] [CrossRef]
  87. Zhang, Y.; Li, X.-Y.; Tang, Y.; Xu, Y.; Guo, W.-H.; Li, Y.-C.; Liu, X.-K.; Huang, C.-Y.; Wang, Y.-S.; Wei, Y.-Q. Rapid increase of serum neuron specific enolase level and tachyphylaxis of EGFR-tyrosine kinase inhibitor indicate small cell lung cancer transformation from EGFR positive lung adenocarcinoma? Lung Cancer 2013, 81, 302–305. [Google Scholar] [CrossRef]
  88. Jiang, S.-Y.; Zhao, J.; Wang, M.-Z.; Huo, Z.; Zhang, J.; Zhong, W.; Xu, Y. Small-Cell Lung Cancer Transformation in Patients with Pulmonary Adenocarcinoma. Medicine 2016, 95, e2752. [Google Scholar] [CrossRef]
  89. Niederst, M.J.; Sequist, L.V.; Poirier, J.T.; Mermel, C.H.; Lockerman, E.L.; Garcia, A.R.; Katayama, R.; Costa, C.; Ross, K.N.; Moran, T.; et al. RB loss in resistant EGFR mutant lung adenocarcinomas that transform to small-cell lung cancer. Nat. Commun. 2015, 6, 6377. [Google Scholar] [CrossRef]
  90. Offin, M.; Chan, J.M.; Tenet, M.; Rizvi, H.A.; Shen, R.; Riely, G.J.; Rekhtman, N.; Daneshbod, Y.; Quintanal-Villalonga, A.; Penson, A.; et al. Concurrent RB1 and TP53 Alterations Define a Subset of EGFR-Mutant Lung Cancers at risk for Histologic Transformation and Inferior Clinical Outcomes. J. Thorac. Oncol. 2019, 14, 1784–1793. [Google Scholar] [CrossRef]
  91. Vokes, N.I.; Chambers, E.; Nguyen, T.; Coolidge, A.; Lydon, C.A.; Le, X.; Sholl, L.; Heymach, J.V.; Nishino, M.; Van Allen, E.M.; et al. Concurrent TP53 mutations facilitate resistance evolution in EGFR mutant lung adenocarcinoma. J. Thorac. Oncol. 2022, S1556-0864, 135907. [Google Scholar] [CrossRef] [PubMed]
  92. Drilon, A.; Cappuzzo, F.; Ou, S.-H.I.; Camidge, D.R. Targeting MET in Lung Cancer: Will Expectations Finally Be MET? J. Thorac. Oncol. 2017, 12, 15–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Raghav, K.P.S.; Gonzalez-Augulo, A.M.; Blumenschein, G.R. Role of HGF/MET axis in resistance of lung cancer to contemporary management. Transl. Lung Cancer Res. 2012, 1, 179–193. [Google Scholar] [CrossRef]
  94. Engelman, J.A.; Zejnullahu, K.; Mitsudomi, T.; Song, Y.; Hyland, C.; Park, J.O.; Lindeman, N.; Gale, C.-M.; Zhao, X.; Christensen, J.; et al. MET Amplification Leads to Gefitinib Resistance in Lung Cancer by Activating ERBB3 Signaling. Science 2007, 316, 1039–1043. [Google Scholar] [CrossRef] [PubMed]
  95. Osimertinib Plus Savolitinib in EGFRm+/MET+ NSCLC Following Prior Osimertinib (SAVANNAH). NCT03778229. Available online: https://clinicaltrials.gov/ct2/show/NCT03778229 (accessed on 31 March 2022).
  96. Park, K.; Haura, E.B.; Leighl, N.B.; Mitchell, P.; Shu, C.A.; Girard, N.; Viteri, S.; Han, J.-Y.; Kim, S.-W.; Lee, C.K.; et al. Amivantamab in EGFR Exon 20 Insertion–Mutated Non–Small-Cell Lung Cancer Progressing on Platinum Chemotherapy: Initial Results From the CHRYSALIS Phase I Study. J. Clin. Oncol. 2021, 39, 3391–3402. [Google Scholar] [CrossRef]
  97. York, E.R.; Varella-Garcia, M.; Bang, T.J.; Aisner, D.L.; Camidge, D.R. Tolerable and Effective Combination of Full-Dose Crizotinib and Osimertinib Targeting MET Amplification Sequentially Emerging after T790M Positivity in EGFR- Mutant Non–Small Cell Lung Cancer. J. Thorac. Oncol. 2017, 12, e85–e88. [Google Scholar] [CrossRef] [Green Version]
  98. Janne, P.A.; Baik, C.S.; Su, W.-C.; Johnson, M.L.; Hayashi, H.; Nishio, M.; Kim, D.-W.; Koczywas, M.; Gold, K.A.; Steuer, C.E.; et al. Efficacy and safety of patritumab deruxtecan (HER3-DXd) in EGFR inhibitor-resistant, EGFR-mutated (EGFRm) non-small cell lung cancer (NSCLC). J. Clin. Oncol. 2021, 39, 9007. [Google Scholar] [CrossRef]
  99. Yan, M.; Schwaederle, M.; Arguello, D.; Millis, S.Z.; Gatalica, Z.; Kurzrock, R. HER2 expression status in diverse cancers: Review of results from 37,992 patients. Cancer Metastasis Rev. 2015, 34, 157–164. [Google Scholar] [CrossRef] [Green Version]
  100. Lin, Y.-C.; Boone, M.; Meuris, L.; Lemmens, I.; Van Roy, N.; Soete, A.; Reumers, J.; Moisse, M.; Plaisance, S.; Drmanac, R.T.; et al. Genome dynamics of the human embryonic kidney 293 lineage in response to cell biology manipulations. Nat. Commun. 2014, 5, 4767. [Google Scholar] [CrossRef] [Green Version]
  101. Wu, L.; Ke, L.; Zhang, Z.; Yu, J.; Meng, X. Development of EGFR TKIs and Options to Manage Resistance of Third-Generation EGFR TKI Osimertinib: Conventional Ways and Immune Checkpoint Inhibitors. Front. Oncol. 2020, 10, 2778. [Google Scholar] [CrossRef]
  102. Du, X.; Yang, B.; An, Q.; Assaraf, Y.G.; Cao, X.; Xia, J. Acquired resistance to third-generation EGFR-TKIs and emerging next-generation EGFR inhibitors. Innovation 2021, 2, 100103. [Google Scholar] [CrossRef] [PubMed]
  103. Zhao, J.; Xia, Y. Targeting HER2 Alterations in Non–Small-Cell Lung Cancer: A Comprehensive Review. JCO Precis. Oncol. 2020, 4, 411–425. [Google Scholar] [CrossRef] [PubMed]
  104. Rich, T.A.; Reckamp, K.L.; Chae, Y.K.; Doebele, R.C.; Iams, W.T.; Oh, M.; Raymond, V.M.; Lanman, R.B.; Riess, J.W.; Stinchcombe, T.E.; et al. Analysis of Cell-Free DNA from 32,989 Advanced Cancers Reveals Novel Co-occurring Activating RET Alterations and Oncogenic Signaling Pathway Aberrations. Clin. Cancer Res. 2019, 25, 5832–5842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Mizukami, T.; Shiraishi, K.; Shimada, Y.; Ogiwara, H.; Tsuta, K.; Ichikawa, H.; Sakamoto, H.; Kato, M.; Shibata, T.; Nakano, T.; et al. Molecular Mechanisms Underlying Oncogenic RET Fusion in Lung Adenocarcinoma. J. Thorac. Oncol. 2014, 9, 622–630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Iams, W.; Chae, Y. P3.02-034 Acquired Resistance to Osimertinib by CCDC6-RET Fusion in a Patient with EGFR T790M Mutant Metastatic Lung Adenocarcinoma. J. Thorac. Oncol. 2017, 12, S2249–S2250. [Google Scholar] [CrossRef]
  107. Sun, Y.; Pei, L.; Luo, N.; Chen, D.; Meng, L. A Novel MYH9-RET Fusion Occurrence and EGFR T790M Loss as an Acquired Resistance Mechanism to Osimertinib in a Patient with Lung Adenocarcinoma: A Case Report. OncoTargets Ther. 2020, 13, 11177–11181. [Google Scholar] [CrossRef]
  108. Klempner, S.; Bazhenova, L.A.; Braiteh, F.S.; Nikolinakos, P.G.; Gowen, K.; Cervantes, C.M.; Chmielecki, J.; Greenbowe, J.R.; Ross, J.S.; Stephens, P.J.; et al. Emergence of RET rearrangement co-existing with activated EGFR mutation in EGFR-mutated NSCLC patients who had progressed on first- or second-generation EGFR TKI. Lung Cancer 2015, 89, 357–359. [Google Scholar] [CrossRef]
  109. Subbiah, V.; Baik, C.; Kirkwood, J.M. Clinical Development of BRAF plus MEK Inhibitor Combinations. Trends Cancer 2020, 6, 797–810. [Google Scholar] [CrossRef]
  110. Ho, C.-C.; Liao, W.-Y.; Lin, C.-A.; Shih, J.-Y.; Yu, C.-J.; Yang, J.C.-H. Acquired BRAF V600E Mutation as Resistant Mechanism after Treatment with Osimertinib. J. Thorac. Oncol. Off. Publ. Int. Assoc. Study Lung Cancer 2017, 12, 567–572. [Google Scholar] [CrossRef] [Green Version]
  111. Huang, Y.; Gan, J.; Guo, K.; Deng, Y.; Fang, W. Acquired BRAF V600E Mutation Mediated Resistance to Osimertinib and Responded to Osimertinib, Dabrafenib, and Trametinib Combination Therapy. J. Thorac. Oncol. Off. Publ. Int. Assoc. Study Lung Cancer 2019, 14, e236–e237. [Google Scholar] [CrossRef]
  112. Roper, N.; Brown, A.-L.; Wei, J.S.; Pack, S.; Trindade, C.; Kim, C.; Restifo, O.; Gao, S.; Sindiri, S.; Mehrabadi, F.; et al. Clonal Evolution and Heterogeneity of Osimertinib Acquired Resistance Mechanisms in EGFR Mutant Lung Cancer. Cell Rep. Med. 2020, 1, 100007. [Google Scholar] [CrossRef] [PubMed]
  113. Vojnic, M.; Kubota, D.; Kurzatkowski, C.; Offin, M.; Suzawa, K.; Benayed, R.; Schoenfeld, A.J.; Plodkowski, A.J.; Poirier, J.; Rudin, C.M.; et al. Acquired BRAF Rearrangements Induce Secondary Resistance to EGFR therapy in EGFR-Mutated Lung Cancers. J. Thorac. Oncol. 2019, 14, 802–815. [Google Scholar] [CrossRef] [PubMed]
  114. Dagogo-Jack, I.; Piotrowska, Z.; Cobb, R.; Banwait, M.; Lennerz, J.K.; Hata, A.N.; Digumarthy, S.R.; Sequist, L.V. Response to the Combination of Osimertinib and Trametinib in a Patient with EGFR-Mutant NSCLC Harboring an Acquired BRAF Fusion. J. Thorac. Oncol. 2019, 14, e226–e228. [Google Scholar] [CrossRef] [PubMed]
  115. Uprety, D.; Adjei, A.A. KRAS: From undruggable to a druggable Cancer Target. Cancer Treat. Rev. 2020, 89, 102070. [Google Scholar] [CrossRef]
  116. Shi, Y.; Xing, P.; Han, X.; Wang, S.; Liu, Y.; Liu, P.; Li, J.; Chang, L.; Guan, Y.; Zhang, Z.; et al. P1.13-18 Exploring the Resistance Mechanism of Osimertinib and Monitoring the Treatment Response Using Plasma ctDNA in Chinese NSCLC Patients. J. Thorac. Oncol. 2018, 13, S589. [Google Scholar] [CrossRef] [Green Version]
  117. Ortiz-Cuaran, S.; Scheffler, M.; Plenker, D.; Dahmen, L.; Scheel, A.H.; Fernandez-Cuesta, L.; Meder, L.; Lovly, C.M.; Persigehl, T.; Merkelbach-Bruse, S.; et al. Heterogeneous Mechanisms of Primary and Acquired Resistance to Third-Generation EGFR Inhibitors. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2016, 22, 4837–4847. [Google Scholar] [CrossRef] [Green Version]
  118. Hong, M.; Kim, M.; Kim, S.-Y.; Heo, S.; Kang, H.-N.; Park, C.-W.; Barrett, J.; Stetson, D.; Chmielecki, J.; Markovets, A.; et al. Molecular landscape of osimertinib resistance revealed by targeted panel sequencing and patient-derived cancer models in non-small cell lung cancer patients. Ann. Oncol. 2018, 29, viii516. [Google Scholar] [CrossRef]
  119. 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]
  120. Fay, A.P.; Moreira, R.B.; Filho, P.R.S.N.; Albuquerque, C.; Barrios, C.H. The management of immune-related adverse events associated with immune checkpoint blockade. Expert Rev. Qual. Life Cancer Care 2016, 1, 89–97. [Google Scholar] [CrossRef] [Green Version]
  121. Choi, Y.J.; Anders, L. Signaling through cyclin D-dependent kinases. Oncogene 2014, 33, 1890–1903. [Google Scholar] [CrossRef] [Green Version]
  122. Osoegawa, A.; Yamaguchi, M.; Nakamura, T.; Morinaga, R.; Tanaka, K.; Kashiwabara, K.; Miura, T.; Suetsugu, T.; Harada, T.; Asoh, T.; et al. High Incidence of C797S Mutation in Patients with Long Treatment History of EGFR Tyrosine Kinase Inhibitors Including Osimertinib. JTO Clin. Res. Rep. 2021, 2, 100191. [Google Scholar] [CrossRef] [PubMed]
  123. Blakely, C.M.; Watkins, T.B.K.; Wu, W.; Gini, B.; Chabon, J.J.; McCoach, C.E.; McGranahan, N.; Wilson, G.A.; Birkbak, N.; Olivas, V.R.; et al. Evolution and clinical impact of co-occurring genetic alterations in advanced-stage EGFR-mutant lung cancers. Nat. Genet. 2017, 49, 1693–1704. [Google Scholar] [CrossRef] [PubMed]
  124. G1T38, a CDK 4/6 Inhibitor, in Combination with Osimertinib in EGFR-Mutant Non-Small Cell Lung Cancer. NCT03455829. Available online: https://clinicaltrials.gov/ct2/show/NCT03455829 (accessed on 31 March 2022).
  125. Antony, J.; Huang, R.Y.-J. AXL-Driven EMT State as a Targetable Conduit in Cancer. Cancer Res. 2017, 77, 3725–3732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Nonagase, Y.; Takeda, M.; Azuma, K.; Hayashi, H.; Haratani, K.; Tanaka, K.; Yonesaka, K.; Ishii, H.; Hoshino, T.; Nakagawa, K. Tumor tissue and plasma levels of AXL and GAS6 before and after tyrosine kinase inhibitor treatment in EGFR-mutated non–small cell lung cancer. Thorac. Cancer 2019, 10, 1928–1935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Yang, Y.M.; Jang, Y.; Lee, S.H.; Kang, B.; Lim, S.M. AXL/MET Dual Inhibitor, CB469, Has Activity in Non-Small Cell Lung Cancer with Acquired Resistance to EGFR TKI with AXL or MET Activation; Elsevier Ireland Ltd.: Shannon, Ireland, 2020; Volume 146, ISBN 8231780343. [Google Scholar]
  128. Namba, K.; Shien, K.; Takahashi, Y.; Torigoe, H.; Sato, H.; Yoshioka, T.; Takeda, T.; Kurihara, E.; Ogoshi, Y.; Yamamoto, H.; et al. Activation of AXL as a Preclinical Acquired Resistance Mechanism Against Osimertinib Treatment in EGFR-Mutant Non–Small Cell Lung Cancer Cells. Mol. Cancer Res. 2018, 17, 499–507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Taniguchi, H.; Yamada, T.; Wang, R.; Tanimura, K.; Adachi, Y.; Nishiyama, A.; Tanimoto, A.; Takeuchi, S.; Araujo, L.H.; Boroni, M.; et al. AXL confers intrinsic resistance to osimertinib and advances the emergence of tolerant cells. Nat. Commun. 2019, 10, 1–14. [Google Scholar] [CrossRef]
  130. Kim, T.M.; Song, A.; Kim, D.-W.; Kim, S.; Ahn, Y.-O.; Keam, B.; Jeon, Y.K.; Lee, S.-H.; Chung, D.H.; Heo, D.S. Mechanisms of Acquired Resistance to AZD9291: A Mutation-Selective, Irreversible EGFR Inhibitor. J. Thorac. Oncol. 2015, 10, 1736–1744. [Google Scholar] [CrossRef] [Green Version]
  131. Taniguchi, Y.; Yamada, T.; Yoshimura, A.; Kaira, K.; Atagi, S.; Yano, S.; Takayama, K. 411P Impact of pre-treatment AXL expression on osimertinib efficacy in patients with non-small cell lung cancer with EGFR mutation. Ann. Oncol. 2020, 31, S1402–S1403. [Google Scholar] [CrossRef]
  132. Li, J.; Choi, E.; Yu, H.; Bai, X.-C. Structural basis of the activation of type 1 insulin-like growth factor receptor. Nat. Commun. 2019, 10, 4567. [Google Scholar] [CrossRef]
  133. Manabe, T.; Yasuda, H.; Terai, H.; Kagiwada, H.; Hamamoto, J.; Ebisudani, T.; Kobayashi, K.; Masuzawa, K.; Ikemura, S.; Kawada, I.; et al. IGF2 Autocrine-Mediated IGF1R Activation Is a Clinically Relevant Mechanism of Osimertinib Resistance in Lung Cancer. Mol. Cancer Res. 2020, 18, 549–559. [Google Scholar] [CrossRef] [Green Version]
  134. Ohara, S.; Suda, K.; Mitsudomi, T. Cell Line Models for Acquired Resistance to First-Line Osimertinib in Lung Cancers—Applications and Limitations. Cells 2021, 10, 354. [Google Scholar] [CrossRef] [PubMed]
  135. Zhu, X.; Chen, L.; Liu, L.; Niu, X. EMT-Mediated Acquired EGFR-TKI Resistance in NSCLC: Mechanisms and Strategies. Front. Oncol. 2019, 9, 1044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Jiang, X.-M.; Xu, Y.-L.; Yuan, L.-W.; Zhang, L.-L.; Huang, M.-Y.; Ye, Z.-H.; Su, M.-X.; Chen, X.-P.; Zhu, H.; Ye, R.D.; et al. TGFβ2-mediated epithelial–mesenchymal transition and NF-κB pathway activation contribute to osimertinib resistance. Acta Pharmacol. Sin. 2020, 42, 451–459. [Google Scholar] [CrossRef] [PubMed]
  137. Qin, Q.; Li, X.; Liang, X.; Zeng, L.; Wang, J.; Sun, L.; Zhong, D. Targeting the EMT transcription factor Snail overcomes resistance to osimertinib in EGFR-mutant non-small cell lung cancer. Thorac. Cancer 2021, 12, 1708–1715. [Google Scholar] [CrossRef]
  138. Zeng, L.; Yang, N.; Zhang, Y. GOPC-ROS1 Rearrangement as an Acquired Resistance Mechanism to Osimertinib and Responding to Crizotinib Combined Treatments in Lung Adenocarcinoma. J. Thorac. Oncol. 2018, 13, e114–e116. [Google Scholar] [CrossRef] [Green Version]
  139. Schrock, A.B.; Zhu, V.W.; Hsieh, W.-S.; Madison, R.; Creelan, B.; Silberberg, J.; Costin, D.; Bharne, A.; Bonta, I.; Bosemani, T.; et al. Receptor Tyrosine Kinase Fusions and BRAF Kinase Fusions are Rare but Actionable Resistance Mechanisms to EGFR Tyrosine Kinase Inhibitors. J. Thorac. Oncol. 2018, 13, 1312–1323. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Resistance mechanisms to first- and second-line osimertinib. Adapted and reprinted from Leonetti et al. [16], Copyright © 2019, under exclusive license to Cancer Research UK. Cell cycle gene alterations (CCND1amp, CCND2amp, CCNE1amp, CDK6amp, CDKN2A E2fs), MAPK-PI3K mutations (BRAF V600E, PIK3CA, KRAS), oncogenic fusions (FGFR, NTRK, RET, ALK, BRAF), amplifications (MET, HER2, PIK3CA), EGFR mutations (C797X, L792X, G796X, exon20 insertions, etc.), histologic transformations (SCLC and SCC).
Figure 1. Resistance mechanisms to first- and second-line osimertinib. Adapted and reprinted from Leonetti et al. [16], Copyright © 2019, under exclusive license to Cancer Research UK. Cell cycle gene alterations (CCND1amp, CCND2amp, CCNE1amp, CDK6amp, CDKN2A E2fs), MAPK-PI3K mutations (BRAF V600E, PIK3CA, KRAS), oncogenic fusions (FGFR, NTRK, RET, ALK, BRAF), amplifications (MET, HER2, PIK3CA), EGFR mutations (C797X, L792X, G796X, exon20 insertions, etc.), histologic transformations (SCLC and SCC).
Cancers 14 01931 g001
Table 1. Mechanism of resistance to osimertinib and proposed therapeutic options.
Table 1. Mechanism of resistance to osimertinib and proposed therapeutic options.
Mechanism of ResistanceTherapeutic StrategiesReferences
C797XGefitinib, erlotinib[17]
Osimertinib + erlotinib[18]
Brigatinib + cetuximab[19]
Patritumab deruxtecan[20]
EAI045[21,22,23]
JBJ-04-125-02[24]
CH7233163[25]
BLU-945[26,27]
Small cell transformationPlatinum-etoposide[28,29]
Squamous cell transformationHistology based approach[30]
MET amplificationOsimertinib + savolitinib[31]
Lazertinib + amivantamab [32]
Tepotinib + osimertinib[33]
Osimertinib + crizotinib[34,35]
Patritumab deruxtecan[20]
HER2 alterationsPatritumab deruxtecan[20]
Osimertinib + lapatinib *[36]
Osimertinib + neratinib *[37]
Osimertinib + T-DXd *[38]
RET alterationsOsimertinib + selpercatinib [39]
BRAF alterationsOsimertinib + dabrafenib + trametinib[40]
Osimertinib + vemurafenib[41]
Osimertinib + selumetinib or trametinib[42]
RASOsimertinib + selumetinib or Aurora kinase b inhibitor[43]
Osimertinib + sotorasib *[44]
Osimertinib + adagrasib *[45]
PIK3EGFR TKIs and everolimus[46,47]
Osimertinib + alpelisib[48]
Cell cycle gene alterationsOsimertinib + palbociclib[49]
Osimertinib + abemaciclib[50]
AXL overexpressionEnapotamab vedotin[51,52]
IGF-1 receptor activationOsimertinib + linsitinib[53]
Non-specific alterationsDatopotamab deruxtecan[54]
* These proposed therapeutic mechanisms to overcome acquired resistance to osimertinib are yet to be tested. PI3K: Phosphoinositide 3-kinase, IGF-1: Insulin-like Growth Factor-1.
Table 2. Cell lines used to study resistance to osimertinib.
Table 2. Cell lines used to study resistance to osimertinib.
Induced Resistance MechanismCell LinesMechanism of InductionReferences
del19/T790M/C797S and L858R/T790M/C797SNIH3T3 cells (immortalized mouse embryonic fibroblast cell line)Transduction with lentiviruses[25]
del19/L858R +/− T790MBa/F3 cells (a murine, IL-3 dependent, hematopoietic cell line)Transduction with retroviral JP1540 or lentiviral JP1698 vectors[55]
MET amplificationHCC827 cells (EGFR del19) Exposure to osimertinib through a stepwise escalation process[56]
HER2 exon 16 skippingHEK293 cells (human embryonic kidney cell line and H1975 (T790M/L858R)Plasmid transfection[57]
BRAF G469APC9 cells (EGFR del 19)Exposure to osimertinib through a stepwise escalation process[42]
RAS alterationsPC9 cellsExposure to osimertinib through a stepwise escalation process and a single concentration of osimertinib [43]
Cell cycle gene alterationsH1975 cells (EGFR L858R/T790M)Exposure to osimertinib through a stepwise escalation process[49]
AXL overexpressionHCC827 cells (EGFR del19), PC9, H1975, and HCC4006 cells (EGFR del19)Exposure to osimertinib through a stepwise escalation process a single concentration of osimertinib [58,59,60]
Activation of IGF-1 receptorPC9 cellsExposure to gefitinib, developing resistance through the T790M, subsequently culture with stepwise escalation with osimertinib [53]
H1975 cellsExposure to osimertinib using a high-concentration method [53]
EMTH1975/AR cells (gefitinib resistant)Exposure to osimertinib through a stepwise escalation process[61]
Other mechanisms:
Src-AKT pathway and EGFR wild-type amplification
PC9 and H1975 cellsExposure to osimertinib through a stepwise escalation process[62]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ríos-Hoyo, A.; Moliner, L.; Arriola, E. Acquired Mechanisms of Resistance to Osimertinib—The Next Challenge. Cancers 2022, 14, 1931. https://doi.org/10.3390/cancers14081931

AMA Style

Ríos-Hoyo A, Moliner L, Arriola E. Acquired Mechanisms of Resistance to Osimertinib—The Next Challenge. Cancers. 2022; 14(8):1931. https://doi.org/10.3390/cancers14081931

Chicago/Turabian Style

Ríos-Hoyo, Alejandro, Laura Moliner, and Edurne Arriola. 2022. "Acquired Mechanisms of Resistance to Osimertinib—The Next Challenge" Cancers 14, no. 8: 1931. https://doi.org/10.3390/cancers14081931

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