Skip to main content
Erschienen in: Journal of Hematology & Oncology 1/2022

Open Access 01.12.2022 | Review

Emerging strategies to overcome resistance to third-generation EGFR inhibitors

verfasst von: Kunyu Shi, Guan Wang, Junping Pei, Jifa Zhang, Jiaxing Wang, Liang Ouyang, Yuxi Wang, Weimin Li

Erschienen in: Journal of Hematology & Oncology | Ausgabe 1/2022

insite
INHALT
download
DOWNLOAD
print
DRUCKEN
insite
SUCHEN

Abstract

Epidermal growth factor receptor (EGFR), the receptor for members of the epidermal growth factor family, regulates cell proliferation and signal transduction; moreover, EGFR is related to the inhibition of tumor cell proliferation, angiogenesis, invasion, metastasis, and apoptosis. Therefore, EGFR has become an important target for the treatment of cancer, including non-small cell lung cancer, head and neck cancer, breast cancer, glioma, cervical cancer, and bladder cancer. First- to third-generation EGFR inhibitors have shown considerable efficacy and have significantly improved disease prognosis. However, most patients develop drug resistance after treatment. The challenge of overcoming intrinsic and acquired resistance in primary and recurrent cancer mediated by EGFR mutations is thus driving the search for alternative strategies in the design of new therapeutic agents. In view of resistance to third-generation inhibitors, understanding the intricate mechanisms of resistance will offer insight for the development of more advanced targeted therapies. In this review, we discuss the molecular mechanisms of resistance to third-generation EGFR inhibitors and review recent strategies for overcoming resistance, new challenges, and future development directions.
Hinweise
Kunyu Shi, Guan Wang and Junping Pei contributed equally to this work

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
ACK1
Activated Cdc42-associated kinase 1
ADC
Antibody drug conjugates
AKR1B1
Aldehyde–ketone reductase family 1 member B1
A-loop
Activation loop
ATP
Adenosine triphosphate
AURK
Aurora kinases
AUTAC
Autophagy-targeting chimera
Bcl-2
B cell lymphoma-2
BRAF
V-RAF murine sarcoma viral oncogene homolog B1
CCND
Cyclin D
CCNE1
Cyclin E1
CDK4/6
Cyclin-dependent kinase 4/6
cIAP1
Cellular inhibitor of apoptosis protein 1
CRBN
Cereblon
DFG
Asp-Phe-Gly
DZ-SIM
DZ-SIMvastatin
EGF
Epidermal growth factor
EGFR
Epidermal growth factor receptor
EGFR-TKIs
Epidermal growth factor receptor tyrosine kinase inhibitors
EMT
Epithelial–mesenchymal transformation
EMT-TFs
EMT-induced transcription factors
ERC1
Excision repair cross-complementation 1
FGF
Fibroblast growth factor
FGFR
Fibroblast growth factor receptor
Fv
Variable region fragment
Grp94
Glucose regulatory protein 94
HB-EGF
EGF-like growth factor
HMGCR
3-Hydroxy-3-methylglutaryl-CoA reductase
HNK
Honokiol
ICIs
Immune checkpoint inhibitors
IGF1R
Insulin-like growth factor receptor 1
LUAD
Lung adenocarcinoma
MDM2
Mouse double minute 2
NCCN
National Comprehensive Cancer Network
NSCLC
Non-small cell lung cancer
NTRK1
Neurotrophic tyrosine receptor kinase 1
PARP
Poly(ADP-ribose) polymerase
PGAM1
Phosphoglycerate mutase 1
PIK3CA
Phosphatidylinositol 4,5-bisphosphate 3-kinase catalytic subunit
PROTAC
Proteolysis-targeting chimera
PTEN
Phosphatase and tensin homolog
ROS
Reactive oxygen species
RTK
Receptor tyrosine kinase
SCLC
Small cell lung cancer
SLC7A11
Solute carrier family 7 member 11
SPP1
Secreted phosphoprotein 1
STAT
Signal sensor and transcription activator
TACC3
Transforming acid helix protein 3
TGF
Transforming growth factor
TMP3
Thrombopoietin mimetic peptide 3
VHL
Von Hippel–Lindau
WT
Wild-type
WT-EGFR
Wild-type EGFR

Introduction

Epidermal growth factor receptor (EGFR) is a member of the receptor tyrosine kinase (RTK) superfamily that consists of exon boundaries and associated extracellular, transmembrane, and intracellular protein domains. EGFR is involved in multiple signaling pathways and regulates numerous cell functions (Fig. 1A). This transmembrane glycoprotein is composed of a cysteine-rich extracellular ligand binding domain, hydrophobic transmembrane domain, cytoplasmic RTK domain, and C-terminal domain. The RTK domain contains an N-lobe consisting of five β-sheet strands and one αC helix and a C-lobe containing the main helices of a highly flexible activation loop (A-loop) [1]. The deep cleft at the junction of these two lobes forms the binding pocket for the adenine ring of ATP. The conformation of three conserved structural elements, namely the Asp-Phe-Gly (DFG) motif, αC helix, and A-loop, critically regulates the activation or inactivation of the catalytic domain. When EGFR is in the active state, the important catalytic residue D855 is located in the ATP binding site, stabilizing the ATP-loaded complex (DFG-in) and αC helix (αC-in). In the inactive state, EGFR forms a Src-like structure, including a closed A-loop, αC-out, and DFG-in [2]. (Fig. 1B). EGFR can dimerize upon binding by ligands, such as amphiregulin, β-cytokines, epidermal growth factor (EGF), heparin-binding EGF-like growth factor (HB-EGF), and transforming growth factor (TGF). The activation of the intracellular tyrosine kinase domain and autophosphorylation, which initiates the Ras/RAF/MEK, signal transducer and activator of transcription (STAT), PI3K/AKT/mTOR and other downstream signaling pathways, are closely related to embryonic development and stem cell division [24]. Overexpression of wild-type (WT) EGFR protein with or without EGFR gene amplification or a kinase-activating mutation further enhances cell proliferation, migration, survival, and antiapoptotic responses through signaling cascades, and these processes are closely related to the occurrence and development of many types of epithelial-derived cancer, such as non-small cell lung cancer (NSCLC), breast cancer, glioma, head and neck cancer, cervical cancer, and bladder cancer. Among these cancers, lung cancer appears to be the most common and has the characteristics of aberrant proliferation, metastasis, and drug resistance [58]. Thus, EGFR has become a promising target for anticancer drug design and development. EGFR tyrosine kinase inhibitors (EGFR-TKIs) have achieved remarkable results in the clinic [9]. However, most patients develop acquired drug resistance to first- and second-generation EGFR-TKIs after 1–2 years. The mechanism of drug resistance for nearly half of cases relates to the T790M mutation. Third-generation EGFR-TKIs that target EGFR-TKI-sensitive mutations and the T790M mutation have been developed [10].
Unfortunately, drug resistance caused by less-common mutations in the EGFR gene and components of signal transduction pathways continues to emerge. In addition to common secondary (T790M) and tertiary (C797S) mutations, other EGFR mutations (such as the L718Q, L796S, and L792H mutations and the exon 20 insertion), MET amplification, phosphatidylinositol 4,5-bisphosphate 3-kinase catalytic subunit alpha (PIK3CA) mutations, HER2 amplification, oncogene fusions, and alterations in cell cycle-related genes have been observed [11] (Fig. 1C). There is an urgent need for better strategies to combat the inevitable molecular-targeted drug resistance associated with third-generation inhibitors. This review aims to provide a comprehensive overview of the mechanisms of resistance to third-generation EGFR-TKIs and to explore new insights and strategies for overcoming acquired resistance.

Third-generation EGFR-TKIs and drug resistance mechanisms

The development of third-generation EGFR-TKIs

The first-generation EGFR-TKIs form hydrogen bonds with Met793 in the ATP binding pocket of EGFR and reversibly compete with ATP for binding. Drug resistance occurs due to the EGFR T790M mutation (Thr790 in the hydrophobic ATP binding site encoded on exon 20 is replaced by methionine), subclonal selection (of a genetically resistant clone), and rare EGFR mutations (such as G719X, S768I, and L861Q). Thereafter, the development of second-generation EGFR-TKIs was reported; these inhibitors have the same quinazoline scaffold as first-generation EGFR-TKIs, but the side chain can irreversibly bind to Cys797 to inhibit the tyrosine kinase activity of EGFR. For example, the anilinoquinazoline derivative forms hydrogen bonds with the backbone of Met793 in the hinge region and interacts with the hydrophobic region. The acrylamide group binds covalently to Cys797 in the active conformation of EGFR, the furanyl group is exposed to solvent, and the 3-chloro-4-fluorophenyl group is situated next to the gatekeeper residue [1214]. However, mutations such as T790M still emerge upon treatment with second-generation EGFR-TKIs, which have limited selectivity against WT-EGFR, resulting in serious side effects [15]. Fortunately, third-generation covalent inhibitors that bind irreversibly to the target and are mutation-selective have been developed. These compounds were designed based on a new aminopyrimidine scaffold and show preferable biological activities [16]. Early clinical trials have proven that these third-generation EGFR-TKIs are effective in patients with double-mutated tumors (EGFR L858R/T790M or ex19del/T790M) and have high selectivity for mutant EGFR, thereby eliminating the side effects in the skin and gastrointestinal system associated with the nonselective inhibition of WT-EGFR [17]. For example, the crystal structures of rociletinib (CO-1686) in complex with EGFR T790M and EGFR L858R have been published; in EGFR T790M, the anilinopyrimidine group of rociletinib forms hydrogen bonds with the Met793 amide and the carbonyl backbone, whereas in EGFR L858R, hydrophobic interactions between rociletinib and the protein were due to hydrogen bonds between nitrogens in the pyrimidine group and between the fluoromethyl and Thr790. In addition, the acrylamide group in rociletinib covalently binds to Cys797 in the DFG-in/αC-in active conformations [18]. The specificity for EGFR T790M may stem from hydrophobic interactions between the large methionine in mutant EGFR and pyrimidines. Drugs that have been approved for marketing include osimertinib (US), almonertinib (China), lazertinib (South Korea), and alflutinib (China) (Fig. 2).

Mechanisms of resistance to third-generation EGFR-TKIs

Due to the covalent bond between the acrylamide (Michael acceptor) of third-generation EGFR-TKIs and the active thiol in the EGFR kinase domain, highly selective inhibitory activity has been achieved by targeting Cys797 and irreversible binding EGFR; thus, these compounds show excellent antitumor activity. Targeted therapy for patients with EGFR T790M and EGFR-activating mutations showed good efficacy in both first- and second-line settings. In patients who developed resistance to third-generation EGFR-TKIs as first-line therapy, genetic changes such as MET amplification, EGFR C797X mutation, PIK3CA amplification and mutation, HER2 amplification and mutation, K-RAS mutation, and BRAF mutation, as well as changes in cell cycle-related genes and oncogene fusions, have been reported, but no T790M mutations have been detected. The mechanism of resistance to second-line therapy is more complicated. Acquisition or deletion of the T790M mutation has been detected in patients [19], and other EGFR mutations (such as L718Q, L796S, L792H, and exon 20 insertion) have also been observed (Fig. 1B). In addition, the mechanisms of acquired resistance to third-generation EGFR-TKIs include alternative pathway activation and histologic and phenotypic transformation (Fig. 3); the details will be discussed in the following sections.

Primary/intrinsic resistance

The differential sensitivity of TKIs to different EGFR mutations is a cause of primary drug resistance. In NSCLC patients, the in-frame deletion of exon 19 (ex19del) and the L858R point mutation in exon 21 are the most common somatic mutations, occurring in approximately 80% of cases. During EGFR-TKI treatment, patients with longer median survival have presented with more than 20 unique deletions of exon 19. Intrinsic drug resistance can all be triggered by other nonclassical sensitizing mutations (mainly exon 20 insertion) and inherent secondary genetic changes. Drug-resistant clones (for example, T790M) may already exist within the cancer cell population, leading to drug resistance during treatment [20]. Some studies have found that in nearly 1% of lung cancer patients, 2–3 simultaneous driver mutations can be detected before treatment. Some molecular and genetic changes have been reported to relate to intrinsic drug resistance, such as the lack of K-RAS/phosphatase and tensin homolog (PTEN) expression. These preexisting molecular and genetic alterations can stimulate the Ras/Raf/MEK/ERK and PI3K/AKT downstream pathways to promote cancer progression [21].

BIM deletion polymorphism

BIM is a proapoptotic member of the B-cell lymphoma-2 (Bcl-2) family [22]. Recent studies showed that lung cancer cells with the BIM deletion polymorphism and EGFR mutation are resistant to third-generation EGFR-TKIs, suggesting that the BIM deletion polymorphism has potential as a biomarker to predict the efficacy of third-generation EGFR-TKIs in patients [22].

EGFR exon 20 insertion

The molecular mechanism of drug resistance caused by the exon 20 insertion is not fully understood. Eck et al. [23] hypothesized that this mutation prevents binding to EGFR-TKIs due to the addition of residues to the N-lobe of EGFR. The crystal structure of EGFR exon 20 with the D770_N771insNPG insertion shows an unchanged ATP binding pocket and a rigid active conformation, leading to steric hindrance of the drug binding pocket and resistance to EGFR-TKIs.

Acquired resistance

Acquired drug resistance refers to the process by which tumor cells with prior sensitivity to treatment circumvent the inhibitory effects of drugs by changing their metabolic pathways. The mechanisms of acquired resistance to third-generation EGFR-TKIs can be divided into EGFR-dependent resistance and EGFR-independent resistance [24].

EGFR-dependent drug resistance mechanisms

Reappearance of an EGFR mutation

C797S mutation

One point mutation of EGFR (C797S) involves the replacement of Cys797 within the ATP binding site (exon 20) with serine [25]. Osimertinib binds covalently and irreversibly to EGFR T790M by interacting with Cys797. When the C797S mutation occurs, the osimertinib binding efficiency decreases [10], resulting in tumor resistance to all third-generation EGFR-TKIs.

G796R/D mutation

The G796R mutation has been detected in cancer patients who received treatment with a third-generation EGFR-TKI. Molecular docking predictions revealed that G796R sterically hinders the covalent binding of osimertinib. Because the bulky side chain and hydrophilic group hinder the binding of osimertinib to the hydrophobic region, the change in binding energy renders binding unfavorable. Compared with samples containing the double-mutant EGFR L858R/T790M, those harboring the triple-mutant EGFR L858R/T790M/G796R are 110 times more resistant to osimertinib [26]. G796D was reported for the first time in osimertinib-resistant NSCLC patients. In vitro studies have shown that the G796D mutation causes a 50-fold increase in the growth inhibitory 50% (GI50) value of osimertinib. Structural modeling showed that the side chain of the mutated G796D residue collides with the surface of osimertinib, resulting in steric hindrance and energy repulsion and ultimately the loss of binding affinity [27].

L792 mutation

The mutations at Leu792 include L792F, L792Y, and L792H. Structural prediction revealed that these mutations introduce a benzene ring or imidazole ring to the side chain of the residue at 792, which spatially disrupts the orientation of osimertinib, thereby potentially affecting the binding of osimertinib to the EGFR ATP binding site [28].

M766Q mutation

The homology simulation with the T790M and M766Q double mutant showed that M766Q seems to position T790M in the inhibitor binding site, thereby weakening osimertinib binding [29].

Mutations in exon 18

EGFR L718Q/V
EGFR L718Q was reported for the first time in a cell model of resistance to third-generation EGFR-TKIs. Subsequent studies have shown that NSCLC with EGFR L858R/T790M/L718Q is resistant to all EGFR-TKIs, but that with only L858R/L718Q remains sensitive to afatinib [30]. The crystallographic model revealed that the L718Q mutation reduces the efficiency of the formation of covalent bonds between the acrylamide warhead and the Cys797 thiol group, thus interfering with the irreversible binding of osimertinib [31, 32]. In addition, L718V resistance mutations in the kinase domain of EGFR have been detected, and these may interfere with the binding of osimertinib to the kinase domain [33]. Of note, EGFR L718Q/V is still sensitive to afatinib [32].
EGFR G724S
The G724S mutation in the ATP binding loop enriches this loop in glycine, which can lead to the development of resistance to EGFR-TKIs by changing the protein structure, enhancing ATP affinity, and stabilizing activating mutations [34]. However, this mutation does not lead to resistance to second-generation EGFR inhibitors [34].

Compound mutations

A compound mutation refers to the simultaneous detection of two or more different types of EGFR mutations in patient cancer cells [35]. The impact of compound mutations on EGFR-TKI sensitivity is listed in descending order: double classic mutations, compound mutations involving classic mutations and rare mutations, and compound mutations of only rare mutations [36, 37]. These EGFR mutations caused by treatment with third-generation EGFR-TKIs confer resistance to irreversible pyrimidine TKIs but not to quinazoline EGFR inhibitors [38].

T790M reduction or deletion

Deletion of T790M may result from third-generation EGFR-TKI treatment or may be one of the reasons for drug resistance related to tumor heterogeneity. In patients with EGFR T790M, resistance mechanisms are often associated with the C797S mutation or aberrant activation of compensatory pathways, whereas patients with the deletion of T790M typically exhibit different resistance mechanisms, most of which are not associated with EGFR signaling pathways [39].

EGFR amplification

Piotrowska and colleagues reported EGFR T790M allele amplification in rociletinib-resistant clones [40]. Nukaga et al. found that amplification of the WT allele of EGFR is sufficient to mediate resistance to third-generation TKIs. The mechanism of drug resistance may be that EGFR gene amplification leads to a relatively low TKI concentration that is insufficient to exert inhibitory activity [41].

EGFR-independent resistance mechanisms

Not all patients develop resistance to TKIs through EGFR mutation; other pathways of acquiring resistance to third-generation EGFR-TKIs include the activation of alternative or downstream signaling pathways, epithelial interstitial resistance, epithelial–mesenchymal transition (EMT), histologic and phenotypic transformation, oncogene fusion, and cell cycle-related gene abnormalities.

Bypass signal pathway activation

Abnormal activation of MET

There are two main drug resistance mechanisms caused by the abnormal activation of MET: the MET exon 14 skipping mutation (METex14) and MET amplification. METex14 leads to the loss of ubiquitin ligase binding sites, a reduction in receptor ubiquitination, and persistent MET activation, resulting in tumor cell survival and acquired resistance [42]. After treatment with third-generation EGFR-TKIs, MET gene amplification can promote drug resistance by activating MAPK/ERK, which is independent of EGFR [43].

HER2 amplification

Hus et al. found that H1975 cells expressing HER2D16 were resistant to osimertinib in vitro. HER2D16 can form a heterodimer with EGFR or a disulfide homodimer, which activates downstream signaling to achieve resistance to osimertinib [44]. HER2D16-driven drug resistance occurs in a manner unrelated to the kinase Src. In addition, other mutations in exon 20 of HER2 have been reported, including point mutations (such as G776C and L755S) and insertions that cause downstream activation [45, 46]. HER2 mutation occurs in approximately 2–4% of NSCLC cases, mostly in lung adenocarcinoma (LUAD) [47]. In NSCLC, HER2 oncogenic amplification occurs in approximately 3% of cases without EGFR-TKI treatment and accounts for approximately 10% of cases with EGFR-TKI resistance [48].

AXL activation

AXL is an RTK that regulates cell survival, proliferation, metastasis, and other cellular functions. Abnormalities in the AXL gene can generate acquired resistance to TKIs by activating relevant downstream signaling pathways. Osimertinib was found to trigger AXL activation by closing the negative feedback loop with SPRY4, thus triggering inherent osimertinib resistance [49].

Overexpression of HGF

Hepatocyte growth factor (HGF) is the ligand of the proto-oncogene c-Met; it can trigger MET activation through EGFR bypass signaling and induce lung cancer resistance to EGFR-TKIs. Yano et al. [50] found that high expression of HGF was related to the acquired and intrinsic drug resistance to EGFR-TKIs in patients with lung cancer. Tumor specimens from patients with acquired drug resistance showed high expression of HGF in the context of MET amplification and the T790M mutation.

Fibroblast growth factor receptor (FGFR) signaling

FGFR is a transmembrane RTK. Studies have shown that FGFR1 is amplified and fibroblast growth factor 2 (FGF2) mRNA levels are increased in patients with osimertinib resistance, suggesting that the FGFR2-FGFR1 autocrine loop may be related to drug resistance [51]. Patients with the T790M mutation have been reported to show disease progression after treatment with osimertinib and nilotinib. The FGFR3-TACC3 fusion was detected in ctDNA [52, 53]. These findings suggest that abnormalities in the FGFR signaling pathway may underlie the mechanism of acquired resistance to third-generation EGFR-TKIs.

Insulin-like growth factor receptor 1 (IGF1R)

IGF1R, a transmembrane heterotetrameric protein encoded by the gene located on chromosome 15q26.3, is involved in promoting the growth of tumor cells. Abnormal activation of IGF1R leads to EGFR-TKI resistance [54].

Aurora kinases (AURKs)

AURKs are an important category of enzymes within the serine/threonine kinase family consisting of three mammalian isoforms: Aurora kinase A (AURK A), AURK B, and AURK C [55, 56]. AURK A and AURK B are highly expressed in dividing cells and play important roles in mitotic progression. Mammalian AURK A and AURK B share approximately 71% similarity in the carboxy-terminal catalytic domain [57]. Aberrant expression of AURK A and AURK B is involved in a broad range of solid cancers and is associated with adverse prognosis and drug resistance [58, 59]. In addition, Tanaka et al. [60] reported that targeting AURK B can prevent and overcome resistance to EGFR inhibitors in lung cancer by enhancing BIM- and PUMA-mediated apoptosis.

Downstream signaling pathway activation

The activation of signaling pathways downstream of oncogenic receptors can regulate cell proliferation, cell cycle progression, and cell survival. Therefore, the direct regulation of downstream signaling pathway-related factors can lead to acquired resistance.

K-RAS mutation

An epidemiological meta-analysis found that K-RAS mutations are present in NSCLC patients, and all patients with K-RAS mutations were resistant to EGFR-TKIs [61]. K-RAS mutation is related to activation of the RAS-MAPK pathway. The common K-RAS mutations include G12S, G12D, G12A, Q61H, and A146T. Studies have found that inhibiting mutant K-RAS can reduce tumor growth and render NSCLC patients sensitive to EGFR inhibitors [62].

BRAF (v-RAF murine sarcoma viral oncogene homologue B1) mutation

BRAF is a serine/threonine protein kinase that plays a key role in the MAPK/ERK pathway, including in EGFR/RAS/RAF signal transduction. BRAF can regulate cell survival, proliferation, differentiation, and apoptosis, as well as tumor induction. Many BRAF mutations (G469A, V600E, and V599E) have been found in cancer, including lung cancer [63]. Ohashi et al. [64] reported that in patients with lung cancer, BRAF mutations can induce acquired resistance to EGFR-TKIs. Preclinical data showed that the BRAF V600E mutation has a strong association with resistance to the third-generation EGFR-TKI osimertinib in patients with T790M-mutated LUAD.

PI3K/AKT/mTOR

PIK3CA is a driver gene of LUAD. Mutation of PIK3CA can promote tumor cell invasion and increase the activity of downstream PI3Ks. Studies have shown that PIK3CA amplification or mutation (including E453K, E545K, and H1047R) may occur in patients with osimertinib resistance [52, 65]. Increased PI3K activity leads to the activation of various downstream kinases, thereby increasing PI3K/AKT/mTOR pathway activity in the absence of coupling to upstream EGFR phosphorylation.

STAT3 activation

STAT proteins, especially STAT3, are key downstream signal sensors of EGFR activation. In studies on NSCLC, Zhao et al. [66] discovered the clinical significance of JAK2/STAT3 in angiogenesis. Chaib et al. [67] found that osimertinib treatment activates not only STAT3 but also SrcYAP1 signaling, which may act downstream of IL-6 to promote disease progression.

Loss of PTEN

PTEN is a tumor suppressor gene that encodes a protein with lipid phosphatase activity and thus regulates cellular protein phosphatase activity. PTEN has dual antitumor effects and is a key component of many signaling pathways in the body. If mutation or deletion of the PTEN gene or downregulation of PTEN expression can reduce or eliminate its antitumor activity [68], loss of PTEN leads to hyperactivation of the PI3K/AKT signaling pathway and resistance to EGFR-TKIs, including osimertinib.

Hyperactivation of activated Cdc42-associated kinase 1 (ACK1)

Hyperphosphorylation of ACK1 and the subsequent activation of antiapoptotic signaling through the AKT pathway are associated with resistance to third-generation EGFR-TKIs [69].

c-Myc gene

The c-Myc gene is an important member of the MYC gene family. The c-Myc gene can induce cells to proliferate indefinitely and can promote cell division; these activities are related to the occurrence and development of various types of cancer. Studies have shown that c-Myc levels are substantially elevated in different EGFR-mutant NSCLC cell lines with acquired resistance to the third-generation EGFR-TKI osimertinib compared with the corresponding parental cell lines; moreover, these increased levels cannot be reduced by osimertinib. Consistently, c-Myc levels are elevated in the majority of EGFR-mutant NSCLC tissues from patients who relapsed on EGFR-TKI treatment compared with the corresponding baseline c-Myc levels prior to treatment [70]. These findings indicate that c-Myc mediates the therapeutic efficacy of third-generation EGFR-TKIs and the development of acquired resistance to these TKIs.

Other mechanisms

Epithelial–mesenchymal transition (EMT)

In EMT, cancer cells lose their epithelial properties through the loss of E-cadherin, leading to increased vimentin expression and transformation into a mesenchymal phenotype. A previous study found that osimertinib-resistant H1975 cells have EMT characteristics in the absence of other EGFR mutations [71]. EMT is a coordinated process involving multiple regulatory factors, such as EMT-induced transcription factors (EMT-TFs), noncoding RNAs (ncRNAs), and various extracellular signals. EMT-TFs play an important role in all stages of EMT; the most well-known EMT-TFs are members of the SNAIL, ZEB, and TWIST families. Many studies have shown that SLUG and SNAIL overexpression can induce drug resistance [72].

miRNAs and EMT

Long noncoding RNAs (lncRNAs) and microRNAs (miRNAs) play important roles in regulating EMT and TKI resistance. Although most miRNAs have been found to inhibit EMT, some have activity that promotes EMT, including miR-21 and miR-155 [73, 74]. Some miRNAs can promote TKI resistance by activating the PI3K/AKT/mTOR signaling pathway; for example, miR-21 and miR-23a can target PTEN and activate AKT, leading to resistance to EGFR-TKIs [75, 76].

Epigenetic alterations

Epigenetic modifications involved in cancer initiation and progression include changes in DNA methylation patterns and histone modifications. Epigenetic changes are common in the development and progression of lung cancer [77]. Studies have shown that epigenetic disorders can make cancer patients susceptible to acquired resistance to EGFR-TKIs [78].

Oncogene fusion

The AURA-3 and FLAURA trials showed that oncogene fusion might be one mechanism of osimertinib resistance; the identified fusions included transforming growth factor receptor (TGFR)-transforming acidic coiled-coil protein 3 (TACC3), neurotrophic receptor tyrosine kinase 1 (NTRK1)-thrombopoietin mimetic peptide 3 (TMP3), ERC1-RET, SPTBN1-ALK, coiled-coil domain-containing protein 6 (CCDC6)-RET, GOPC-ROS1, AGK-BRAF, NCOA4-RET, ESYT2-BRAF, and echinoderm microtubule-associated protein-like 4 (EML4)-ALK. Oncogene fusions can coexist with the EGFR C797S mutation, MET amplification, and BRAF mutation [79].
Recent studies have shown that changes in cell cycle-related genes, including the CDKN2A E27fs mutation, cyclin D (CCND) amplification, cyclin-dependent kinase 4/6 (CDK4/6) amplification, and cyclin E1 (CCNE1) amplification, can cause resistance to third-generation EGFR-TKIs [65].

Histologic and phenotypic transformation

Histopathological transformation to small cell lung cancer (SCLC) from NSCLC has been reported as a mechanism of acquired resistance to EGFR-TKIs in 3–15% of patients [8083]. Transformed SCLC mainly occurs in Asian patients with adenocarcinoma harboring EGFR-TKI-sensitive mutations (such as the EGFR ex19del/T790M mutation) who are nonsmokers. The widely accepted hypothesis for this transformation posits that adenocarcinoma and SCLC originate from type II alveolar cells. RB1 and TP53 mutations might be involved in SCLC transformation but are not sufficient for the induction of complete transformation. Additional genomic alterations, including those that activate the PI3K/AKT family and downregulate NOTCH signaling and those affecting the MYC and SOX families, AKT pathway activation and other molecules, also participate in the transformation from EGFR-mutant NSCLC. However, the precise mechanisms in other cases are unclear [84]. In addition, squamous cell transformation was recently identified as a mechanism of acquired EGFR-TKI resistance that occurs in approximately 15% of patients who received osimertinib as both first- and second-line therapy. Similar to the case in SCLC transformation, the primary EGFR mutation is preserved in squamous cell transformation [85].

Immune escape

EGFR is expressed in different hematopoietic cell types, including macrophages, monocytes, and certain T-cell subsets. Therefore, it is likely that EGFR inhibitors can interfere with the function of these leukocytes. Immune checkpoint inhibitors (ICIs) have adverse effects and poor efficacy in patients with an EGFR mutation or a secondary T790M mutation, largely because of low tumor mutational burden and a noninflamed tumor microenvironment [8688]. A previous study showed that secreted phosphoprotein 1 (SPP1) promotes macrophage M2 polarization and PD-L1 expression in LUAD, which may influence the response to immunotherapy. SPP1 levels might be a useful marker of immunosuppression in patients with an EGFR mutation and could provide therapeutic insight [89]. In addition, HGF, MET amplification, and EGFR T790M lead to the upregulation of PD-L1 expression in NSCLC and promote immune escape by tumor cells through different mechanisms mediated by the PI3K-Akt, MAPK, and NF-κB pathways [90].

Strategies for overcoming third-generation EGFR-TKI resistance

Fourth-generation EGFR-TKIs: overcoming the L858R/T790M and C797S resistance mutations

Third-generation EGFR-TKIs had the potential for remarkable achievements, if not for the numerous mutations. The C797S mutation, which is a covalent anchor mutation, is located in the ATP binding site of the EGFR tyrosine kinase domain. This missense mutation in exon 20 at position Cys797 blocks the ability of third-generation EGFR-TKIs to form a covalent bond in the ATP binding region, with a consequent decrease in the binding affinity between EGFR and an EGFR-TKI [91]. The combination of the C797S mutation with exon 19 deletion, L858R mutation, or T790M mutation was reported both in vitro and in vivo [91]. Studies have shown that drug-resistant lung cancer cells with two mutations (EGFR-activating mutation/C797S) are sensitive to first- and second-generation EGFR-TKIs. However, lung cancer cells with three mutations (EGFR-activating mutation/T790M/C797S) show resistant to third-generation EGFR-TKIs if the C797S and T790M mutations are both in the trans conformation. Nonetheless, these cells are still sensitive to the combination of first- and third-generation EGFR-TKIs [92]. Of note, if C797S and T790M are mutated in the cis conformation, the cells show resistance to all existing EGFR-TKIs (either alone or in combination) [93]. The resistance to third-generation EGFR-TKIs caused by the trans-C797S mutation can be overcome by drugs targeting different kinase binding sites, including allosteric inhibitors, ATP-competitive inhibitors, and “dual-site” inhibitors that occupy both the ATP binding site and an allosteric site.

Allosteric inhibitors

EGFR has three binding sites: an inactive site, a competitive ATP binding site, and an allosteric site. Ligands and drugs cannot bind the inactive site. Recent studies have mostly focused on either ATP-competitive inhibitors targeting the ATP binding site or molecules that bind the allosteric site, which causes a conformational change in the protein that inhibits the signaling cascade [94]. To overcome the resistance of EGFR-TKIs mediated by the T790M and C797S mutations and to further identify and explore compounds that bind outside the ATP binding domain of EGFR, researchers have pursued the development of allosteric inhibitors, and this appears to be a promising strategy. The newly developed fourth-generation mutant-selective allosteric inhibitors can overcome the T790M and C797S mutations that develop in response to third-generation EGFR-TKIs by binding to sites outside the ATP binding pocket of EGFR.
Through molecular phenotypic screening, Engel et al. obtained quinazoline compound 1a (1), which specifically inhibits the drug-resistant H1975 cell line (L858R/T790M); further modification addressed the problem of off-target activity (nonspecific inhibition). X-ray crystallography verified that compound 1a (1) fits well in the tyrosine kinase domain of c-Src [95] (Fig. 4).
Jia et al. conducted counter-screening of active compounds against WT-EGFR and discovered the first non-ATP-competitive allosteric EGFR L858R/T790M/C797S inhibitor based on the thiazolamide scaffold (EAI001, 2) [96]. The X-ray crystal structure of EAI001 (2) in complex with EGFR T790M shows that EAI001 (2) can bind to the allosteric site of this receptor in the form of a “three-bladed propeller,” partly due to the outward displacement of the C-helix in the inactive conformation of the kinase. The hydrophilic side chain of the WT gatekeeper residue (Thr) cannot adapt to the thiazole of EAI001; therefore, there is no favorable interaction. The thiazole of EAI001 closely interacts with the hydrophobic side chain of Met790; specifically, aminothiazole group of EAI001 directly binds to Met790. The carbonyl oxygen of the isoindoline-1-one moiety is inserted between the mutant gatekeeper residue (Met) and the active site residue Lys745, forming another hydrogen bond with the ε-amine of the Lys745 side chain. The NH group of formamide acts as a hydrogen bond donor for Asp855 in the DFG motif. The cationic phenyl group occupies the hydrophobic pocket formed by Met766, Leu777, and Phe856. The 1-oxindolinyl group is exposed along the C-helix and extends to the solvent-accessible area. The ATP analog adenylyl imidodiphosphate (AMP-PNP) binds the active site cavity in an expected manner. The half maximal inhibitory concentration (IC50) of EAI001 (2) for EGFR L858R/T790M is 24 nmol/L, which is lower than that for WT-EGFR (IC50 > 50 μmol/L). The IC50s of EAI001 (2) for EGFR L858R and EGFR T790M is 0.75 μmol/L and 1.7 μmol/L, respectively. By introducing ortho-hydroxyl and meta-fluorine atoms on the benzene ring of EAI001 (2), the researchers synthesized another compound, EAI045 (3), that binds more tightly than EAI001 (2) to EGFR [96]. However, EAI045 has a major drawback: it must be used in combination with cetuximab to preserve its efficacy. While EAI045 (3) has good selectivity for WT-EGFR, cetuximab is expected to have off-target effects in clinical use. Lee et al. [97] designed the EGFR allosteric inhibitor TREA-0236 (4) based on the structure–activity relationships of EAI045 (3). The structure of EAI045 (3) was modified by cyclization, wherein the 2-aminothiazole amide was converted to quinazoline-4-one. To minimize hematological and methemoglobinemia toxicity and to obtain better safety and pharmacokinetic parameters, To et al. linked the 5-indole substituent to the isoindolinone of EAI001 (2) and obtained a new EGFR allosteric compound, JBJ-02-112-05 (5), with an IC50 of 15 nmol/L for EGFR L858R/T790M [98]. Additionally, EAI045 (3) was further optimized to generate another EGFR allosteric inhibitor, JBJ-04-125-02 (6), in which the 2-hydroxy-5-fluorophenyl of EAI045 (3) was combined with the phenylpiperazine on isoindolinone. This compound showed a significantly increased ability to inhibit EGFR L858R/T790M, with an IC50 of 0.26 nmol/L. Interestingly, combination with osimertinib enhanced the efficacy of JBJ-04-125-02 (6) and improved the targeting of JBJ-04-125-02 (6) to cancer cells [98], indicating that the combined use of covalent mutant-selective ATP-competitive inhibitors and EGFR allosteric inhibitors may be an effective treatment strategy for patients with EGFR-mutant disease (Fig. 5). Encouraged by the advantages of inhibiting allosteric sites in the EGFR tyrosine kinase domain, researchers have extensively designed and optimized allosteric inhibitors for EGFR [98101], as shown in Table 1.
Table 1
EGFR allosteric inhibitors
Compound (reference)
Structure
Activity
Interaction with EGFR allosteric site
7, DDC4002 [100]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figa_HTML.png
EGFRL858R/T790M
IC50 = 10 nmol/L
EGFRL858R/T790M/C797S
IC50 = 59 nmol/L
Phe856
8 [100]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figb_HTML.png
EGFRL858R/T790M/C797S
IC50 = 19 nmol/L
Phe856
9 [100]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figc_HTML.png
EGFRL858R/T790M/C797S
IC50 = 23 nmol/L
Phe856
10 [100]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figd_HTML.png
EGFRL858R/T790M/C797S
IC50 = 13 nmol/L
Phe856
11 [100, 101]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Fige_HTML.png
EGFRL858R/T790M
IC50 = 1 nmol/L
EGFRL858R/T790M/C797S
IC50 = 5 nmol/L
Phe856
12 [100, 101]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figf_HTML.png
EGFRL858R/T790M
IC50 = 3 nmol/L
EGFRL858R/T790M/C797S
IC50 = 4 nmol/L
Phe856
The novelty of allosteric sites has attracted the attention of researchers, and these sites have become the most promising targets for the development of drugs for NSCLC and other diseases. Fourth-generation EGFR-TKIs require further investigation and development so that they are suitable as single-agent drugs targeting EGFR ex19del/T790M/C797S [98]. Allosteric inhibitors have now entered the stage of rapid development and are expected to enter clinical trials soon, with the goal of benefitting more patients.

ATP-competitive inhibitors

ATP-competitive inhibitors form one to three hydrogen bonds with amino acids in the hinge region of the target kinase, thereby mimicking the characteristic hydrogen bonds formed by the adenine ring of ATP. This type of inhibitor usually consists of a heterocyclic ring system that occupies the purine binding site, where it acts as a side chain scaffold that occupies the adjacent hydrophobic regions I and II. A high physiological or intracellular concentration of ATP may block the phosphotransferase activity of the target. The size of the amino acid side chain at the gatekeeper residue determines the relative accessibility of the hydrophobic pocket near the ATP binding site. To overcome drug resistance related to triple-mutant EGFR, it is particularly crucial to develop new ATP-competitive inhibitors based on structural design and optimization. Many ATP-competitive inhibitors have been reported; below, we summarize recent ATP-competitive inhibitors that can overcome the resistance to third-generation EGFR inhibitors (Table 2) [102119].
Table 2
ATP-competitive EGFR inhibitors
Compound (reference)
Structure
Enzymatic activity
Biological activity
DMPK profile
13, JND3229 [102]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figg_HTML.png
EGFRL858R/T790M/C797S
IC50 = 5.8 nmol/L
BaF3-EGFRL858R/T790M/C797S
IC50 = 510 nmol/L
BaF3-EGFRDel19/T790M/C797S
IC50 = 320 nmol/L
N/A*
14 [103]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figh_HTML.png
EGFRWT
IC50 = 16 nmol/L
EGFRL858R/T790M/C797S
IC50 = 88 nmol/L
A431
GI50 = 3600 nmol/L
H1975
GI50 = 140 nmol/L
N/A
15 [104]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figi_HTML.png
EGFRL858R/T790M/C797S IC50 = 8 nmol/L
A431
EC50 = 4000 nmol/L
H1975
EC50 = 400 nmol/L
N/A
16 [105, 106]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figj_HTML.png
EGFRL858R/T790M/C797S
IC50 = 630 nmol/L
H1975
IC50 = 1200 nmol/L
Liver microsomes (Human):
t1/2 (min) = 66.6, CLint (mL/min/kg) = 20.8
17 [107]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figk_HTML.png
EGFRWT
IC50 > 1000 nmol/L
EGFRL858R/T790M/C797S
IC50 = 27.5 nmol/L
BaF3-EGFRL858R/T790M/C797S
IC50 = 662 nmol/L
T1/2 (rat, minutes) = 8.36,
CLint (mL/min/kg) = 297.12
18 [108]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figl_HTML.png
EGFRL858R/T790M/C797S
IC50 = 7.2 nmol/L
HCC827
IC50 = 44 nmol/L
H1975
IC50 = 400 nmol/L
N/A
19 [109, 110]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figm_HTML.png
EGFRL858R/T790M/C797S
IC50 = 18 nmol/L
HCC827
IC50 = 0.88 nmol/L
H1975
IC50 = 200 nmol/L
A549
IC50 = 2910 nmol/L
A431
IC50 > 10,000 nmol/L
N/A
20 [111]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Fign_HTML.png
EGFRL858R/T790M/C797S
IC50 = 8.5 nmol/L
EGFR CTG
EC50:
HCC827
IC50 < 14 nmol/L
H1975
IC50 = 51 ± 19 nmol/L
A431
IC50 = 1675 ± 402 nmol/L
A549
IC50 = 3700 nmol/L
H358
IC50 = 3700 nmol/L
In vivo PK (mice, IP, 20 mg/kg):
AUCfree (h∙ng/mL) = 8.6
t1/2 (h) = 1.2
Cmax,free (μmol/L) = 0.012
21 [112]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figo_HTML.png
EGFRL858R/T790M/C797S
IC50 = 3.1 nmol/L
H1975
IC50 = 0.12 ± 0.09 μmol/L
BaF3-EGFRL858R/T790M/C797S
IC50 = 0.29 μmol/L
BaF3-EGFR19D/T790M/C797S IC50 = 0.31 μmol/L
N/A
22 [113]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figp_HTML.png
EGFRWT
IC50 > 1000 nmol/L
EGFRL858R/T790M/C797S
IC50 = 218.3 nmol/L
EGFRDel19/T790M/C797S
IC50 = 15.3 nmol/L
H1975
IC50 = 16,180 nmol/L
A431
IC50 = 20,480 nmol/L
BaF3-EGFRDel19/T790M/C797S
IC50 = 8510 nmol/L
N/A
23 [114]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figq_HTML.png
EGFRWT
IC50 = 430 nmol/L
EGFRDel19/T790M/C797S
IC50 = 0.2 nmol/L
BaF3-EGFRWT
IC50 = 1000 nmol/L
BaF3-EGFRDel19
IC50 = 180 nmol/L
BaF3-EGFRDel19/T790M
IC50 = 99 nmol/L
BaF3-EGFRDel19/T790M/C797S
IC50 = 63 nmol/L
N/A
24 [115]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figr_HTML.png
N/A
Biochemical potency:
BaF3 cells IC50 < 100 nmol/L
Antiproliferative activity:
BaF3 cells IC50 < 100 nmol/L
N/A
25 [116]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figs_HTML.png
N/A
Antiproliferative activity:
PC-9-EGFRL858R/T790M/C797S IC50 = 595.7 nmol/L
PC-9-EGFRDel19/T790M/C797S IC50 = 739.9 nmol/L
A549 IC50 = 2861.7 nmol/L
BaF3-EGFRWT IC50 = 519.82 nmol/L
BaF3-EGFRL858R/T790M/C797S IC50 = 0.16 nmol/L
BaF3-EGFRDel19/T790M/C797S IC50 = 0.23 nmol/L
N/A
26 [117]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figt_HTML.png
N/A
BaF3-EGFRWT IC50 = 540 nmol/L
BaF3-EGFRL858R/T790M/C797S IC50 = 48.2 nmol/L
BaF3-EGFRDel19/T790M/C797S IC50 = 12 nmol/L
N/A
27 [115]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figu_HTML.png
EGFRWT
IC50 = 7.92 nmol/L
EGFRL858R/T790M/C797S
IC50 = 0.218 nmol/L
EGFRDel19/T790M/C797S
IC50 = 0.16 nmol/L
Antiproliferative Activity:
A431
IC50 = 154 nmol/L
BaF3-EGFRDel19/T790M/C797S
IC50 = 22 nmol/L
In vivo PK (mice, per os, 15 mg/kg):
AUC0-last = 57,037 (nmol/L∙h)
t1/2 (h) = 10.0
Plasma (nmol/L), 2 h = 3553
Tumor (nmol/kg), 2 h = 16,667
28 [118]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figv_HTML.png
EGFRWT
IC50 = 3.8 nmol/L
EGFRL858R/T790M/C797S
IC50 = 38.1 nmol/L
BaF3-EGFRL858R/T790M/C797S
IC50 < 1000 nmol/L
N/A
29, UPR1444 [119]
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figw_HTML.png
EGFRWT
IC50 = 30 ± 4.8 nmol/L
EGFRL858R/T790M/C797S
IC50 = 110 ± 33 nmol/L
N/A
N/A
*N/A not available

“Dual-site” inhibitors: occupying both the ATP binding site and the allosteric site

Based on the non-ATP-competitive EGFR L858R/T790M/C797S inhibitor EAI001 reported by Jia et al., the more potent compound EAI045 (3) was obtained through structural optimization [96]. EAI045 (3) binds to the allosteric site created by the outward displacement of the αC helix of EGFR, located next to the ATP binding pocket. Facilitated by molecular docking, researchers developed a series of new compounds that noncovalently occupy both the EGFR ATP binding site and the allosteric site; these fourth-generation reversible EGFR inhibitors have improved binding affinity for EGFR L858R/T790M/C797S, effectively compete with ATP, and further overcome resistance to third-generation EGFR inhibitors.
The compound vandetanib (30) [120] is a known EGFR inhibitor that shows moderate efficacy against EGFR L858R/T790M/C797S, with an IC50 value of 369.2 nmol/L. Via molecular docking simulation, Li et al. found that vandetanib can extend to the EGFR ATP binding pocket (gscore =  − 8.2 kcal/mol). The docking model of vandetanib with EGFR T790M/V948R shows that the phenyl group of vandetanib binds the ATP binding site of EGFR, occupying a position such that it resembles the thiazole moiety of EAI001 (2). EAI001 (2) binds as a Y-shaped constellation in the allosteric site [121]. Modifying vandetanib to occupy both the ATP binding site and the allosteric site may be an effective way to improve its biological activity against EGFR L858R/T790M/C797S. To promote occupation of the allosteric site of EGFR, the structure of EAI045 (3) was modified such that the hydrophobic group oxyisoindole-2 phenylacetamide was introduced with an amide bond as the linker, generating compound 31. With this compound as a new lead, three moieties, namely the allosteric targeting region, the hinge targeting region, and the solvent exposure region, were studied and optimized. Finally, the EGFR L858R/T790M/C797S reversible inhibitor compound 32 (Fig. 6) was obtained, with an IC50 value of 2.2 nmol/L. The docking simulation showed that compound 32 occupies both the ATP binding region and the allosteric region. In addition, it extensively interacts with residues in the allosteric region, the solvent-accessible region, and the hinge region. The phenyl of the Y-shaped group (oxoisoindolin-2-phenylacetamide) and Phe856 of the allosteric cavity form ππ stacking interactions. Inside the ATP binding region, hydrogen bonds are formed between the quinazoline ring and the hinge residue Met793. In addition, the piperidine tail is surrounded by the solvent-exposed region. At a concentration of 0.1 μmol/L, compound 32 almost completely inhibited the phosphorylation of EGFR, showing comparable potency to that of EAI045 (3).
To further design more potent inhibitors spanning both binding sites, considering the proximity of the ortho and allosteric positions, Wittlinger F et al. compared the binding of the EGFR ATP site inhibitor LN2057 (33) with the allosteric inhibitor EAI045 (3) and found that the 4-fluorophenyl of LN2057 (33) and the thiazole of EAI045 (3) had the same binding position [122]. Based on this, researchers designed and synthesized a series of compounds that combined a large portion of the isomerization inhibitor EAI045 (3) with the pyridyl-imidazole skeleton. For compound 34, the pyridinylimidazole scaffold partly binds the 2-fluoro-5-hydroxyphenyl moiety of EAI045 (3); 1-oxoisoindoline-2-yl was introduced into compound 35; and 1,3-dioxoisoindoline-2-yl was added to generate compound 36 to further explore the structure–activity relationship of the allosteric site. In addition, an N-(4-methoxyphenyl)acrylamide warhead was introduced to produce compound 37, and the influence of the C797-targeting capacity of these chimeric compounds, which are expected to form a covalent bond with C797, was assessed. The X-ray cocrystal analysis of the binding mode with EGFR T790M/V948R (Fig. 7) revealed that compounds 34 and 36 bind in the same way. Taking compound 36 as an example, the aminopyridine moiety forms a hydrogen bond with the M793 residue in the hinge region. The inhibitor is anchored at the ATP binding site, and the N atom of the imidazole moiety forms a hydrogen bond with K745, which is essential for the strong reversible binding of the imidazole skeleton. The phenylamide bond extending into the allosteric pocket is directed toward the T790M mutation, and the N atom on the amide forms hydrogen bonds with the T854 and D855 residues. Despite considerable efforts, the X-ray crystal structure of compound 37 in complex with EGFR was not obtained. Compound 37 was computationally docked to the EGFR T790M/V948R kinase domain, and the result was the same as that for compound 36. The methoxyphenyl acrylamide formed a covalent bond with C797. Importantly, no covalent binding of compound 37 to the EGFR L858R/T790M/C797S kinase domain was observed, confirming that this compound is a noncovalent inhibitor.
The inhibitory activity of the above compounds was tested, and the results showed that compound 34 exhibited strong inhibitory activity against all mutants, with IC50 values of 5–32 nmol/L, indicating that the introduction of 2-fluoro-5-hydroxyphenyl alone did not increase selectivity. With the introduction of oxyisoindolin-2-yl, the inhibitory activity of compound 35 decreased, but a certain inhibitory effect against EGFR L858R was observed. Furthermore, compound 1 inhibited all three EGFR mutants at the low nanomolar range. Compound 37 showed a moderate degree of mutation selectivity for WT-EGFR, possibly due to the methoxyphenyl acrylamide group. To assess the kinase selectivity of compound 37, a kinome screen including 335 WT kinases was performed; compound 37 exhibited high selectivity, with a selectivity score of 0.006 at an inhibitor concentration of 1 μmol/L. Next, the antiproliferative activity of these compounds was evaluated in Ba/F3 cells stably transfected with WT-EGFR, EGFR L858R, EGFR L858R/T790M, or EGFR L858R/T790M/C797S. Among the compounds, compound 37 showed an antiproliferative effect in the EGFR L858R and EGFR L858R/T790M cell lines, with IC50 values in the micromolar range in the presence and absence of cetuximab. The IC50 value of compound 37 in EGFR L858R Ba/F3 cells (1.2 ± 0.07 μmol/L) was comparable to that of EAI045 (3) combined with cetuximab (840 ± 700 nmol/L). Although compound 37 is potent and selective for kinases, its cellular activity is suboptimal. Kinase selectivity was achieved by increasing the molecular weight of the lead compound and increasing the number of hydrogen bond donors and acceptors, but these changes may have produced limited cell permeability and effects on cell viability; thus, this compound lacked sufficient activity in cells expressing EGFR L858R/T790M/C797S.
The selective EGFR inhibitor (compound 37) designed and developed in this study can bind to both the ATP site and the allosteric site of the EGFR kinase domain. Adding allosteric inhibitor elements to the compound skeleton at the ATP binding site contributes to the mutation selectivity of these compounds. The designed compound 37 has good kinase activity but nonideal cell activity. Future research and development could optimize the structure of this lead compound to further enhance its cellular activity.

PROTAC technology

Allosteric EGFR degrader

Resistance to third-generation EGFR-TKIs is a major obstacle to clinical targeted therapy. Due to changes in the EGFR protein [123], some kinase inhibitors are restricted to the catalytic pocket [124]. A proteolysis-targeting chimera (PROTAC) induces the proteasomal degradation of the target by recruiting it to a specific E3 ligase. The eradication of EGFR protein from cancer cells provides a promising strategy for overcoming drug resistance. The allosteric EGFR degrader is a heterobifunctional compound based on allosteric EGFR inhibitors. It includes a small molecule (protein-of-interest (POI) ligand) that binds the target protein and a small-molecule E3 ligase ligand that recruits cereblon (CRBN), von Hippel–Lindau (VHL), cellular inhibitor of apoptosis protein 1 (cIAP1) or murine double minute 2 (MDM2). After the addition of a linker connecting the two parts [125, 126], these chimeras can degrade mutant EGFR without affecting WT-EGFR.
Compared with classic “occupying” inhibitors, allosteric EGFR degraders can completely eliminate the function of the target protein, thereby improving the phenotypic potency. Moreover, since PROTAC molecules usually do not require strong binding to targets or long-term retention to achieve protein degradation, the development of drug-induced resistance mutations may be prevented. Compared with kinase inhibitors, PROTACs have the advantages of activity at lower concentrations, limited dose-dependent toxicity, and the potential to overcome drug resistance and target drug refractory disease [127132]. These molecules have attracted considerable attention from academia and industry and have become an attractive therapeutic strategy in drug discovery.
Based on EAI001 (2), a compound that buries deeply in the allosteric pocket [96], Jang et al. introduced 1-(pyridin-2-yl)piperazine at the 6 position of isoindolinone and synthesized JBJ-07-149 (38), which has an IC50 value of 1.1 nmol/L for EGFR L858R/T790M. In combination with cetuximab, JBJ-07-149 has a half maximal effective concentration (EC50) of 0.148 nmol/L for EGFR L858R/T790M. However, this compound was less potent in the proliferation assay (EC50 = 4.9 nmol/L) [133].
Based on JBJ-07-149 (38), different linkers that bind the piperazine group and connect the CRBN ligand were evaluated. The compound with 3-PEG as the linker (DDC-01-163, 39) showed the strongest antiproliferative activity for EGFR L858R/T790M (Fig. 8). DDC-01-163 (39) induced the selective degradation mutant EGFR and inhibited the proliferation of cells expressing mutant EGFR in a dose- and time-dependent manner. DDC-01-163 (39) showed no activity in WT-EGFR Ba/F3 cells (EC50 > 10 μmol/L) but inhibited the proliferation of EGFR L858R/T790M Ba/F3 cells, including those expressing EGFR L858R/T790M (EC50 = 0.096 μmol/L), EGFR L858R/T790M/C797S (EC50 = 0.041 μmol/L) and EGFR L858R/T790M/L718Q (EC50 = 0.028 μmol/L). The results in H1975 cells were consistent with those in Ba/F3 cells. Osimertinib-resistant cell lines treated with 0.1 μmol/L DDC-01-163 (39) showed EGFR L858R/T790M/C797S and EGFR L858R/T790M/L718Q degradation rates of 74% and 71%, respectively.
Jang et al. also identified the 2-hydroxy-5-fluorophenyl allosteric inhibitor JBJ-04-125-02 (6), which can be used as a single agent to inhibit the proliferation of Ba/F3 cells. Following the same strategy as that used to develop DDC-01-163 (39), this group designed JBJ-04-125-02 (6) as a PROTAC molecule and synthesized the allosteric EGFR degrader JBJ-07–038 (40) (EC50 = 0.48 μmol/L). In addition, JBJ-07-200 (41) (EC50 = 0.15 μmol/L) was obtained by replacing the hydroxyl group of JBJ-04-125-02 (6) with fluorine (Fig. 9), which could potentially improve membrane permeability [133]. It is highly anticipated that the further characteristic optimization and development of allosteric EGFR PROTACs will produce a valuable therapeutic strategy that will benefit more patients with EGFR-mutant disease.
According to the first report by Zhao et al., EGFR degradation induced by PROTACs may be related to the autophagy pathway [134]. Qu et al. [135] demonstrated for the first time that in addition to the well-known ubiquitin/proteasome pathway, the ubiquitin/autophagy/lysosomal pathway participates in PROTAC-induced EGFR degradation. Based on the EGFR inhibitor canertinib (41) and the CRBN ligand pomalidomide (an E3 ubiquitin ligase ligand), researchers generated two novel EGFR PROTACs (Fig. 10), namely SIAIS125 (42) and SIAIS126 (43). These two EGFR degraders showed effective and selective antitumor activity in EGFR-TKI-resistant lung cancer cells.

Dual PROTACs

The basic goal of modern drug discovery is to develop efficient and selective drugs for specific targets. However, complex diseases such as cancer usually result from interactions among multiple factors, synergistic effects of multiple disease-modifying factors, the upregulation of multiple receptors, and crosstalk between signaling networks. Tumor cells readily gain drug resistance by upregulating an alternative factor or transforming the signaling pathway that promotes proliferation; therefore, treatment focused on only a single target has limitations. In addition to its issues related to drug resistance, single-target drugs also show reduced efficacy and can decrease the quality of life of patients due to side effects and tissue toxicity.
To overcome the deficiencies of single-target drugs, single hybrid molecules fused to two or more pharmacophores have been designed to simultaneously target two or more antitumor epitopes or targets. These hybrid molecules can simultaneously modulate multiple targets or pathways and thus generally have better efficacy with fewer side effects. Based on this information and inspired by the great success of dual-targeted drugs, especially dual-specific antibodies, Professor Li et al. combined the concepts of PROTACs and dual targeting; this group used trifunctional natural amino acids as starlike core linkers to connect two independent inhibitors, gefitinib (44) and olaparib (45), that are linked to CRBN or VHL E3 ligands. The synthesized novel dual PROTACs can successfully and simultaneously degrade EGFR and poly(ADP-ribose) polymerase (PARP) in cancer cells [136]. Among the developed compounds, compound DP-V-4 (46) exhibited the best ability to degrade EGFR and PARP in a dose- and time-dependent manner in H1299 cells and human epidermal carcinoma A431 cells (Fig. 11). As the first successful example of a dual PROTAC, this research will inject new vitality into the field of combination therapy for cancer. Moreover, these findings will broaden the potential applications of the PROTAC method, open new fields of drug discovery, and overcome the limitations of single-target therapy against EGFR.
Another new technology is the autophagy-targeting chimera (AUTAC), a small molecule that targets protein degradation through autophagy and contains both a degradation tag (guanine derivative) and a warhead to provide target specificity; AUTACs have a wider substrate panel than the ubiquitin–proteasome system [137139]. Therefore, there is considerable potential for the design and development of AUTAC molecules to degrade EGFR.

Monoclonal antibodies and ADCs

For patients with EGFR-mutant disease, there are targeted therapies for tumors harboring EGFR-TKI-sensitizing mutations [140]. The EGFR monoclonal antibody can bind to the extracellular domain of EGFR to compete with EGF binding, thereby blocking downstream signaling. The variable fragment (Fv) is composed of parts of the light chain and heavy chain of the antibody and has unique antigen recognition function. The constant region (Fc) mediates innate immunity related to monoclonal antibodies, mainly by binding immune factors or cells to exert antitumor effects. These properties make antibodies a favorable approach in targeted therapy, especially in combination with other strategies. In addition, the internalization and degradation of EGFR monoclonal antibody and receptor complexes can downregulate EGFR on the surface of cancer cells. EGFR monoclonal antibodies are now standard-of-care therapies for head and neck cancer and colorectal cancer. Common EGFR monoclonal antibodies include cetuximab, necitumumab, panitumumab, matuzumab, and nimotuzumab. Antibody–drug conjugates (ADCs) are composed of three moieties: the antibody, linker, and drug (especially those with potential cytotoxicity) (Fig. 12). Antibodies are equivalent to precise arrows, and highly active cytotoxic drugs (the payload) correspond to the gunpowder on the arrows; these drugs mainly include tubulin inhibitors (monomethyl auristatin E, monomethyl auristatin F, mertansine, and ravtansine) and DNA-damaging agents (those that cause DNA double-strand breaks, DNA alkylation, DNA intercalation, and DNA cross-linking). It is difficult to effectively kill tumor cells with only cytotoxic drugs, but monoclonal antibodies alone are too inefficient. ADCs composed of both the cytotoxic drug and a monoclonal antibody represent a more powerful combination. ADCs can precisely target tumor cells by combining highly specific monoclonal antibodies with highly toxic cytotoxic drugs, thereby achieving a precise attack on EGFR-TKI-resistant cancer cells and filling the gap between antibody drugs and traditional chemotherapy drugs. The ADC approach can improve both the drug specificity and the treatment window. Being precise and efficient, ADCs have therapeutic potential across cancer types and can also induce tumor cell death via the bystander effect [141].
He et al. developed a new ADC targeting EGFR, namely SHR-A1307 (47) (Fig. 13), for the treatment of solid tumors resistant or refractory to EGFR-targeted therapy [142]. SHR-A1307 (47) has intermediate ability to block EGFR affinity for hR3 and selectively binds to cancer cells expressing EGFR while avoiding inhibitory effects on normal cells. In addition to increasing stability and reducing systemic toxicity, Fc domain engineering improved the pharmacokinetics. Although less frequent drug administration may reduce toxin accumulation, effective tumor cell killing with minimal toxicity were observed. In addition, SHR-A1307 (47) can effectively kill cancer cells that do not respond to current EGFR inhibitors and shows low nanomolar in vitro cytotoxicity in a broad spectrum of cancer cells with different drug resistance mutations, thus providing an attractive treatment opportunity to overcome the drug resistance of patients with EGFR-overexpressing tumors.
MRG003 (48) [143, 144], the first EGFR ADC to enter the clinical trials in China, is composed of a humanized anti-EGFR monoclonal antibody and the tubulin inhibitor MMAE coupled through a degradable VC (Val-Cit) linker (Fig. 13). The phase I dose escalation and expansion study for patients with refractory solid tumors has been completed. Based on the results of the phase Ia and Ib clinical trials, Lepu Biosciences is currently conducting phase II clinical trials of MRG003 monotherapy in China for recurrent or metastatic advanced head and neck squamous cell carcinoma, advanced NSCLC, biliary tract cancer, and nasopharyngeal carcinoma.

Combination therapy strategy

Resistance to third-generation EGFR inhibitors mediated by EGFR-independent mechanisms can develop through the activation of alternative bypass pathways and abnormal downstream signal transduction closely related to tumor growth, invasion and metastasis. In the clinic, HER2 mutation, high HGF expression, and abnormal activation of MET, AXL, IGF1R and the FGFR pathway were found in patients with acquired resistance to third-generation EGFR-TKIs. Mutation or abnormal expression of EGFR signaling pathway-related genes involved in the Ras/Raf/MEK/ERK/MARK, PI3K/PDK1/Akt, PLC-γ and JAK/STAT pathways was also found. Importantly, these aberrations can coexist in the same tumor and with EGFR-TKI tertiary mutations, which are the basis for the complexity and heterogeneity of cancer evolution in response to EGFR-TKI treatment. Therefore, in combination with third-generation EGFR-TKIs, targeting important components of alternative bypass pathways (Table 3) [145154] and downstream signal transduction pathways (Table 4) [155164] appears to be a promising treatment strategy.
Table 3
Combination therapy with the bypass pathway target
Target
Representative compound*
Structure
Reference
MET
Cabozantinib
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figx_HTML.png
[145, 146]
MET
Crizotinib
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figy_HTML.png
[147]
MET
Savolitinib
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figz_HTML.png
[148]
FGFR
AZD4547
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figaa_HTML.png
[149]
ALK
Lorlatinib
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figab_HTML.png
[150]
ALK
Brigatinib
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figac_HTML.png
[151]
HER2
JQ1
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figad_HTML.png
[152]
HER2
Trastuzumab-DM1
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figae_HTML.png
[153]
BRAF V600E
Encorafenib (LGX818)
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figaf_HTML.png
[64]
AURK B
PF-03814735
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figag_HTML.png
[60, 154]
*Osimertinib is a representative third-generation EGFR-TKI
Table 4
Combination therapy with targets in downstream signaling pathways
Target
Representative compound*
Structure
Reference
MEK
Trametinib
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figah_HTML.png
[155, 156]
MEK
Selumetinib
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figai_HTML.png
[155157]
MEK
PD0325901
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figaj_HTML.png
[155]
AKT
Uprosertib (GSK2141795)
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figak_HTML.png
[158]
AKT
Capivasertib (AZD5363)
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figal_HTML.png
[158]
AXL
Cabozantinib
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figam_HTML.png
[159]
AXL
DS-1205b
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figan_HTML.png
[160]
AXL
Yuanhuadine (YD)
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figao_HTML.png
[161]
AXL
Bemcentinib (R428)
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figap_HTML.png
[162]
ACK1
(R)-9b
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figaq_HTML.png
[163, 164]
*Osimertinib is a representative third-generation EGFR-TKI

Multitarget inhibitors

Cancer is a multifactorial disease, and single-target treatments may have poor efficacy. As clinical targeted therapy, EGFR kinase inhibitors are effective only when the cancer cells contain specific EGFR-activating mutations that alter downstream signaling [165]. Moreover, only a small proportion of patients benefit from EGFR inhibitors [2]. In addition to activating mutations at the EGFR locus that lead to drug resistance, a large number of genetic and epigenetic abnormalities may also lead to resistance to third-generation EGFR-TKIs. The emergence of intrinsic and acquired resistance requires appropriate strategies to prevent serious side effects. Combination therapy has additive or even synergistic effects, but due to various dose-limiting toxicities and drug–drug interactions caused by changes in pharmacokinetics, the simultaneous use of two or more drugs in the clinic is challenging. Therefore, as an alternative to combination therapy, drugs targeting two or more objects have a lower risk of drug–drug interactions and better pharmacokinetic and safety profiles, which helps mitigate poor patient compliance, off-target effects, and high development costs. Such treatment regimens are more flexible and can represent an effective strategy for cancer therapy [166, 167]. The effectiveness of multitarget kinase inhibitors of WT and/or mutant EGFR has been extensively studied (Table 5) [59, 168197]. Some EGFR-mutant cell lines are sensitive to multitarget inhibition and maintain certain levels of activity, highlighting the selectivity of multitarget compounds and suggesting that multitarget inhibition can be used to circumvent acquired multidrug resistance to EGFR-targeted therapy without serious side effects.
Table 5
Multitarget inhibitors
Number
Target
Pharmacophores
Structure
Activity
69 [168]
EGFR/FGFR1
Pyrimidine-2,4-diamines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figar_HTML.png
EGFRL858R/T790M
IC50 = 43.1 nmol/L
EGFRWT
IC50 = 1138.7 nmol/L
FGFR1WT
IC50 = 17.6 nmol/L
H1975 cells
IC50 = 336.3 nmol/L
70 [169]
EGFR/Src
Pyrimidine-4-amines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figas_HTML.png
K562 cells
IC50 = 220 nmol/L
A549 cells
IC50 = 250 nmol/L
EGFR inhibition rate = 33.15% (10 µmol/L)
Src inhibition rate = 72.12% (1 µmol/L)
71 [170]
EGFR/HER4
3-Cyanoquizolines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figat_HTML.png
EGFRL858R
IC50 = 419 nmol/L
EGFRWT
IC50 = 2.4 nmol/L
HER4
IC50 = 0.03 nmol/L
72 [171]
EGFR/COX2
1,3,4-Oxadiazole scaffold
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figau_HTML.png
EGFR
IC50 = 280 nmol/L
COX2
IC50 = 170 nmol/L
UO-31 cells
IC50 = 5800 nmol/L
73 [172]
EGFR/BRAF
Spirobenzo[h]chromene derivatives
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figav_HTML.png
EGFR
IC50 = 1200 nmol/L
BRAF
IC50 = 2600 nmol/L
A549 cells
IC50 = 1780 nmol/L
MCF-7 cells
IC50 = 4090 nmol/L
HT-29 cells
IC50 = 4450 nmol/L
74 [173]
EGFRT790M/ALK
Pyrimidine-2,4-diamines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figaw_HTML.png
ALK
IC50 = 18 nmol/L
EGFRWT
IC50 = 151 nmol/L
EGFRT790M
IC50 = 2 nmol/L
EGFRL858R/T790M
IC50 = 4 nmol/L
DFCI032 cells
IC50 = 170 nmol/L
DFCI076 cells
IC50 = 820 nmol/L
75 [174]
EGFRWT and mutant EGFR/ALK
Pyrimidine-2,4-diamines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figax_HTML.png
EGFRWT
IC50 = 108 nmol/L
EGFRT790M
IC50 = 3.9 nmol/L
EGFRL858R/T790M
IC50 = 3.6 nmol/L
ALKWT
IC50 = 9.8 nmol/L
ALKR1275Q
IC50 = 0.82 nmol/L
ALKL1196M
IC50 = 0.59 nmol/L
ALKF1174L
IC50 = 0.92 nmol/L
ALKC1156Y
IC50 = 1.0 nmol/L
H1975 cells
GI50 = 15 nmol/L
H3112 cells
GI50 < 0.3 nmol/L
76 [175]
EGFR/ATX
Pyrimidine-4-amines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figay_HTML.png
EGFR
IC50 = 24.2 nmol/L
ATX
IC50 = 29.1 nmol/L
A549 cells
IC50 = 4960 nmol/L
MKN-45 cells
IC50 = 3430 nmol/L
SGC cells
IC50 = 2910 nmol/L
CFs cells
IC50 = 1490 nmol/L
77 [176]
EGFR/AURK A
Pyrimidine-4-amines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figaz_HTML.png
AURK A
IC50 = 1990 nmol/L
EGFR
IC50 = 3.76 nmol/L
78 [177]
EGFR/IGF1R
Pyrimidine-2-amines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figba_HTML.png
EGFR
IC50 = 35.5 nmol/L
EGFRT790M
IC50 = 66.0 nmol/L
IGF1R
IC50 = 52.0 nmol/L
79 [178]
EGFR/tubulin
Pyrimidine-4-amines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbb_HTML.png
EGFR
IC50 = 30 nmol/L
Tubulin assembly
IC50 = 710 nmol/L
HeLa cells
IC50 = 1 nmol/L
HT-29 cells
IC50 = 20 nmol/L
Jurkat cells
IC50 = 1 nmol/L
RS4;11 cells
IC50 = 1 nmol/L
80 [179]
EGFR/tubulin
Chalcones
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbc_HTML.png
EGFR
IC50 = 39 nmol/L
Tubulin polymerization
IC50 = 8840 nmol/L
MCF-7 cells
IC50 = 1650 nmol/L
HCT-116 cells
IC50 = 3610 nmol/L
81 [180]
EGFR/AKT
Chalcones
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbd_HTML.png
A549 cells
IC50 = 3820 nmol/L
MDA-MB-231 cells
IC50 = 5890 nmol/L
SKBR3 cells
IC50 = 4790 nmol/L
82 [181]
EGFR/HDACs
Pyrimidine-2-amines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbe_HTML.png
EGFRWT
IC50 = 5700 nmol/L
EGFRT790M
IC50 = 5000 nmol/L
HDACs
IC50 = 85 nmol/L
A549 cells
IC50 = 2190 nmol/L
HeLa cells
IC50 = 1850 nmol/L
MDA-MB-231 cells
IC50 = 600 nmol/L
MDA-MB-468 cells
IC50 = 230 nmol/L
83 [182]
EGFR/PDGFR-β
Pyrimidine-2,4-diamines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbf_HTML.png
EGFR Ki
IC50 = 170 nmol/L
PDGFR-β
IC50 = 81 nmol/L
84 [183]
EGFR/NF-κB
Quinazolines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbg_HTML.png
EGFR
IC50 = 60.1 nmol/L
NF-κB
IC50 = 300 nmol/L
85 [184]
EGFR/c-Met
1,2,4-Oxadiazole derivate
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbh_HTML.png
MDA-MB-231 cells
IC50 = 200 nmol/L
A459 cells IC50 = 200 nmol/L
PC9 cells IC50 = 500 nmol/L
H1975 cells IC50 = 300 nmol/L
CL68 cells IC50 = 400 nmol/L
CL97 cells
IC50 = 500 nmol/L
86 [185]
EGFRT790M/c-Met
Pyrimidine-2-amines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbi_HTML.png
EGFRT790M
IC50 = 97 nmol/L
c-Met
IC50 = 518 nmol/L
87 [186]
EGFR/VEGFR-2
Pyrimidine-2,4-diamines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbj_HTML.png
EGFR Ki
IC50 = 80 nmol/L
VEGFR-2 Ki
IC50 = 3240 nmol/L
NCI-H460 cells GI = 25% (10 µmol/L)
88 [187]
EGFR/VEGFR-2
Quinazolines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbk_HTML.png
EGFR
IC50 = 1.0 nmol/L
VEGFR-2
IC50 = 79.0 nmol/L
HT-29 cells
IC50 = 1760 nmol/L
MCF7 cells
IC50 = 7280 nmol/L
89 [188]
EGFR/VEGFR-2
Quinazolines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbl_HTML.png
EGFR
IC50 = 0.69 nmol/L
VEGFR-2
IC50 = 67.84 nmol/L
90 [59]
EGFR/VEGFR-2
Quinazolines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbm_HTML.png
EGFR
IC50 = 2.0 nmol/L
VEGFR-2
IC50 = 103.0 nmol/L
A431 cells
IC50 = 14.0 nmol/L
H1975 cells
IC50 = 130.0 nmol/L
91 [189]
EGFR/VEGFR-2
Quinazolines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbn_HTML.png
EGFR
IC50 = 20 nmol/L
VEGFR-2
IC50 = 50 nmol/L
92 [190]
EGFR/HER2
Quinazolines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbo_HTML.png
EGFR
IC50 = 0.69 nmol/L
HER2
IC50 = 42.1 nmol/L
NCI-H1975 cells
IC50 = 12.20 nmol/L
HCC827 cells
IC50 = 0.31 nmol/L
A431 cells
IC50 = 1.52 nmol/L
MDA-MB-453 cells
IC50 = 0.62 nmol/L
93 [191]
EGFR/HER2
Pyrimidine-4-amines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbp_HTML.png
EGFR
IC50 = 186 nmol/L
VEGFR-2
IC50 = 254 nmol/L
94 [192]
EGFR/HER2
3-Cyanoquizolines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbq_HTML.png
EGFR
IC50 = 597 nmol/L
IGF1R
IC50 = 908 nmol/L
A431 cells
IC50 = 1890 nmol/L
SKBR3 cells
IC50 = 1930 nmol/L
95 [193195]
EGFR/HER2
Pyrimidinones
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbr_HTML.png
EGFR
IC50 = 60 nmol/L
HER2
IC50 = 300 nmol/L
A549 cells
IC50 = 280 nmol/L
96 [196]
EGFR/CSK
Chalcones
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbs_HTML.png
EGFR
IC50 = 11,120 nmol/L
CSK
IC50 = 5160 nmol/L
97 [197]
EGFR/CAIX
Quinazolines
https://static-content.springer.com/image/art%3A10.1186%2Fs13045-022-01311-6/MediaObjects/13045_2022_1311_Figbt_HTML.png
EGFRWT
IC50 = 27.0 nmol/L
EGFRT790M
IC50 = 9.2 nmol/L
hCAII
IC50 = 278.2 nmol/L
hCAIX
IC50 = 115.0 nmol/L
A549 cells (hypoxia)
IC50 = 2210 nmol/L
A549 cells (normoxia)
IC50 = 6450 nmol/L
HeLa cells
IC50 = 1850 nmol/L
H1975 cells (hypoxia)
IC50 = 1050 nmol/L
H1975 cells (normoxia)
IC50 = 1940 nmol/L

Natural products

The discovery of natural products offers new scaffolds for drug development. Natural products are an important source of compounds to overcome resistance to third-generation TKIs and provide ample possibilities for new drug discovery. Honokiol (HNK) (98) is a natural product purified from Magnolia used as a human nutritional supplement, with good tolerance and safety profiles. Many preclinical studies have shown that HNK (98) has potential antitumor activity against different types of cancer. Zang et al. proved that the decrease in Mcl-1 and the increase in BIM are the key mechanisms by which osimertinib induces the apoptosis of NSCLC cells with EGFR-TKI-sensitive mutations. HNK (98) and its derivative CAz-p (99) in combination with osimertinib effectively reduced the survival and induced the apoptosis of EGFR ex19del/C797S (trans) double-mutant PC-9/2 M cells and EGFR ex19del/T790M/C797S (cis) triple-mutant PC-9/3 M cells [198]. It is highly encouraging that HNK (98) and its derivatives may overcome clinical resistance to third-generation TKIs.
Overexpression of MCL-1 induces acquired resistance to osimertinib. Combination therapy with MCL-1 inhibitors and osimertinib is a potential strategy to overcome resistance. Bufalin (100) is a natural product that belongs to the class of bufadienolide analogs. A recent study found that bufalin (100) can reverse acquired resistance to osimertinib by inducing Ku70-mediated Mcl-1 degradation. Moreover, combined treatment with bufalin (100) and osimertinib triggered significant cell apoptosis and increased the levels of cleaved caspase-3 and PARP [199].
Wighteone (101) is a natural flavonoid compound widely found in plants. Sun et al. reported that wighteone (101) docks at the ATP binding site of EGFR L858R/T790M and forms two hydrogen bonds with the carbonyl group of Gln791 and the amino group of Met793, indicating that it may directly bind to EGFR L858R/T790M. Wighteone has a significant inhibitory effect on Ba/F3 and NCI-H1975 cells expressing EGFR L858R/T790M, with IC50 values of 1.88 μmol/L and 5.70 μmol/L, respectively [200] (Fig. 14).

Other strategies

EGFR degradation based on the FBXL2-Grp94-EGFR axis

Xiao’s research group found that the F-box protein Fbxl2 (an E3 ubiquitin ligase) can target EGFR and EGFR-TKI-resistant mutants for proteasome-mediated degradation independent of EGF stimulation. They also discovered that glucose regulatory protein 94 (Grp94) protects EGFR from degradation by blocking the binding of Fbxl2 to EGFR. Through virtual screening of the DrugBank database, small compounds that can bind to the Fbxo3-apag domain were scored. Nebivolol (102) can be placed in the dumbbell-shaped cavity of the APAG region. There are 5 amino acid residues in the center of this cavity (I331, E341, T367, T368 and F369); T367 and T368 project into the cavities of complementary shapes, forming hydrophobic interactions with the ligand. The binding affinity of the Fbxo3 protein for endogenous Fbxl2 is greatly reduced when these five amino acids are mutated individually or in combination. Data suggest the potential of nebivolol (102) as a small molecule that can disrupt the Fbxo3–Fbxl2 interaction. Increasing Fbxl2 levels with nebivolol (102) (Fig. 14) in combination with osimertinib or a Grp94 inhibitor (ganetespib) to target the FBXL2-Grp94-EGFR axis and thus destabilize EGFR is a possible therapeutic strategy to overcome resistance to third-generation EGFR-TKIs [201].

AKR1B1 inhibitors

Zhang et al. discovered that aldehyde ketone reductase family 1 member B1 (AKR1B1) interacts with STAT3 and activates the cystine transporter solute carrier family 7 member 11 (SLC7A11), which in turn leads to enhanced cystine uptake, glutathione synthesis flux, clearance of reactive oxygen species (ROS), protection against cell death, and EGFR-TKI resistance. The use of selective inhibitors (including the clinically approved anti-diabetic drug epalrestat) to inhibit AKR1B1 can restore the sensitivity of drug-resistant cell lines to EGFR-TKIs and delay drug resistance in mice harboring xenografted tumors derived from lung cancer patients [202].

PGAM1 inhibitors

Phosphoglycerate mutase 1 (PGAM1) is an important enzyme in the glycolysis pathway and is related to tumor cell metastasis [203]. HKB99 (103) (Fig. 14) is an allosteric inhibitor of PGAM1 that significantly inhibits the growth and metastasis of NSCLC by affecting the metabolic activity and nonmetabolic functions of PGAM1 [204]. The docking model of the PGAM1-HKB99 complex shows that HKB99 (103) binds to the allosteric site of the adjacent substrate-binding pocket of PGAM1, thereby inhibiting the conversion of 3-PG to 2-PG and significantly reducing the metabolic activity of PGAM1. In addition, HKB99 (103) can allosterically bind to PGAM1, weaken the interaction between PGAM1 and ACTA2, and inhibit the growth and metastasis of erlotinib-resistant lung cancer cells [205, 206]. Therefore, PGAM1 is a metabolic enzyme that may overcome EGFR-TKI resistance.

Nonoverlapping allosteric pockets—the X-Pocket

Qiu et al. revealed the underlying mechanism of reverse allosteric communication in dual-targeted therapy. Allosteric sites can be affected by orthomorphic drugs. The nonoverlapping allosteric pocket X-Pocket was discovered in EGFR mutants; this pocket is mainly composed of nonconserved residues, including the hot spots K867, S895, and K960, that can cooperate with traditional TKIs [207]. It is a promising target for the design of selective conformationally restricted drugs, with great potential in terms of affinity, efficacy, and selectivity.

DZ-SIM inhibitors

In addition, researchers found that a group of near-infrared heptamethine carbocyanine (DZ) fluorescent dyes, the prototype of which is heptamethylamine carbocyanine dye (IR-783) (104) (Fig. 14), have tumor-targeting activity through differentially expressed organic anion transport peptides on cancer cells [208]. This group of organic dyes can specifically deliver therapeutic payloads to tumor cells in the form of chemical conjugates. DZ-SIM was preliminarily synthesized; SIM specifically targets 3-hydroxy-3-methylglutaryl-CoA reductase (HMGCR) in the endoplasmic reticulum. After specific uptake by tumor cells, DZ-SIM was enriched in subcellular organelles (including mitochondria and lysosomes). NSCLC cells were killed by mitochondrial damage, which mainly led to cytochrome C release into the cytoplasm, thereby activating the caspase-3-dependent apoptosis cascade. DZ-SIM inhibited the formation of cancer cell colonies resistant to first-generation (H1650 and H1975) and third-generation EGFR-TKIs (PC9AR), and most IC50 values were lower than 10 μmol/L. DZ-SIM represents a promising new therapy to overcome drug resistance in patients with EGFR-mutant disease.

Selection of individualized combination therapy

For patients who experience SCLC transformation, chemotherapy after the development of osimertinib resistance is an option. Research has shown that patients with transformation to SCLC have higher response rates to etoposide, cisplatin, and paclitaxel. For patients with unclear resistance mechanisms, chemotherapy is still a treatment option. If the patient is asymptomatic or has symptomatic local progression, osimertinib can be combined with local treatment according to National Comprehensive Cancer Network (NCCN) guidelines. Carboplatin, paclitaxel, bevacizumab, and atezolizumab (anti-PD-L1 antibody) are also options for patients who experience systemic progression after osimertinib treatment [209]. Whether chemotherapy can delay the development of resistance to third-generation EGFR-TKIs remains unknown. A study on osimertinib with or without chemotherapy as first-line therapy for patients with EGFR-mutant NSCLC is currently recruiting (NCT04035486) [210].
For most patients, the PD-1/PD-L1 pathway is not the sole rate-limiting factor for antitumor immunity, and blocking the PD-1/PD-L1 axis is insufficient to activate an effective antitumor immune response [211]. Strategies that lead to acquired EGFR-TKI resistance, such as HGF, MET amplification, and EGFR T790M, also promote immune escape in lung cancer by upregulating the expression of PD-L1. Many combination strategies, including α-PD-1/PD-L1 plus chemotherapy, radiotherapy, angiogenesis inhibitors, targeted therapy, other ICIs, agonists of the costimulatory molecule, stimulator of interferon gene agonists, epigenetic modulators, or metabolic modulators, have been confirmed to have superior antitumor efficacy and a higher response rate. The immunomodulatory effect of chemotherapy suggests that it might be a suitable partner for combination with α-PD-1/PD-L1 to achieve rapid and long-term cancer control. During the KEYNOTE series of clinical trials (such as KEYNOTE-021, KEYNOTE-189, and KEYNOTE-407), pembrolizumab combined with standard chemotherapy led to better overall survival (OS) and progression-free survival (PFS) in NSCLC patients and has been approved by the FDA as first-line treatment for advanced nonsquamous NSCLC [212, 213]. In addition, the National Medical Products Administration (NMPA) approved sintilimab plus gemcitabine and platinum as first-line treatment for advanced squamous NSCLC based on the results of ORIENT-12 [214]. In addition to α-PD-1-based approaches, α-PD-L1-based chemoimmunotherapy has also attracted intense attention. The IMpower150 trial was the pioneer of this series of studies, and the FDA-approved atezolizumab plus bevacizumab, paclitaxel, and carboplatin as first-line treatment for advanced nonsquamous NSCLC [215]. Subsequently, the FDA-approved atezolizumab plus nab-paclitaxel and carboplatin for nonsquamous NSCLC (based on the results of IMpower130). Radiotherapy can also induce immunogenic cell death and enhance the antitumor immune response. The results of a phase 1 study showed that α-PD-1/PD-L1 plus chemoradiotherapy was tolerable in advanced NSCLC (NCT02621398), with promising clinical outcomes. In multiple clinical studies, such as IMpower150, angiogenesis inhibitors enhanced the efficacy of α-PD-1/PD-L1 [216]. Moreover, dual immune checkpoint blockade or costimulatory molecule agonists plus α-PD-1/PD-L1 are also promising strategies. To date, the FDA has approved ipilimumab plus nivolumab for NSCLC and melanoma, among others. Agonists targeting costimulatory pathways such as CD27/CD70, CD40/CD40L, and 4-1BB/4-1BBL could also enhance T-cell activity and restore the antitumor immune response. However, bispecific/bifunctional antibodies simultaneously block two molecules and thus have a strategic advantage over combination therapy. For example, in the phase 1 NCT03710265 trial, SHR-1701 (TGF-β × PD-L1 bifunctional antibody) showed encouraging antitumor activity [217].
Given the heterogeneity of mutations across patients, the selection of individualized combination treatment strategies could improve outcomes and mitigate treatment resistance.

Discussion and future perspectives

EGFR is an important target on tumor cells that promotes mitosis and transformation. It is overexpressed in many diseases and is particularly related to the occurrence and development of cancer [3, 7, 8]. Tumors often have prominent genomic and transcriptional heterogeneity that is closely related to EGFR-TKI resistance [40, 218]. Studies have shown that drug resistance can develop through EGFR-dependent and EGFR-independent mechanisms [24, 219]. The emergence of resistance to third-generation EGFR-TKIs limits the clinical benefits for patients, thus necessitating the further development of more effective strategies.
To date, fourth-generation EGFR-TKIs show prominent antitumor activity. Recent research has shown that fourth-generation inhibitors targeting allosteric sites and ATP-competitive sites of EGFR can achieve remarkable results against EGFR L858R/T790M and C797S. In addition to fourth-generation EGFR-TKIs, combination treatments, monoclonal antibodies, and bispecific antibodies are significantly contributing to the treatment of tumors harboring the C797S mutation. While the C797S mutation is only one of the numerous drug resistance mechanisms, it is necessary to overcome other mutations by designing and developing new noncovalent ATP-competitive inhibitors that form hydrogen bonds with mutated residues in the EGFR ATP pocket (such as Lys745 and Asn842). The rational design of selective EGFR inhibitors that bind to both the ATP and allosteric sites of the EGFR kinase domain, that is, adding allosteric inhibitor elements to the compound skeleton at the ATP binding site, will help optimize and improve the mutation selectivity of compounds and lead to the identification of small molecules with good kinase inhibitory activity. However, the cellular activity of such compounds needs to be further improved, and future research directions should focus on the structural optimization of current lead compounds to obtain EGFR inhibitors with better mutation selectivity. Targeted protein degradation technology provides a new research direction for overcoming resistance to third-generation EGFR inhibitors. Considering the significance of overcoming allosteric hindrance by triple-mutant EGFR, allosteric EGFR degraders were developed. In addition, dual PROTACs have emerged in the field of cancer combination therapy; dual PROTACs can be designed with two targets, such as tumor immune targets plus adjuvant immune targets or energy metabolism targets and epigenetic targets plus antiapoptotic targets, to further overcome resistance of third-generation EGFR inhibitors and provide a better curative effect. Of course, the larger molecular weight of dual PROTACs will affect their druggability and pharmacokinetics, but perhaps nanodrug delivery systems can be utilized to improve drug absorption or optimized by simplifying the inhibitor moiety and maintaining the minimum pharmacophore. In addition, ADCs containing a small-molecule cytotoxic compound and a monoclonal antibody targeting a cancer target have attracted attention. The ADC MRG003 has entered clinical trials with great development and application prospects. The activation of alternative pathways and histological transformation are important mechanisms of resistance to third-generation EGFR inhibitors. The combined use of third-generation inhibitors and related pathway blockers is another important approach. To prevent the toxicity and side effects of multidrug combinations, drugs with multiple pharmacological activities were developed and proven to have more advantages than combination therapy. Multitarget kinase drugs have become a favorable choice due to their attractive pharmacokinetic characteristics and safety profiles. Natural compounds have received much research attention due to their potential antitumor effects. Based on the molecular mechanism of inhibition, natural compounds can be modified to provide new insights for effectively overcoming resistance to third-generation EGFR-TKIs. Last but not least, the discovery of EGFR degraders based on the FBXL2-Grp94-EGFR axis, AKR1B1 and PGAM1 inhibitors, DZ-SIM, and the nonoverlapping allosteric pocket X-Pocket provides promising support for the further development of strategies to overcome resistance to third-generation EGFR inhibitors. The mechanism of resistance to third-generation EGFR-TKIs is very complex, is impacted by EGFR mutations, and differs among patients and tumor sites. Thus, next-generation sequencing (NGS) of blood-based circulating tumor DNA (ctDNA) or tissue samples to elucidate the resistance mechanism will be valuable for guiding future therapeutic approaches and for clinical research on novel combination therapies to overcome drug resistance. Moreover, individualized combination treatment strategies could also improve treatment efficacy and mitigate treatment resistance.
EGFR is a verified target for antitumor therapy in a broad spectrum of cancers. Facilitated by versatile strategies in the field of medicinal chemistry, better approaches are anticipated for overcoming the hurdle of drug resistance to provide new hope for patients.

Conclusion

As a crucial “controller” that is related to the inhibition of tumor cell proliferation, angiogenesis, invasion, metastasis, and apoptosis, EGFR actively participates in malignant disease progression. However, the intrinsic and acquired resistance in primary and recurrent cancer which is mediated by EGFR mutations after target treatment leads to difficult therapeutic. Understanding the complex resistance mechanisms of EGFR-TKIs and developing potential strategies to combat it could be of potential interest for improving the individual therapeutic strategies for cancer.

Acknowledgements

Not applicable.

Declarations

This is not applicable for this review.
This is not applicable for this review.

Competing interests

The authors declare no competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
insite
INHALT
download
DOWNLOAD
print
DRUCKEN

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

e.Med Innere Medizin

Kombi-Abonnement

Mit e.Med Innere Medizin erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes Innere Medizin, den Premium-Inhalten der internistischen Fachzeitschriften, inklusive einer gedruckten internistischen Zeitschrift Ihrer Wahl.

Literatur
1.
2.
Zurück zum Zitat da Cunha SG, Shepherd FA, Tsao MS. EGFR mutations and lung cancer. Annu Rev Pathol. 2011;6:49–69.CrossRef da Cunha SG, Shepherd FA, Tsao MS. EGFR mutations and lung cancer. Annu Rev Pathol. 2011;6:49–69.CrossRef
3.
Zurück zum Zitat Campbell ID, Bork P. Epidermal growth factor-like modules. Curr Opin Struct Biol. 1993;3:385–92.CrossRef Campbell ID, Bork P. Epidermal growth factor-like modules. Curr Opin Struct Biol. 1993;3:385–92.CrossRef
4.
Zurück zum Zitat Roskoski R. The ERBB/HER family of protein-tyrosine kinases and cancer. Pharmacol Res. 2014;79:34–74.PubMedCrossRef Roskoski R. The ERBB/HER family of protein-tyrosine kinases and cancer. Pharmacol Res. 2014;79:34–74.PubMedCrossRef
5.
Zurück zum Zitat Gazdar AF. Activating and resistance mutations of EGFR in non-small-cell lung cancer: role in clinical response to EGFR tyrosine kinase inhibitors. Oncogene. 2009;28(Suppl 1):S24–31.PubMedPubMedCentralCrossRef Gazdar AF. Activating and resistance mutations of EGFR in non-small-cell lung cancer: role in clinical response to EGFR tyrosine kinase inhibitors. Oncogene. 2009;28(Suppl 1):S24–31.PubMedPubMedCentralCrossRef
6.
Zurück zum Zitat Herbst RS, Langer CJ. Epidermal growth factor receptors as a target for cancer treatment: the emerging role of IMC-C225 in the treatment of lung and head and neck cancers. Semin Oncol. 2002;29:27–36.PubMedCrossRef Herbst RS, Langer CJ. Epidermal growth factor receptors as a target for cancer treatment: the emerging role of IMC-C225 in the treatment of lung and head and neck cancers. Semin Oncol. 2002;29:27–36.PubMedCrossRef
7.
Zurück zum Zitat Normanno N, Bianco C, De Luca A, Salomon DS. The role of EGF-related peptides in tumor growth. Front Biosci. 2001;6:D685–707.PubMedCrossRef Normanno N, Bianco C, De Luca A, Salomon DS. The role of EGF-related peptides in tumor growth. Front Biosci. 2001;6:D685–707.PubMedCrossRef
9.
Zurück zum Zitat Sabbah DA, Hajjo R, Sweidan K. Review on epidermal growth factor receptor (EGFR) structure, signaling pathways, interactions, and recent updates of EGFR inhibitors. Curr Top Med Chem. 2020;20:815–34.PubMedCrossRef Sabbah DA, Hajjo R, Sweidan K. Review on epidermal growth factor receptor (EGFR) structure, signaling pathways, interactions, and recent updates of EGFR inhibitors. Curr Top Med Chem. 2020;20:815–34.PubMedCrossRef
10.
Zurück zum Zitat Cross DA, Ashton SE, Ghiorghiu S, Eberlein C, Nebhan CA, Spitzler PJ, et al. AZD9291, an irreversible EGFR TKI, overcomes T790M-mediated resistance to EGFR inhibitors in lung cancer. Cancer Discov. 2014;4:1046–61.PubMedPubMedCentralCrossRef Cross DA, Ashton SE, Ghiorghiu S, Eberlein C, Nebhan CA, Spitzler PJ, et al. AZD9291, an irreversible EGFR TKI, overcomes T790M-mediated resistance to EGFR inhibitors in lung cancer. Cancer Discov. 2014;4:1046–61.PubMedPubMedCentralCrossRef
11.
Zurück zum Zitat Ricordel C, Friboulet L, Facchinetti F, Soria JC. Molecular mechanisms of acquired resistance to third-generation EGFR-TKIs in EGFR T790M-mutant lung cancer. Ann Oncol. 2018;29:i28–37.PubMedCrossRef Ricordel C, Friboulet L, Facchinetti F, Soria JC. Molecular mechanisms of acquired resistance to third-generation EGFR-TKIs in EGFR T790M-mutant lung cancer. Ann Oncol. 2018;29:i28–37.PubMedCrossRef
12.
Zurück zum Zitat Roskoski R Jr. Classification of small molecule protein kinase inhibitors based upon the structures of their drug-enzyme complexes. Pharmacol Res. 2016;103:26–48.PubMedCrossRef Roskoski R Jr. Classification of small molecule protein kinase inhibitors based upon the structures of their drug-enzyme complexes. Pharmacol Res. 2016;103:26–48.PubMedCrossRef
13.
Zurück zum Zitat Sequist LV, Yang JC, Yamamoto N, O’Byrne K, Hirsh V, Mok T, et al. Phase iii study of afatinib or cisplatin plus pemetrexed in patients with metastatic lung adenocarcinoma with EGFR mutations. J Clin Oncol. 2013;31:3327–34.PubMedCrossRef Sequist LV, Yang JC, Yamamoto N, O’Byrne K, Hirsh V, Mok T, et al. Phase iii study of afatinib or cisplatin plus pemetrexed in patients with metastatic lung adenocarcinoma with EGFR mutations. J Clin Oncol. 2013;31:3327–34.PubMedCrossRef
14.
Zurück zum Zitat Wu YL, Zhou C, Hu CP, Feng J, Lu S, Huang Y, et al. Afatinib versus cisplatin plus gemcitabine for first-line treatment of Asian patients with advanced non-small-cell lung cancer harbouring EGFR mutations (LUX-Lung 6): an open-label, randomised phase 3 trial. Lancet Oncol. 2014;15:213–22.PubMedCrossRef Wu YL, Zhou C, Hu CP, Feng J, Lu S, Huang Y, et al. Afatinib versus cisplatin plus gemcitabine for first-line treatment of Asian patients with advanced non-small-cell lung cancer harbouring EGFR mutations (LUX-Lung 6): an open-label, randomised phase 3 trial. Lancet Oncol. 2014;15:213–22.PubMedCrossRef
15.
Zurück zum Zitat Zhang H. Three generations of epidermal growth factor receptor tyrosine kinase inhibitors developed to revolutionize the therapy of lung cancer. Drug Des Devel Ther. 2016;10:3867–72.PubMedPubMedCentralCrossRef Zhang H. Three generations of epidermal growth factor receptor tyrosine kinase inhibitors developed to revolutionize the therapy of lung cancer. Drug Des Devel Ther. 2016;10:3867–72.PubMedPubMedCentralCrossRef
16.
17.
Zurück zum Zitat Jiang T, Zhou C. Clinical activity of the mutant-selective EGFR inhibitor AZD9291 in patients with EGFR inhibitor-resistant non-small cell lung cancer. Transl Lung Cancer Res. 2014;3:370–2.PubMedPubMedCentral Jiang T, Zhou C. Clinical activity of the mutant-selective EGFR inhibitor AZD9291 in patients with EGFR inhibitor-resistant non-small cell lung cancer. Transl Lung Cancer Res. 2014;3:370–2.PubMedPubMedCentral
18.
Zurück zum Zitat Yan XE, Zhu SJ, Liang L, Zhao P, Choi HG, Yun CH. Structural basis of mutant-selectivity and drug-resistance related to CO-1686. Oncotarget. 2017;8:53508–17.PubMedPubMedCentralCrossRef Yan XE, Zhu SJ, Liang L, Zhao P, Choi HG, Yun CH. Structural basis of mutant-selectivity and drug-resistance related to CO-1686. Oncotarget. 2017;8:53508–17.PubMedPubMedCentralCrossRef
19.
Zurück zum Zitat 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;121:725–37.PubMedPubMedCentralCrossRef 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;121:725–37.PubMedPubMedCentralCrossRef
20.
Zurück zum Zitat Hata AN, Niederst MJ, Archibald HL, Gomez-Caraballo M, Siddiqui FM, Mulvey HE, et al. Tumor cells can follow distinct evolutionary paths to become resistant to epidermal growth factor receptor inhibition. Nat Med. 2016;22:262–9.PubMedPubMedCentralCrossRef Hata AN, Niederst MJ, Archibald HL, Gomez-Caraballo M, Siddiqui FM, Mulvey HE, et al. Tumor cells can follow distinct evolutionary paths to become resistant to epidermal growth factor receptor inhibition. Nat Med. 2016;22:262–9.PubMedPubMedCentralCrossRef
21.
Zurück zum Zitat Guibert N, Barlesi F, Descourt R, Lena H, Besse B, Beau-Faller M, et al. Characteristics and outcomes of patients with lung cancer harboring multiple molecular alterations: results from the IFCT study biomarkers france. J Thorac Oncol. 2017;12:963–73.PubMedCrossRef Guibert N, Barlesi F, Descourt R, Lena H, Besse B, Beau-Faller M, et al. Characteristics and outcomes of patients with lung cancer harboring multiple molecular alterations: results from the IFCT study biomarkers france. J Thorac Oncol. 2017;12:963–73.PubMedCrossRef
22.
Zurück zum Zitat Li X, Wang S, Li B, Wang Z, Shang S, Shao Y, et al. Bim deletion polymorphism confers resistance to osimertinib in EGFR T790M lung cancer: a case report and literature review. Target Oncol. 2018;13:517–23.PubMedCrossRef Li X, Wang S, Li B, Wang Z, Shang S, Shao Y, et al. Bim deletion polymorphism confers resistance to osimertinib in EGFR T790M lung cancer: a case report and literature review. Target Oncol. 2018;13:517–23.PubMedCrossRef
23.
Zurück zum Zitat Eck MJ, Yun CH. Structural and mechanistic underpinnings of the differential drug sensitivity of EGFR mutations in non-small cell lung cancer. Biochim Biophys Acta. 2010;1804:559–66.PubMedCrossRef Eck MJ, Yun CH. Structural and mechanistic underpinnings of the differential drug sensitivity of EGFR mutations in non-small cell lung cancer. Biochim Biophys Acta. 2010;1804:559–66.PubMedCrossRef
24.
Zurück zum Zitat Westover D, Zugazagoitia J, Cho BC, Lovly CM, Paz-Ares L. Mechanisms of acquired resistance to first- and second-generation EGFR tyrosine kinase inhibitors. Ann Oncol. 2018;29:i10–9.PubMedPubMedCentralCrossRef Westover D, Zugazagoitia J, Cho BC, Lovly CM, Paz-Ares L. Mechanisms of acquired resistance to first- and second-generation EGFR tyrosine kinase inhibitors. Ann Oncol. 2018;29:i10–9.PubMedPubMedCentralCrossRef
25.
Zurück zum Zitat Zhou W, Ercan D, Chen L, Yun CH, Li D, Capelletti M, et al. Novel mutant-selective EGFR kinase inhibitors against EGFR T790M. Nature. 2009;462:1070–4.PubMedPubMedCentralCrossRef Zhou W, Ercan D, Chen L, Yun CH, Li D, Capelletti M, et al. Novel mutant-selective EGFR kinase inhibitors against EGFR T790M. Nature. 2009;462:1070–4.PubMedPubMedCentralCrossRef
26.
Zurück zum Zitat Zhang Q, Zhang XC, Yang JJ, Yang ZF, Bai Y, Su J, et al. EGFR L792H and G796R: two novel mutations mediating resistance to the third-generation EGFR tyrosine kinase inhibitor osimertinib. J Thorac Oncol. 2018;13:1415–21.PubMedCrossRef Zhang Q, Zhang XC, Yang JJ, Yang ZF, Bai Y, Su J, et al. EGFR L792H and G796R: two novel mutations mediating resistance to the third-generation EGFR tyrosine kinase inhibitor osimertinib. J Thorac Oncol. 2018;13:1415–21.PubMedCrossRef
27.
28.
Zurück zum Zitat Ou SI, Cui J, Schrock AB, Goldberg ME, Zhu VW, Albacker L, et al. Emergence of novel and dominant acquired EGFR solvent-front mutations at Gly796 (G796S/R) together with C797S/R and L792F/H mutations in one EGFR (L858R/T790M) NSCLC patient who progressed on osimertinib. Lung Cancer. 2017;108:228–31.PubMedCrossRef Ou SI, Cui J, Schrock AB, Goldberg ME, Zhu VW, Albacker L, et al. Emergence of novel and dominant acquired EGFR solvent-front mutations at Gly796 (G796S/R) together with C797S/R and L792F/H mutations in one EGFR (L858R/T790M) NSCLC patient who progressed on osimertinib. Lung Cancer. 2017;108:228–31.PubMedCrossRef
29.
Zurück zum Zitat Castellano GM, Aisner J, Burley SK, Vallat B, Yu HA, Pine SR, et al. A novel acquired exon 20 EGFR M766Q mutation in lung adenocarcinoma mediates osimertinib resistance but is sensitive to neratinib and poziotinib. J Thorac Oncol. 2019;14:1982–8.PubMedPubMedCentralCrossRef Castellano GM, Aisner J, Burley SK, Vallat B, Yu HA, Pine SR, et al. A novel acquired exon 20 EGFR M766Q mutation in lung adenocarcinoma mediates osimertinib resistance but is sensitive to neratinib and poziotinib. J Thorac Oncol. 2019;14:1982–8.PubMedPubMedCentralCrossRef
30.
Zurück zum Zitat Liu J, Jin B, Su H, Qu X, Liu Y. Afatinib helped overcome subsequent resistance to osimertinib in a patient with NSCLC having leptomeningeal metastasis baring acquired EGFR L718Q mutation: a case report. BMC Cancer. 2019;19:702.PubMedPubMedCentralCrossRef Liu J, Jin B, Su H, Qu X, Liu Y. Afatinib helped overcome subsequent resistance to osimertinib in a patient with NSCLC having leptomeningeal metastasis baring acquired EGFR L718Q mutation: a case report. BMC Cancer. 2019;19:702.PubMedPubMedCentralCrossRef
31.
Zurück zum Zitat Bersanelli M, Minari R, Bordi P, Gnetti L, Bozzetti C, Squadrilli A, et al. L718Q mutation as new mechanism of acquired resistance to AZD9291 in EGFR -mutated NSCLC. J Thorac Oncol. 2016;11:e121–3.PubMedCrossRef Bersanelli M, Minari R, Bordi P, Gnetti L, Bozzetti C, Squadrilli A, et al. L718Q mutation as new mechanism of acquired resistance to AZD9291 in EGFR -mutated NSCLC. J Thorac Oncol. 2016;11:e121–3.PubMedCrossRef
32.
Zurück zum Zitat Callegari D, Ranaghan KE, Woods CJ, Minari R, Tiseo M, Mor M, et al. L718Q mutant EGFR escapes covalent inhibition by stabilizing a non-reactive conformation of the lung cancer drug osimertinib. Chem Sci. 2018;9:2740–9.PubMedPubMedCentralCrossRef Callegari D, Ranaghan KE, Woods CJ, Minari R, Tiseo M, Mor M, et al. L718Q mutant EGFR escapes covalent inhibition by stabilizing a non-reactive conformation of the lung cancer drug osimertinib. Chem Sci. 2018;9:2740–9.PubMedPubMedCentralCrossRef
33.
Zurück zum Zitat Yang Z, Yang J, Chen Y, Shao YW, Wang X. Acquired EGFR L718V mutation as the mechanism for osimertinib resistance in a T790M-negative non-small-cell lung cancer patient. Target Oncol. 2019;14:369–74.PubMedCrossRef Yang Z, Yang J, Chen Y, Shao YW, Wang X. Acquired EGFR L718V mutation as the mechanism for osimertinib resistance in a T790M-negative non-small-cell lung cancer patient. Target Oncol. 2019;14:369–74.PubMedCrossRef
34.
Zurück zum Zitat Fassunke J, Muller F, Keul M, Michels S, Dammert MA, Schmitt A, et al. Overcoming EGFR (G724S)-mediated osimertinib resistance through unique binding characteristics of second-generation EGFR inhibitors. Nat Commun. 2018;9:4655.PubMedPubMedCentralCrossRef Fassunke J, Muller F, Keul M, Michels S, Dammert MA, Schmitt A, et al. Overcoming EGFR (G724S)-mediated osimertinib resistance through unique binding characteristics of second-generation EGFR inhibitors. Nat Commun. 2018;9:4655.PubMedPubMedCentralCrossRef
35.
Zurück zum Zitat Tu HY, Ke EE, Yang JJ, Sun YL, Yan HH, Zheng MY, et al. A comprehensive review of uncommon EGFR mutations in patients with non-small cell lung cancer. Lung Cancer. 2017;114:96–102.PubMedCrossRef Tu HY, Ke EE, Yang JJ, Sun YL, Yan HH, Zheng MY, et al. A comprehensive review of uncommon EGFR mutations in patients with non-small cell lung cancer. Lung Cancer. 2017;114:96–102.PubMedCrossRef
36.
Zurück zum Zitat Xu J, Jin B, Chu T, Dong X, Yang H, Zhang Y, et al. EGFR tyrosine kinase inhibitor (TKI) in patients with advanced non-small cell lung cancer (NSCLC) harboring uncommon EGFR mutations: a real-world study in china. Lung Cancer. 2016;96:87–92.PubMedCrossRef Xu J, Jin B, Chu T, Dong X, Yang H, Zhang Y, et al. EGFR tyrosine kinase inhibitor (TKI) in patients with advanced non-small cell lung cancer (NSCLC) harboring uncommon EGFR mutations: a real-world study in china. Lung Cancer. 2016;96:87–92.PubMedCrossRef
37.
Zurück zum Zitat Shen YC, Tseng GC, Tu CY, Chen WC, Liao WC, Chen WC, et al. Comparing the effects of afatinib with gefitinib or erlotinib in patients with advanced-stage lung adenocarcinoma harboring non-classical epidermal growth factor receptor mutations. Lung Cancer. 2017;110:56–62.PubMedCrossRef Shen YC, Tseng GC, Tu CY, Chen WC, Liao WC, Chen WC, et al. Comparing the effects of afatinib with gefitinib or erlotinib in patients with advanced-stage lung adenocarcinoma harboring non-classical epidermal growth factor receptor mutations. Lung Cancer. 2017;110:56–62.PubMedCrossRef
38.
Zurück zum Zitat Ercan D, Choi HG, Yun CH, Capelletti M, Xie T, Eck MJ, et al. EGFR mutations and resistance to irreversible pyrimidine-based EGFR inhibitors. Clin Cancer Res. 2015;21:3913–23.PubMedPubMedCentralCrossRef Ercan D, Choi HG, Yun CH, Capelletti M, Xie T, Eck MJ, et al. EGFR mutations and resistance to irreversible pyrimidine-based EGFR inhibitors. Clin Cancer Res. 2015;21:3913–23.PubMedPubMedCentralCrossRef
39.
Zurück zum Zitat Piotrowska Z, Isozaki H, Lennerz JK, Gainor JF, Lennes IT, Zhu VW, 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;8:1529–39.PubMedPubMedCentralCrossRef Piotrowska Z, Isozaki H, Lennerz JK, Gainor JF, Lennes IT, Zhu VW, 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;8:1529–39.PubMedPubMedCentralCrossRef
40.
Zurück zum Zitat Piotrowska Z, Niederst MJ, Karlovich CA, Wakelee HA, Neal JW, Mino-Kenudson M, et al. Heterogeneity underlies the emergence of EGFRT790 wild-type clones following treatment of T790M-positive cancers with a third-generation EGFR inhibitor. Cancer Discov. 2015;5:713–22.PubMedPubMedCentralCrossRef Piotrowska Z, Niederst MJ, Karlovich CA, Wakelee HA, Neal JW, Mino-Kenudson M, et al. Heterogeneity underlies the emergence of EGFRT790 wild-type clones following treatment of T790M-positive cancers with a third-generation EGFR inhibitor. Cancer Discov. 2015;5:713–22.PubMedPubMedCentralCrossRef
41.
Zurück zum Zitat Nukaga S, Yasuda H, Tsuchihara K, Hamamoto J, Masuzawa K, Kawada I, 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;77:2078–89.PubMedCrossRef Nukaga S, Yasuda H, Tsuchihara K, Hamamoto J, Masuzawa K, Kawada I, 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;77:2078–89.PubMedCrossRef
42.
Zurück zum Zitat Huang C, Zou Q, Liu H, Qiu B, Li Q, Lin Y, et al. Management of non-small cell lung cancer patients with MET exon 14 skipping mutations. Curr Treat Opt Oncol. 2020;21:33.CrossRef Huang C, Zou Q, Liu H, Qiu B, Li Q, Lin Y, et al. Management of non-small cell lung cancer patients with MET exon 14 skipping mutations. Curr Treat Opt Oncol. 2020;21:33.CrossRef
43.
Zurück zum Zitat Mueller KL, Madden JM, Zoratti GL, Kuperwasser C, List K, Boerner JL. Fibroblast-secreted hepatocyte growth factor mediates epidermal growth factor receptor tyrosine kinase inhibitor resistance in triple-negative breast cancers through paracrine activation of MET. Breast Cancer Res. 2012;14:R104.PubMedPubMedCentralCrossRef Mueller KL, Madden JM, Zoratti GL, Kuperwasser C, List K, Boerner JL. Fibroblast-secreted hepatocyte growth factor mediates epidermal growth factor receptor tyrosine kinase inhibitor resistance in triple-negative breast cancers through paracrine activation of MET. Breast Cancer Res. 2012;14:R104.PubMedPubMedCentralCrossRef
44.
Zurück zum Zitat Hsu CC, Liao BC, Liao WY, Markovets A, Stetson D, Thress K, et al. Exon 16-skipping HER2 as a novel mechanism of osimertinib resistance in EGFR L858R/T790M-positive non-small cell lung cancer. J Thorac Oncol. 2020;15:50–61.PubMedCrossRef Hsu CC, Liao BC, Liao WY, Markovets A, Stetson D, Thress K, et al. Exon 16-skipping HER2 as a novel mechanism of osimertinib resistance in EGFR L858R/T790M-positive non-small cell lung cancer. J Thorac Oncol. 2020;15:50–61.PubMedCrossRef
45.
Zurück zum Zitat Ou S-HI, Madison R, Robichaux JP, Ross JS, Miller VA, Ali SM, et al. Characterization of 648 non-small cell lung cancer (NSCLC) cases with 28 unique HER2 exon 20 insertions. J Clin Oncol. 2019;37:9063–63.CrossRef Ou S-HI, Madison R, Robichaux JP, Ross JS, Miller VA, Ali SM, et al. Characterization of 648 non-small cell lung cancer (NSCLC) cases with 28 unique HER2 exon 20 insertions. J Clin Oncol. 2019;37:9063–63.CrossRef
46.
Zurück zum Zitat Gao G, Li X, Wang Q, Zhang Y, Chen J, Shu Y, et al. Single-arm, phase ii study of pyrotinib in advanced non-small cell lung cancer (NSCLC) patients with HER2 exon 20 mutation. J Clin Oncol. 2019;37:9089–189.CrossRef Gao G, Li X, Wang Q, Zhang Y, Chen J, Shu Y, et al. Single-arm, phase ii study of pyrotinib in advanced non-small cell lung cancer (NSCLC) patients with HER2 exon 20 mutation. J Clin Oncol. 2019;37:9089–189.CrossRef
47.
48.
Zurück zum Zitat Zhu J, Yang Q, Xu W. Iterative upgrading of small molecular tyrosine kinase inhibitors for EGFR mutation in NSCLC: necessity and perspective. Pharmaceutics. 2021;13:1500.PubMedPubMedCentralCrossRef Zhu J, Yang Q, Xu W. Iterative upgrading of small molecular tyrosine kinase inhibitors for EGFR mutation in NSCLC: necessity and perspective. Pharmaceutics. 2021;13:1500.PubMedPubMedCentralCrossRef
49.
Zurück zum Zitat Taniguchi H, Yamada T, Wang R, Tanimura K, Adachi Y, Nishiyama A, et al. AXL confers intrinsic resistance to osimertinib and advances the emergence of tolerant cells. Nat Commun. 2019;10:259.PubMedPubMedCentralCrossRef Taniguchi H, Yamada T, Wang R, Tanimura K, Adachi Y, Nishiyama A, et al. AXL confers intrinsic resistance to osimertinib and advances the emergence of tolerant cells. Nat Commun. 2019;10:259.PubMedPubMedCentralCrossRef
50.
Zurück zum Zitat Yano S, Yamada T, Takeuchi S, Tachibana K, Minami Y, Yatabe Y, et al. Hepatocyte growth factor expression in EGFR mutant lung cancer with intrinsic and acquired resistance to tyrosine kinase inhibitors in a Japanese cohort. J Thorac Oncol. 2011;6:2011–7.PubMedCrossRef Yano S, Yamada T, Takeuchi S, Tachibana K, Minami Y, Yatabe Y, et al. Hepatocyte growth factor expression in EGFR mutant lung cancer with intrinsic and acquired resistance to tyrosine kinase inhibitors in a Japanese cohort. J Thorac Oncol. 2011;6:2011–7.PubMedCrossRef
51.
Zurück zum Zitat Kim TM, Song A, Kim DW, Kim S, Ahn YO, Keam B, et al. Mechanisms of acquired resistance to AZD9291 a mutation-selective, irreversible EGFR inhibitor. J Thorac Oncol. 2015;10:1736–44.PubMedCrossRef Kim TM, Song A, Kim DW, Kim S, Ahn YO, Keam B, et al. Mechanisms of acquired resistance to AZD9291 a mutation-selective, irreversible EGFR inhibitor. J Thorac Oncol. 2015;10:1736–44.PubMedCrossRef
52.
Zurück zum Zitat Papadimitrakopoulou VA, Wu YL, Han JY, Ahn MJ, Ramalingam SS, John T, et al. Analysis of resistance mechanisms to osimertinib in patients with EGFR T790M advanced NSCLC from the AURA3 study. Ann Oncol. 2018;29:741–841.CrossRef Papadimitrakopoulou VA, Wu YL, Han JY, Ahn MJ, Ramalingam SS, John T, et al. Analysis of resistance mechanisms to osimertinib in patients with EGFR T790M advanced NSCLC from the AURA3 study. Ann Oncol. 2018;29:741–841.CrossRef
53.
Zurück zum Zitat Tanaka H, Sakagami H, Kaneko N, Konagai S, Yamamoto H, Matsuya T, et al. Mutant-selective irreversible EGFR inhibitor, naquotinib, inhibits tumor growth in NSCLC models with EGFR-activating mutations, T790M mutation, and AXL overexpression. Mol Cancer Ther. 2019;18:1366–73.PubMedCrossRef Tanaka H, Sakagami H, Kaneko N, Konagai S, Yamamoto H, Matsuya T, et al. Mutant-selective irreversible EGFR inhibitor, naquotinib, inhibits tumor growth in NSCLC models with EGFR-activating mutations, T790M mutation, and AXL overexpression. Mol Cancer Ther. 2019;18:1366–73.PubMedCrossRef
54.
Zurück zum Zitat Park JH, Choi YJ, Kim SY, Lee JE, Sung KJ, Park S, et al. Activation of the IGF1R pathway potentially mediates acquired resistance to mutant-selective 3rd-generation EGF receptor tyrosine kinase inhibitors in advanced non-small cell lung cancer. Oncotarget. 2016;7:22005–15.PubMedPubMedCentralCrossRef Park JH, Choi YJ, Kim SY, Lee JE, Sung KJ, Park S, et al. Activation of the IGF1R pathway potentially mediates acquired resistance to mutant-selective 3rd-generation EGF receptor tyrosine kinase inhibitors in advanced non-small cell lung cancer. Oncotarget. 2016;7:22005–15.PubMedPubMedCentralCrossRef
55.
Zurück zum Zitat Carmena M, Earnshaw WC. The cellular geography of aurora kinases. Nat Rev Mol Cell Biol. 2003;4:842–54.PubMedCrossRef Carmena M, Earnshaw WC. The cellular geography of aurora kinases. Nat Rev Mol Cell Biol. 2003;4:842–54.PubMedCrossRef
56.
Zurück zum Zitat Cheetham GMT, Knegtel RMA, Coll JT, Renwick SB, Swenson L, Weber P, et al. Crystal structure of aurora-2, an oncogenic serine/threonine kinase*. J Biol Chem. 2002;277:42419–22.PubMedCrossRef Cheetham GMT, Knegtel RMA, Coll JT, Renwick SB, Swenson L, Weber P, et al. Crystal structure of aurora-2, an oncogenic serine/threonine kinase*. J Biol Chem. 2002;277:42419–22.PubMedCrossRef
57.
Zurück zum Zitat Pradhan T, Gupta O, Singh G, Monga V. Aurora kinase inhibitors as potential anticancer agents: recent advances. Eur J Med Chem. 2021;221:113495.PubMedCrossRef Pradhan T, Gupta O, Singh G, Monga V. Aurora kinase inhibitors as potential anticancer agents: recent advances. Eur J Med Chem. 2021;221:113495.PubMedCrossRef
58.
Zurück zum Zitat Falchook GS, Bastida CC, Kurzrock R. Aurora kinase inhibitors in oncology clinical trials: current state of the progress. Semin Oncol. 2015;42:832–48.PubMedCrossRef Falchook GS, Bastida CC, Kurzrock R. Aurora kinase inhibitors in oncology clinical trials: current state of the progress. Semin Oncol. 2015;42:832–48.PubMedCrossRef
59.
Zurück zum Zitat Hu L, Fan M, Shi S, Song X, Wang F, He H, et al. Dual target inhibitors based on EGFR: promising anticancer agents for the treatment of cancers (2017-). Eur J Med Chem. 2022;227:113963.PubMedCrossRef Hu L, Fan M, Shi S, Song X, Wang F, He H, et al. Dual target inhibitors based on EGFR: promising anticancer agents for the treatment of cancers (2017-). Eur J Med Chem. 2022;227:113963.PubMedCrossRef
60.
Zurück zum Zitat Tanaka K, Yu HA, Yang S, Han S, Selcuklu SD, Kim K, et al. Targeting aurora b kinase prevents and overcomes resistance to EGFR inhibitors in lung cancer by enhancing BIM- and PUMA-mediated apoptosis. Cancer Cell. 2021;39(1245–61):e6. Tanaka K, Yu HA, Yang S, Han S, Selcuklu SD, Kim K, et al. Targeting aurora b kinase prevents and overcomes resistance to EGFR inhibitors in lung cancer by enhancing BIM- and PUMA-mediated apoptosis. Cancer Cell. 2021;39(1245–61):e6.
61.
Zurück zum Zitat Mao C, Qiu LX, Liao RY, Du FB, Ding H, Yang WC, et al. KRAS mutations and resistance to EGFR-TKIs treatment in patients with non-small cell lung cancer: a meta-analysis of 22 studies. Lung Cancer. 2010;69:272–8.PubMedCrossRef Mao C, Qiu LX, Liao RY, Du FB, Ding H, Yang WC, et al. KRAS mutations and resistance to EGFR-TKIs treatment in patients with non-small cell lung cancer: a meta-analysis of 22 studies. Lung Cancer. 2010;69:272–8.PubMedCrossRef
62.
Zurück zum Zitat Sunaga N, Shames DS, Girard L, Peyton M, Larsen JE, Imai H, et al. Knockdown of oncogenic KRAS in non-small cell lung cancers suppresses tumor growth and sensitizes tumor cells to targeted therapy. Mol Cancer Ther. 2011;10:336–46.PubMedPubMedCentralCrossRef Sunaga N, Shames DS, Girard L, Peyton M, Larsen JE, Imai H, et al. Knockdown of oncogenic KRAS in non-small cell lung cancers suppresses tumor growth and sensitizes tumor cells to targeted therapy. Mol Cancer Ther. 2011;10:336–46.PubMedPubMedCentralCrossRef
63.
Zurück zum Zitat Leonetti A, Facchinetti F, Rossi G, Minari R, Conti A, Friboulet L, et al. Braf in non-small cell lung cancer (NSCLC): pickaxing another brick in the wall. Cancer Treat Rev. 2018;66:82–94.PubMedCrossRef Leonetti A, Facchinetti F, Rossi G, Minari R, Conti A, Friboulet L, et al. Braf in non-small cell lung cancer (NSCLC): pickaxing another brick in the wall. Cancer Treat Rev. 2018;66:82–94.PubMedCrossRef
64.
Zurück zum Zitat Ho CC, Liao WY, Lin CA, Shih JY, Yu CJ, Yang JC. Acquired BRAF V600E mutation as resistant mechanism after treatment with osimertinib. J Thorac Oncol. 2017;12:567–72.PubMedCrossRef Ho CC, Liao WY, Lin CA, Shih JY, Yu CJ, Yang JC. Acquired BRAF V600E mutation as resistant mechanism after treatment with osimertinib. J Thorac Oncol. 2017;12:567–72.PubMedCrossRef
65.
Zurück zum Zitat Ramalingam SS, Cheng Y, Zhou C, Ohe Y, Imamura F, Cho BC, et al. Mechanisms of acquired resistance to first-line osimertinib: preliminary data from the phase iii flaura study. Ann Oncol. 2018;29:viii740.CrossRef Ramalingam SS, Cheng Y, Zhou C, Ohe Y, Imamura F, Cho BC, et al. Mechanisms of acquired resistance to first-line osimertinib: preliminary data from the phase iii flaura study. Ann Oncol. 2018;29:viii740.CrossRef
66.
Zurück zum Zitat Zhao M, Gao FH, Wang JY, Liu F, Yuan HH, Zhang WY, et al. JAK2/STAT3 signaling pathway activation mediates tumor angiogenesis by upregulation of VEGF and bFGF in non-small-cell lung cancer. Lung Cancer. 2011;73:366–74.PubMedCrossRef Zhao M, Gao FH, Wang JY, Liu F, Yuan HH, Zhang WY, et al. JAK2/STAT3 signaling pathway activation mediates tumor angiogenesis by upregulation of VEGF and bFGF in non-small-cell lung cancer. Lung Cancer. 2011;73:366–74.PubMedCrossRef
67.
Zurück zum Zitat Chaib I, Karachaliou N, Pilotto S, Codony Servat J, Cai X, Li X, et al. Co-activation of STAT3 and YES-associated protein 1 (YAP1) pathway in EGFR-mutant NSCLC. J Natl Cancer Inst. 2017;109. Chaib I, Karachaliou N, Pilotto S, Codony Servat J, Cai X, Li X, et al. Co-activation of STAT3 and YES-associated protein 1 (YAP1) pathway in EGFR-mutant NSCLC. J Natl Cancer Inst. 2017;109.
68.
Zurück zum Zitat Soria J-C, Lee H-Y, Lee JI, Wang L, Issa J-P, Kemp BL, et al. Lack of PTEN expression in non-small cell lung cancer could be related to promoter methylation. Clin Cancer Res. 2002;8:1178–84.PubMed Soria J-C, Lee H-Y, Lee JI, Wang L, Issa J-P, Kemp BL, et al. Lack of PTEN expression in non-small cell lung cancer could be related to promoter methylation. Clin Cancer Res. 2002;8:1178–84.PubMed
69.
Zurück zum Zitat Zhang T, Qu R, Chan S, Lai M, Tong L, Feng F, et al. Discovery of a novel third-generation EGFR inhibitor and identification of a potential combination strategy to overcome resistance. Mol Cancer. 2020;19:90.PubMedPubMedCentralCrossRef Zhang T, Qu R, Chan S, Lai M, Tong L, Feng F, et al. Discovery of a novel third-generation EGFR inhibitor and identification of a potential combination strategy to overcome resistance. Mol Cancer. 2020;19:90.PubMedPubMedCentralCrossRef
70.
Zurück zum Zitat Zhu L, Chen Z, Zang H, Fan S, Gu J, Zhang G, et al. Targeting c-Myc to overcome acquired resistance of EGFR mutant NSCLC cells to the third-generation EGFR tyrosine kinase inhibitor, osimertinib. Cancer Res. 2021;81:4822–34.PubMedPubMedCentralCrossRef Zhu L, Chen Z, Zang H, Fan S, Gu J, Zhang G, et al. Targeting c-Myc to overcome acquired resistance of EGFR mutant NSCLC cells to the third-generation EGFR tyrosine kinase inhibitor, osimertinib. Cancer Res. 2021;81:4822–34.PubMedPubMedCentralCrossRef
71.
Zurück zum Zitat Weng CH, Chen LY, Lin YC, Shih JY, Lin YC, Tseng RY, et al. Epithelial–mesenchymal transition (emtEMT) beyond EGFR mutations per se is a common mechanism for acquired resistance to EGFR TKI. Oncogene. 2019;38:455–68.PubMedCrossRef Weng CH, Chen LY, Lin YC, Shih JY, Lin YC, Tseng RY, et al. Epithelial–mesenchymal transition (emtEMT) beyond EGFR mutations per se is a common mechanism for acquired resistance to EGFR TKI. Oncogene. 2019;38:455–68.PubMedCrossRef
72.
Zurück zum Zitat Brabletz T, Kalluri R, Nieto MA, Weinberg RA. EMT in cancer. Nat Rev Cancer. 2018;18:128–34.PubMedCrossRef Brabletz T, Kalluri R, Nieto MA, Weinberg RA. EMT in cancer. Nat Rev Cancer. 2018;18:128–34.PubMedCrossRef
73.
Zurück zum Zitat Kong W, Yang H, He L, Zhao JJ, Coppola D, Dalton WS, et al. MicroRNA-155 is regulated by the transforming growth factor beta/Smad pathway and contributes to epithelial cell plasticity by targeting RhoA. Mol Cell Biol. 2008;28:6773–84.PubMedPubMedCentralCrossRef Kong W, Yang H, He L, Zhao JJ, Coppola D, Dalton WS, et al. MicroRNA-155 is regulated by the transforming growth factor beta/Smad pathway and contributes to epithelial cell plasticity by targeting RhoA. Mol Cell Biol. 2008;28:6773–84.PubMedPubMedCentralCrossRef
74.
Zurück zum Zitat Liu CH, Huang Q, Jin ZY, Zhu CL, Liu Z, Wang C. miR-21 and KLF4 jointly augment epithelialmesenchymal transition via the Akt/ERK1/2 pathway. Int J Oncol. 2017;50:1109–15.PubMedPubMedCentralCrossRef Liu CH, Huang Q, Jin ZY, Zhu CL, Liu Z, Wang C. miR-21 and KLF4 jointly augment epithelialmesenchymal transition via the Akt/ERK1/2 pathway. Int J Oncol. 2017;50:1109–15.PubMedPubMedCentralCrossRef
75.
Zurück zum Zitat Han Z, Zhou X, Li S, Qin Y, Chen Y, Liu H. Inhibition of miR-23a increases the sensitivity of lung cancer stem cells to erlotinib through PTEN/PI3K/Akt pathway. Oncol Rep. 2017;38:3064–70.PubMedCrossRef Han Z, Zhou X, Li S, Qin Y, Chen Y, Liu H. Inhibition of miR-23a increases the sensitivity of lung cancer stem cells to erlotinib through PTEN/PI3K/Akt pathway. Oncol Rep. 2017;38:3064–70.PubMedCrossRef
76.
Zurück zum Zitat Shen H, Zhu F, Liu J, Xu T, Pei D, Wang R, et al. Alteration in Mir-21/PTEN expression modulates gefitinib resistance in non-small cell lung cancer. PLoS ONE. 2014;9:e103305.PubMedPubMedCentralCrossRef Shen H, Zhu F, Liu J, Xu T, Pei D, Wang R, et al. Alteration in Mir-21/PTEN expression modulates gefitinib resistance in non-small cell lung cancer. PLoS ONE. 2014;9:e103305.PubMedPubMedCentralCrossRef
77.
78.
Zurück zum Zitat Del Re M, Arrigoni E, Restante G, Passaro A, Rofi E, Crucitta S, et al. Concise review: resistance to tyrosine kinase inhibitors in non-small cell lung cancer: the role of cancer stem cells. Stem Cells. 2018;36:633–40.PubMedCrossRef Del Re M, Arrigoni E, Restante G, Passaro A, Rofi E, Crucitta S, et al. Concise review: resistance to tyrosine kinase inhibitors in non-small cell lung cancer: the role of cancer stem cells. Stem Cells. 2018;36:633–40.PubMedCrossRef
79.
Zurück zum Zitat Papadimitrakopoulou VA, Wu YL, Han JY, Ahn MJ, Ramalingam SS, John T, et al. Analysis of resistance mechanisms to osimertinib in patients with EGFR T790M advanced NSCLC from the AURA3 study. Ann Oncol. 2018;29:viii741.CrossRef Papadimitrakopoulou VA, Wu YL, Han JY, Ahn MJ, Ramalingam SS, John T, et al. Analysis of resistance mechanisms to osimertinib in patients with EGFR T790M advanced NSCLC from the AURA3 study. Ann Oncol. 2018;29:viii741.CrossRef
80.
Zurück zum Zitat Dorantes-Heredia R, Ruiz-Morales JM, Cano-Garcia F. Histopathological transformation to small-cell lung carcinoma in non-small cell lung carcinoma tumors. Transl Lung Cancer Res. 2016;5:401–12.PubMedPubMedCentralCrossRef Dorantes-Heredia R, Ruiz-Morales JM, Cano-Garcia F. Histopathological transformation to small-cell lung carcinoma in non-small cell lung carcinoma tumors. Transl Lung Cancer Res. 2016;5:401–12.PubMedPubMedCentralCrossRef
81.
Zurück zum Zitat Sequist LV, Waltman BA, Dias-Santagata D, Digumarthy S, Turke AB, Fidias P, et al. Genotypic and histological evolution of lung cancers acquiring resistance to EGFR inhibitors. Sci Transl Med. 2011;3:75ra26.PubMedPubMedCentralCrossRef Sequist LV, Waltman BA, Dias-Santagata D, Digumarthy S, Turke AB, Fidias P, et al. Genotypic and histological evolution of lung cancers acquiring resistance to EGFR inhibitors. Sci Transl Med. 2011;3:75ra26.PubMedPubMedCentralCrossRef
82.
Zurück zum Zitat Oser MG, Niederst MJ, Sequist LV, Engelman JA. Transformation from non-small-cell lung cancer to small-cell lung cancer: Molecular drivers and cells of origin. Lancet Oncol. 2015;16:e165-172.PubMedPubMedCentralCrossRef Oser MG, Niederst MJ, Sequist LV, Engelman JA. Transformation from non-small-cell lung cancer to small-cell lung cancer: Molecular drivers and cells of origin. Lancet Oncol. 2015;16:e165-172.PubMedPubMedCentralCrossRef
83.
Zurück zum Zitat Norkowski E, Ghigna MR, Lacroix L, Le Chevalier T, Fadel E, Dartevelle P, et al. Small-cell carcinoma in the setting of pulmonary adenocarcinoma: new insights in the era of molecular pathology. J Thorac Oncol. 2013;8:1265–71.PubMedCrossRef Norkowski E, Ghigna MR, Lacroix L, Le Chevalier T, Fadel E, Dartevelle P, et al. Small-cell carcinoma in the setting of pulmonary adenocarcinoma: new insights in the era of molecular pathology. J Thorac Oncol. 2013;8:1265–71.PubMedCrossRef
85.
Zurück zum Zitat Schoenfeld AJ, Chan JM, Kubota D, Sato H, Rizvi H, Daneshbod Y, 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;26:2654–63.PubMedPubMedCentralCrossRef Schoenfeld AJ, Chan JM, Kubota D, Sato H, Rizvi H, Daneshbod Y, 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;26:2654–63.PubMedPubMedCentralCrossRef
86.
Zurück zum Zitat Garassino MC, Cho BC, Kim JH, Mazieres J, Vansteenkiste J, Lena H, et al. Durvalumab as third-line or later treatment for advanced non-small-cell lung cancer (ATLANTIC): an open-label, single-arm, phase 2 study. Lancet Oncol. 2018;19:521–36.PubMedPubMedCentralCrossRef Garassino MC, Cho BC, Kim JH, Mazieres J, Vansteenkiste J, Lena H, et al. Durvalumab as third-line or later treatment for advanced non-small-cell lung cancer (ATLANTIC): an open-label, single-arm, phase 2 study. Lancet Oncol. 2018;19:521–36.PubMedPubMedCentralCrossRef
87.
Zurück zum Zitat Gainor JF, Shaw AT, Sequist LV, Fu X, Azzoli CG, Piotrowska Z, et al. EGFR mutations and ALK rearrangements are associated with low response rates to PD-1 pathway blockade in non-small cell lung cancer: a retrospective analysis. Clin Cancer Res. 2016;22:4585–93.PubMedPubMedCentralCrossRef Gainor JF, Shaw AT, Sequist LV, Fu X, Azzoli CG, Piotrowska Z, et al. EGFR mutations and ALK rearrangements are associated with low response rates to PD-1 pathway blockade in non-small cell lung cancer: a retrospective analysis. Clin Cancer Res. 2016;22:4585–93.PubMedPubMedCentralCrossRef
88.
Zurück zum Zitat Haratani K, Hayashi H, Tanaka T, Kaneda H, Togashi Y, Sakai K, et al. Tumor immune microenvironment and nivolumab efficacy in EGFR mutation-positive non-small-cell lung cancer based on T790M status after disease progression during EGFR-TKI treatment. Ann Oncol. 2017;28:1532–9.PubMedCrossRef Haratani K, Hayashi H, Tanaka T, Kaneda H, Togashi Y, Sakai K, et al. Tumor immune microenvironment and nivolumab efficacy in EGFR mutation-positive non-small-cell lung cancer based on T790M status after disease progression during EGFR-TKI treatment. Ann Oncol. 2017;28:1532–9.PubMedCrossRef
89.
Zurück zum Zitat Zheng Y, Hao S, Xiang C, Han Y, Shang Y, Zhen Q, et al. The correlation between SPP1 and immune escape of EGFR mutant lung adenocarcinoma was explored by bioinformatics analysis. Front Oncol. 2021;11:592854.PubMedPubMedCentralCrossRef Zheng Y, Hao S, Xiang C, Han Y, Shang Y, Zhen Q, et al. The correlation between SPP1 and immune escape of EGFR mutant lung adenocarcinoma was explored by bioinformatics analysis. Front Oncol. 2021;11:592854.PubMedPubMedCentralCrossRef
90.
Zurück zum Zitat Peng S, Wang R, Zhang X, Ma Y, Zhong L, Li K, et al. EGFR-TKI resistance promotes immune escape in lung cancer via increased PD-L1 expression. Mol Cancer. 2019;18:165.PubMedPubMedCentralCrossRef Peng S, Wang R, Zhang X, Ma Y, Zhong L, Li K, et al. EGFR-TKI resistance promotes immune escape in lung cancer via increased PD-L1 expression. Mol Cancer. 2019;18:165.PubMedPubMedCentralCrossRef
91.
Zurück zum Zitat Yu HA, Tian SK, Drilon AE, Borsu L, Riely GJ, Arcila ME, et al. Acquired resistance of EGFR-mutant lung cancer to a T790M-specific EGFR inhibitor: emergence of a third mutation (C797S) in the EGFR tyrosine kinase domain. JAMA Oncol. 2015;1:982–4.PubMedPubMedCentralCrossRef Yu HA, Tian SK, Drilon AE, Borsu L, Riely GJ, Arcila ME, et al. Acquired resistance of EGFR-mutant lung cancer to a T790M-specific EGFR inhibitor: emergence of a third mutation (C797S) in the EGFR tyrosine kinase domain. JAMA Oncol. 2015;1:982–4.PubMedPubMedCentralCrossRef
92.
Zurück zum Zitat Wang Z, Yang JJ, Huang J, Ye JY, Zhang XC, Tu HY, et al. 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;12:1723–7.PubMedCrossRef Wang Z, Yang JJ, Huang J, Ye JY, Zhang XC, Tu HY, et al. 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;12:1723–7.PubMedCrossRef
93.
Zurück zum Zitat Niederst MJ, Hu H, Mulvey HE, Lockerman EL, Garcia AR, Piotrowska Z, et al. 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;21:3924–33.PubMedPubMedCentralCrossRef Niederst MJ, Hu H, Mulvey HE, Lockerman EL, Garcia AR, Piotrowska Z, et al. 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;21:3924–33.PubMedPubMedCentralCrossRef
94.
95.
Zurück zum Zitat Engel J, Richters A, Getlik M, Tomassi S, Keul M, Termathe M, et al. Targeting drug resistance in EGFR with covalent inhibitors: a structure-based design approach. J Med Chem. 2015;58:6844–63.PubMedCrossRef Engel J, Richters A, Getlik M, Tomassi S, Keul M, Termathe M, et al. Targeting drug resistance in EGFR with covalent inhibitors: a structure-based design approach. J Med Chem. 2015;58:6844–63.PubMedCrossRef
96.
Zurück zum Zitat Jia Y, Yun CH, Park E, Ercan D, Manuia M, Juarez J, et al. Overcoming EGFR(T790M) and EGFR(C797S) resistance with mutant-selective allosteric inhibitors. Nature. 2016;534:129–32.PubMedPubMedCentralCrossRef Jia Y, Yun CH, Park E, Ercan D, Manuia M, Juarez J, et al. Overcoming EGFR(T790M) and EGFR(C797S) resistance with mutant-selective allosteric inhibitors. Nature. 2016;534:129–32.PubMedPubMedCentralCrossRef
97.
Zurück zum Zitat Lee S, Kim J, Duggirala KB, Go A, Shin I, Cho BC, et al. Allosteric inhibitor TREA-0236 containing non-hydrolysable quinazoline-4-one for EGFR T790M/C797S mutants inhibition. Bull Korean Chem Soc. 2018;39:895–8.CrossRef Lee S, Kim J, Duggirala KB, Go A, Shin I, Cho BC, et al. Allosteric inhibitor TREA-0236 containing non-hydrolysable quinazoline-4-one for EGFR T790M/C797S mutants inhibition. Bull Korean Chem Soc. 2018;39:895–8.CrossRef
98.
Zurück zum Zitat To C, Jang J, Chen T, Park E, Mushajiang M, De Clercq DJH, et al. Single and dual targeting of mutant EGFR with an allosteric inhibitor. Cancer Discov. 2019;9:926–43.PubMedPubMedCentralCrossRef To C, Jang J, Chen T, Park E, Mushajiang M, De Clercq DJH, et al. Single and dual targeting of mutant EGFR with an allosteric inhibitor. Cancer Discov. 2019;9:926–43.PubMedPubMedCentralCrossRef
99.
Zurück zum Zitat Maity S, Pai KSR, Nayak Y. Advances in targeting EGFR allosteric site as anti-NSCLC therapy to overcome the drug resistance. Pharmacol Rep. 2020;72:799–813.PubMedPubMedCentralCrossRef Maity S, Pai KSR, Nayak Y. Advances in targeting EGFR allosteric site as anti-NSCLC therapy to overcome the drug resistance. Pharmacol Rep. 2020;72:799–813.PubMedPubMedCentralCrossRef
100.
Zurück zum Zitat De Clercq DJH, Heppner DE, To C, Jang J, Park E, Yun C-H, et al. Discovery and optimization of dibenzodiazepinones as allosteric mutant-selective EGFR inhibitors. ACS Med Chem Lett. 2019;10:1549–53.PubMedPubMedCentralCrossRef De Clercq DJH, Heppner DE, To C, Jang J, Park E, Yun C-H, et al. Discovery and optimization of dibenzodiazepinones as allosteric mutant-selective EGFR inhibitors. ACS Med Chem Lett. 2019;10:1549–53.PubMedPubMedCentralCrossRef
101.
Zurück zum Zitat Duplessis M, Goergler A, Jaeschke G, Kocer B, Kuhn B, Lazarski K, et al. COMPOUNDS. Publication number: 20210079005, March 18, 2021. Duplessis M, Goergler A, Jaeschke G, Kocer B, Kuhn B, Lazarski K, et al. COMPOUNDS. Publication number: 20210079005, March 18, 2021.
102.
Zurück zum Zitat Lu X, Zhang T, Zhu SJ, Xun Q, Tong L, Hu X, et al. Discovery of JND3229 as a new EGFR(C797S) mutant inhibitor with in vivo monodrug efficacy. ACS Med Chem Lett. 2018;9:1123–7.PubMedPubMedCentralCrossRef Lu X, Zhang T, Zhu SJ, Xun Q, Tong L, Hu X, et al. Discovery of JND3229 as a new EGFR(C797S) mutant inhibitor with in vivo monodrug efficacy. ACS Med Chem Lett. 2018;9:1123–7.PubMedPubMedCentralCrossRef
103.
Zurück zum Zitat Engel J, Becker C, Lategahn J, Keul M, Ketzer J, Muhlenberg T, et al. Insight into the inhibition of drug-resistant mutants of the receptor tyrosine kinase EGFR. Angew Chem Int Ed Engl. 2016;55:10909–12.PubMedCrossRef Engel J, Becker C, Lategahn J, Keul M, Ketzer J, Muhlenberg T, et al. Insight into the inhibition of drug-resistant mutants of the receptor tyrosine kinase EGFR. Angew Chem Int Ed Engl. 2016;55:10909–12.PubMedCrossRef
104.
Zurück zum Zitat Gunther M, Lategahn J, Juchum M, Doring E, Keul M, Engel J, et al. Trisubstituted pyridinylimidazoles as potent inhibitors of the clinically resistant L858R/T790M/C797S EGFR mutant: targeting of both hydrophobic regions and the phosphate binding site. J Med Chem. 2017;60:5613–37.PubMedCrossRef Gunther M, Lategahn J, Juchum M, Doring E, Keul M, Engel J, et al. Trisubstituted pyridinylimidazoles as potent inhibitors of the clinically resistant L858R/T790M/C797S EGFR mutant: targeting of both hydrophobic regions and the phosphate binding site. J Med Chem. 2017;60:5613–37.PubMedCrossRef
105.
Zurück zum Zitat Park H, Jung HY, Mah S, Hong S. Discovery of EGF receptor inhibitors that are selective for the D746–750/T790M/C797S mutant through structure-based de novo design. Angew Chem Int Ed Engl. 2017;56:7634–8.PubMedCrossRef Park H, Jung HY, Mah S, Hong S. Discovery of EGF receptor inhibitors that are selective for the D746–750/T790M/C797S mutant through structure-based de novo design. Angew Chem Int Ed Engl. 2017;56:7634–8.PubMedCrossRef
106.
Zurück zum Zitat Zhang M, Wang Y, Wang J, Liu Z, Shi J, Li M, et al. Design, synthesis and biological evaluation of the quinazoline derivatives as L858R/T790M/C797S triple mutant epidermal growth factor receptor tyrosine kinase inhibitors. Chem Pharm Bull (Tokyo). 2020;68:971–80.CrossRef Zhang M, Wang Y, Wang J, Liu Z, Shi J, Li M, et al. Design, synthesis and biological evaluation of the quinazoline derivatives as L858R/T790M/C797S triple mutant epidermal growth factor receptor tyrosine kinase inhibitors. Chem Pharm Bull (Tokyo). 2020;68:971–80.CrossRef
107.
Zurück zum Zitat Shen J, Zhang T, Zhu SJ, Sun M, Tong L, Lai M, et al. Structure-based design of 5-methylpyrimidopyridone derivatives as new wild-type sparing inhibitors of the epidermal growth factor receptor triple mutant (EGFR(L858R/T790M/C797S). J Med Chem. 2019;62:7302–8.PubMedCrossRef Shen J, Zhang T, Zhu SJ, Sun M, Tong L, Lai M, et al. Structure-based design of 5-methylpyrimidopyridone derivatives as new wild-type sparing inhibitors of the epidermal growth factor receptor triple mutant (EGFR(L858R/T790M/C797S). J Med Chem. 2019;62:7302–8.PubMedCrossRef
108.
Zurück zum Zitat Zhang H, Wang J, Shen Y, Wang HY, Duan WM, Zhao HY, et al. Discovery of 2,4,6-trisubstitued pyrido[3,4-d]pyrimidine derivatives as new EGFR-TKIs. Eur J Med Chem. 2018;148:221–37.PubMedCrossRef Zhang H, Wang J, Shen Y, Wang HY, Duan WM, Zhao HY, et al. Discovery of 2,4,6-trisubstitued pyrido[3,4-d]pyrimidine derivatives as new EGFR-TKIs. Eur J Med Chem. 2018;148:221–37.PubMedCrossRef
109.
Zurück zum Zitat Hei YY, Shen Y, Wang J, Zhang H, Zhao HY, Xin M, et al. Synthesis and evaluation of 2,9-disubstituted 8-phenylthio/phenylsulfinyl-9H-purine as new EGFR inhibitors. Bioorg Med Chem. 2018;26:2173–85.PubMedCrossRef Hei YY, Shen Y, Wang J, Zhang H, Zhao HY, Xin M, et al. Synthesis and evaluation of 2,9-disubstituted 8-phenylthio/phenylsulfinyl-9H-purine as new EGFR inhibitors. Bioorg Med Chem. 2018;26:2173–85.PubMedCrossRef
110.
Zurück zum Zitat Lei H, Fan S, Zhang H, Liu YJ, Hei YY, Zhang JJ, et al. Discovery of novel 9-heterocyclyl substituted 9H-purines as L858R/T790M/C797S mutant EGFR tyrosine kinase inhibitors. Eur J Med Chem. 2020;186:111888.PubMedCrossRef Lei H, Fan S, Zhang H, Liu YJ, Hei YY, Zhang JJ, et al. Discovery of novel 9-heterocyclyl substituted 9H-purines as L858R/T790M/C797S mutant EGFR tyrosine kinase inhibitors. Eur J Med Chem. 2020;186:111888.PubMedCrossRef
111.
Zurück zum Zitat Lategahn J, Keul M, Klovekorn P, Tumbrink HL, Niggenaber J, Muller MP, et al. Inhibition of osimertinib-resistant epidermal growth factor receptor EGFR-T790M/C797S. Chem Sci. 2019;10:10789–801.PubMedPubMedCentralCrossRef Lategahn J, Keul M, Klovekorn P, Tumbrink HL, Niggenaber J, Muller MP, et al. Inhibition of osimertinib-resistant epidermal growth factor receptor EGFR-T790M/C797S. Chem Sci. 2019;10:10789–801.PubMedPubMedCentralCrossRef
112.
Zurück zum Zitat Hu X, Xun Q, Zhang T, Zhu S-J, Li Q, Tong L, et al. 2-Oxo-3,4-dihydropyrimido[4,5-d] pyrimidines as new reversible inhibitors of EGFR C797S (Cys797 to Ser797) mutant. Chin Chem Lett. 2020;31:1281–7.CrossRef Hu X, Xun Q, Zhang T, Zhu S-J, Li Q, Tong L, et al. 2-Oxo-3,4-dihydropyrimido[4,5-d] pyrimidines as new reversible inhibitors of EGFR C797S (Cys797 to Ser797) mutant. Chin Chem Lett. 2020;31:1281–7.CrossRef
113.
Zurück zum Zitat Su Z, Yang T, Wang J, Lai M, Tong L, Wumaier G, et al. Design, synthesis and biological evaluation of potent EGFR kinase inhibitors against 19D/T790M/C797S mutation. Bioorg Med Chem Lett. 2020;30:127327.PubMedCrossRef Su Z, Yang T, Wang J, Lai M, Tong L, Wumaier G, et al. Design, synthesis and biological evaluation of potent EGFR kinase inhibitors against 19D/T790M/C797S mutation. Bioorg Med Chem Lett. 2020;30:127327.PubMedCrossRef
114.
Zurück zum Zitat Lee Kwangho, SHIN Inji, CHOI Gildon, CHAE Chong Hak, Choe Hyeon Jeong, JUNG Myoung Eun, et al. N2,N4-diphenylpyrimidine-2,4-diamine derivative, method for preparing same, and pharmaceutical composition containing same as active ingredient for prevention or treatment of cancer. WO2018230934, 2018. Lee Kwangho, SHIN Inji, CHOI Gildon, CHAE Chong Hak, Choe Hyeon Jeong, JUNG Myoung Eun, et al. N2,N4-diphenylpyrimidine-2,4-diamine derivative, method for preparing same, and pharmaceutical composition containing same as active ingredient for prevention or treatment of cancer. WO2018230934, 2018.
115.
Zurück zum Zitat Wu L, Liu X, Ding CZ, Chen S, Hu L, Zhao L, et al. Spiro-aryl-phosphorus-oxygen compound as fourth generation of EGFR kinase inhibitor. WO 2018108064 A1, 2016. Wu L, Liu X, Ding CZ, Chen S, Hu L, Zhao L, et al. Spiro-aryl-phosphorus-oxygen compound as fourth generation of EGFR kinase inhibitor. WO 2018108064 A1, 2016.
116.
Zurück zum Zitat Iwao M, Fukuda T, Ishibashi F, Uehara Y, Nishiya N, Oku Y, et al. Fourth-generation EGFR tyrosine kinase inhibitor. CN 110461850 A, 2019. Iwao M, Fukuda T, Ishibashi F, Uehara Y, Nishiya N, Oku Y, et al. Fourth-generation EGFR tyrosine kinase inhibitor. CN 110461850 A, 2019.
117.
Zurück zum Zitat Boese D, Dahmann G, Engelhardt H, Petronczki M, Scharn D. New benzimidazole compounds and derivatives as EGFR inhibitors. WO 2019162323 A1, 2019. Boese D, Dahmann G, Engelhardt H, Petronczki M, Scharn D. New benzimidazole compounds and derivatives as EGFR inhibitors. WO 2019162323 A1, 2019.
118.
Zurück zum Zitat Ding K, Ding J, Shen J, Geng M, Lu X, Xie H, et al. Pyrimidopyridone or pyridopyridone compound and use thereof. WO 2019015593 A1, 2019. Ding K, Ding J, Shen J, Geng M, Lu X, Xie H, et al. Pyrimidopyridone or pyridopyridone compound and use thereof. WO 2019015593 A1, 2019.
119.
Zurück zum Zitat Ferlenghi F, Scalvini L, Vacondio F, Castelli R, Bozza N, Marseglia G, et al. A sulfonyl fluoride derivative inhibits EGFR(L858R/T790M/C797S) by covalent modification of the catalytic lysine. Eur J Med Chem. 2021;225:113786.PubMedCrossRef Ferlenghi F, Scalvini L, Vacondio F, Castelli R, Bozza N, Marseglia G, et al. A sulfonyl fluoride derivative inhibits EGFR(L858R/T790M/C797S) by covalent modification of the catalytic lysine. Eur J Med Chem. 2021;225:113786.PubMedCrossRef
120.
Zurück zum Zitat Morabito A, Piccirillo MC, Falasconi F, De Feo G, Del Giudice A, Bryce J, et al. Vandetanib (ZD6474), a dual inhibitor of vascular endothelial growth factor receptor (vEGFR) and epidermal growth factor receptor (EGFR) tyrosine kinases: Current status and future directions. Oncologist. 2009;14:378–90.PubMedCrossRef Morabito A, Piccirillo MC, Falasconi F, De Feo G, Del Giudice A, Bryce J, et al. Vandetanib (ZD6474), a dual inhibitor of vascular endothelial growth factor receptor (vEGFR) and epidermal growth factor receptor (EGFR) tyrosine kinases: Current status and future directions. Oncologist. 2009;14:378–90.PubMedCrossRef
121.
Zurück zum Zitat Li Q, Zhang T, Li S, Tong L, Li J, Su Z, et al. Discovery of potent and noncovalent reversible EGFR kinase inhibitors of EGFR(L858R/T790M/C797S). ACS Med Chem Lett. 2019;10:869–73.PubMedPubMedCentralCrossRef Li Q, Zhang T, Li S, Tong L, Li J, Su Z, et al. Discovery of potent and noncovalent reversible EGFR kinase inhibitors of EGFR(L858R/T790M/C797S). ACS Med Chem Lett. 2019;10:869–73.PubMedPubMedCentralCrossRef
122.
Zurück zum Zitat Wittlinger F, Heppner DE, To C, Gunther M, Shin BH, Rana JK, et al. Design of a “two-in-one” mutant-selective epidermal growth factor receptor inhibitor that spans the orthosteric and allosteric sites. J Med Chem. 2022;65:1370–83.PubMedCrossRef Wittlinger F, Heppner DE, To C, Gunther M, Shin BH, Rana JK, et al. Design of a “two-in-one” mutant-selective epidermal growth factor receptor inhibitor that spans the orthosteric and allosteric sites. J Med Chem. 2022;65:1370–83.PubMedCrossRef
123.
Zurück zum Zitat Yu HA, Arcila ME, Rekhtman N, Sima CS, Zakowski MF, Pao W, et al. Analysis of tumor specimens at the time of acquired resistance to EGFR-TKI therapy in 155 patients with EGFR-mutant lung cancers. Clin Cancer Res. 2013;19:2240–7.PubMedPubMedCentralCrossRef Yu HA, Arcila ME, Rekhtman N, Sima CS, Zakowski MF, Pao W, et al. Analysis of tumor specimens at the time of acquired resistance to EGFR-TKI therapy in 155 patients with EGFR-mutant lung cancers. Clin Cancer Res. 2013;19:2240–7.PubMedPubMedCentralCrossRef
124.
Zurück zum Zitat Noble ME, Endicott JA, Johnson LN. Protein kinase inhibitors: Insights into drug design from structure. Science. 2004;303:1800–5.PubMedCrossRef Noble ME, Endicott JA, Johnson LN. Protein kinase inhibitors: Insights into drug design from structure. Science. 2004;303:1800–5.PubMedCrossRef
125.
Zurück zum Zitat Bondeson DP, Mares A, Smith IE, Ko E, Campos S, Miah AH, et al. Catalytic in vivo protein knockdown by small-molecule protacs. Nat Chem Biol. 2015;11:611–7.PubMedPubMedCentralCrossRef Bondeson DP, Mares A, Smith IE, Ko E, Campos S, Miah AH, et al. Catalytic in vivo protein knockdown by small-molecule protacs. Nat Chem Biol. 2015;11:611–7.PubMedPubMedCentralCrossRef
126.
Zurück zum Zitat Winter GE, Buckley DL, Paulk J, Roberts JM, Souza A, Dhe-Paganon S, et al. Drug development. Phthalimide conjugation as a strategy for in vivo target protein degradation. Science. 2015;348:1376–81.PubMedPubMedCentralCrossRef Winter GE, Buckley DL, Paulk J, Roberts JM, Souza A, Dhe-Paganon S, et al. Drug development. Phthalimide conjugation as a strategy for in vivo target protein degradation. Science. 2015;348:1376–81.PubMedPubMedCentralCrossRef
127.
Zurück zum Zitat Scheepstra M, Hekking KFW, van Hijfte L, Folmer RHA. Bivalent ligands for protein degradation in drug discovery. Comput Struct Biotechnol J. 2019;17:160–76.PubMedPubMedCentralCrossRef Scheepstra M, Hekking KFW, van Hijfte L, Folmer RHA. Bivalent ligands for protein degradation in drug discovery. Comput Struct Biotechnol J. 2019;17:160–76.PubMedPubMedCentralCrossRef
128.
Zurück zum Zitat An S, Fu L. Small-molecule protacs: an emerging and promising approach for the development of targeted therapy drugs. EBioMedicine. 2018;36:553–62.PubMedPubMedCentralCrossRef An S, Fu L. Small-molecule protacs: an emerging and promising approach for the development of targeted therapy drugs. EBioMedicine. 2018;36:553–62.PubMedPubMedCentralCrossRef
129.
Zurück zum Zitat Chamberlain PP, Hamann LG. Development of targeted protein degradation therapeutics. Nat Chem Biol. 2019;15:937–44.PubMedCrossRef Chamberlain PP, Hamann LG. Development of targeted protein degradation therapeutics. Nat Chem Biol. 2019;15:937–44.PubMedCrossRef
130.
Zurück zum Zitat Churcher I. Protac-induced protein degradation in drug discovery: breaking the rules or just making new ones? J Med Chem. 2018;61:444–52.PubMedCrossRef Churcher I. Protac-induced protein degradation in drug discovery: breaking the rules or just making new ones? J Med Chem. 2018;61:444–52.PubMedCrossRef
132.
Zurück zum Zitat Toure M, Crews CM. Small-molecule protacs: new approaches to protein degradation. Angew Chem Int Ed Engl. 2016;55:1966–73.PubMedCrossRef Toure M, Crews CM. Small-molecule protacs: new approaches to protein degradation. Angew Chem Int Ed Engl. 2016;55:1966–73.PubMedCrossRef
133.
Zurück zum Zitat Jang J, To C, De Clercq DJH, Park E, Ponthier CM, Shin BH, et al. Mutant-selective allosteric EGFR degraders are effective against a broad range of drug-resistant mutations. Angew Chem Int Ed Engl. 2020;59:14481–9.PubMedPubMedCentralCrossRef Jang J, To C, De Clercq DJH, Park E, Ponthier CM, Shin BH, et al. Mutant-selective allosteric EGFR degraders are effective against a broad range of drug-resistant mutations. Angew Chem Int Ed Engl. 2020;59:14481–9.PubMedPubMedCentralCrossRef
134.
Zurück zum Zitat Zhao HY, Yang XY, Lei H, Xi XX, Lu SM, Zhang JJ, et al. Discovery of potent small molecule protacs targeting mutant EGFR. Eur J Med Chem. 2020;208:112781.PubMedCrossRef Zhao HY, Yang XY, Lei H, Xi XX, Lu SM, Zhang JJ, et al. Discovery of potent small molecule protacs targeting mutant EGFR. Eur J Med Chem. 2020;208:112781.PubMedCrossRef
135.
Zurück zum Zitat Qu X, Liu H, Song X, Sun N, Zhong H, Qiu X, et al. Effective degradation of EGFRL858R+T790M mutant proteins by CRBN-based PROTAC s through both proteosome and autophagy/lysosome degradation systems. Eur J Med Chem. 2021;218:113328.PubMedCrossRef Qu X, Liu H, Song X, Sun N, Zhong H, Qiu X, et al. Effective degradation of EGFRL858R+T790M mutant proteins by CRBN-based PROTAC s through both proteosome and autophagy/lysosome degradation systems. Eur J Med Chem. 2021;218:113328.PubMedCrossRef
136.
Zurück zum Zitat Zheng M, Huo J, Gu X, Wang Y, Wu C, Zhang Q, et al. Rational design and synthesis of novel dual PROTACS for simultaneous degradation of EGFR and PARP. J Med Chem. 2021;64:7839–52.PubMedCrossRef Zheng M, Huo J, Gu X, Wang Y, Wu C, Zhang Q, et al. Rational design and synthesis of novel dual PROTACS for simultaneous degradation of EGFR and PARP. J Med Chem. 2021;64:7839–52.PubMedCrossRef
137.
Zurück zum Zitat Kim JH, Nam B, Choi YJ, Kim SY, Lee JE, Sung KJ, et al. Enhanced glycolysis supports cell survival in EGFR-mutant lung adenocarcinoma by inhibiting autophagy-mediated EGFR degradation. Cancer Res. 2018;78:4482–96.PubMedCrossRef Kim JH, Nam B, Choi YJ, Kim SY, Lee JE, Sung KJ, et al. Enhanced glycolysis supports cell survival in EGFR-mutant lung adenocarcinoma by inhibiting autophagy-mediated EGFR degradation. Cancer Res. 2018;78:4482–96.PubMedCrossRef
139.
Zurück zum Zitat Takahashi D, Moriyama J, Nakamura T, Miki E, Takahashi E, Sato A, et al. Autacs: Cargo-specific degraders using selective autophagy. Mol Cell. 2019;76(797–810):e10. Takahashi D, Moriyama J, Nakamura T, Miki E, Takahashi E, Sato A, et al. Autacs: Cargo-specific degraders using selective autophagy. Mol Cell. 2019;76(797–810):e10.
140.
Zurück zum Zitat Ramalingam SS, Vansteenkiste J, Planchard D, Cho BC, Gray JE, Ohe Y, et al. Overall survival with osimertinib in untreated, EGFR-mutated advanced NSCLC. N Engl J Med. 2020;382:41–50.PubMedCrossRef Ramalingam SS, Vansteenkiste J, Planchard D, Cho BC, Gray JE, Ohe Y, et al. Overall survival with osimertinib in untreated, EGFR-mutated advanced NSCLC. N Engl J Med. 2020;382:41–50.PubMedCrossRef
141.
142.
Zurück zum Zitat He K, Xu J, Liang J, Jiang J, Tang M, Ye X, et al. Discovery of a novel EGFR-targeting antibody-drug conjugate, SHR-A1307, for the treatment of solid tumors resistant or refractory to anti-EGFR therapies. Mol Cancer Ther. 2019;18:1104–14.PubMedCrossRef He K, Xu J, Liang J, Jiang J, Tang M, Ye X, et al. Discovery of a novel EGFR-targeting antibody-drug conjugate, SHR-A1307, for the treatment of solid tumors resistant or refractory to anti-EGFR therapies. Mol Cancer Ther. 2019;18:1104–14.PubMedCrossRef
143.
Zurück zum Zitat Xu R-h, Qiu M-Z, Zhang Y, Wei X-L, Hu C. First-in-human dose-escalation study of anti-EGFR adc MRG003 in patients with relapsed/refractory solid tumors. J Clin Oncol. 2020;38:3550–50.CrossRef Xu R-h, Qiu M-Z, Zhang Y, Wei X-L, Hu C. First-in-human dose-escalation study of anti-EGFR adc MRG003 in patients with relapsed/refractory solid tumors. J Clin Oncol. 2020;38:3550–50.CrossRef
144.
Zurück zum Zitat Li Z, Wang M, Yao X, Luo W, Qu Y, Yu D, et al. Development of a novel EGFR-targeting antibody-drug conjugate for pancreatic cancer therapy. Target Oncol. 2019;14:93–105.PubMedCrossRef Li Z, Wang M, Yao X, Luo W, Qu Y, Yu D, et al. Development of a novel EGFR-targeting antibody-drug conjugate for pancreatic cancer therapy. Target Oncol. 2019;14:93–105.PubMedCrossRef
145.
Zurück zum Zitat Shi P, Oh YT, Zhang G, Yao W, Yue P, Li Y, 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.PubMedCrossRef Shi P, Oh YT, Zhang G, Yao W, Yue P, Li Y, 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.PubMedCrossRef
146.
Zurück zum Zitat 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:e232–4.PubMedCrossRef 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:e232–4.PubMedCrossRef
147.
Zurück zum Zitat Kang J, Chen HJ, Wang Z, Liu J, Li B, Zhang T, et al. Osimertinib and cabozantinib combinatorial therapy in an EGFR-mutant lung adenocarcinoma patient with multiple MET secondary-site mutations after resistance to crizotinib. J Thorac Oncol. 2018;13:e49–53.PubMedCrossRef Kang J, Chen HJ, Wang Z, Liu J, Li B, Zhang T, et al. Osimertinib and cabozantinib combinatorial therapy in an EGFR-mutant lung adenocarcinoma patient with multiple MET secondary-site mutations after resistance to crizotinib. J Thorac Oncol. 2018;13:e49–53.PubMedCrossRef
148.
Zurück zum Zitat Fujino T, Suda K, Mitsudomi T. Emerging MET tyrosine kinase inhibitors for the treatment of non-small cell lung cancer. Expert Opin Emerg Drugs. 2020;25:229–49.PubMedCrossRef Fujino T, Suda K, Mitsudomi T. Emerging MET tyrosine kinase inhibitors for the treatment of non-small cell lung cancer. Expert Opin Emerg Drugs. 2020;25:229–49.PubMedCrossRef
149.
Zurück zum Zitat Quintanal-Villalonga A, Molina-Pinelo S, Cirauqui C, Ojeda-Marquez L, Marrugal A, Suarez R, et al. FGFR1 cooperates with EGFR in lung cancer oncogenesis, and their combined inhibition shows improved efficacy. J Thorac Oncol. 2019;14:641–55.PubMedCrossRef Quintanal-Villalonga A, Molina-Pinelo S, Cirauqui C, Ojeda-Marquez L, Marrugal A, Suarez R, et al. FGFR1 cooperates with EGFR in lung cancer oncogenesis, and their combined inhibition shows improved efficacy. J Thorac Oncol. 2019;14:641–55.PubMedCrossRef
150.
Zurück zum Zitat Shaw AT, Felip E, Bauer TM, Besse B, Navarro A, Postel-Vinay S, et al. Lorlatinib in non-small-cell lung cancer with ALK or ROS1 rearrangement: an international, multicentre, open-label, single-arm first-in-man phase 1 trial. Lancet Oncol. 2017;18:1590–9.PubMedPubMedCentralCrossRef Shaw AT, Felip E, Bauer TM, Besse B, Navarro A, Postel-Vinay S, et al. Lorlatinib in non-small-cell lung cancer with ALK or ROS1 rearrangement: an international, multicentre, open-label, single-arm first-in-man phase 1 trial. Lancet Oncol. 2017;18:1590–9.PubMedPubMedCentralCrossRef
151.
Zurück zum Zitat Uchibori K, Inase N, Araki M, Kamada M, Sato S, Okuno Y, et al. Brigatinib combined with anti-EGFR antibody overcomes osimertinib resistance in EGFR-mutated non-small-cell lung cancer. Nat Commun. 2017;8:14768.PubMedPubMedCentralCrossRef Uchibori K, Inase N, Araki M, Kamada M, Sato S, Okuno Y, et al. Brigatinib combined with anti-EGFR antibody overcomes osimertinib resistance in EGFR-mutated non-small-cell lung cancer. Nat Commun. 2017;8:14768.PubMedPubMedCentralCrossRef
152.
Zurück zum Zitat Liu S, Li S, Hai J, Wang X, Chen T, Quinn MM, et al. Targeting HER2 aberrations in non-small cell lung cancer with osimertinib. Clin Cancer Res. 2018;24:2594–604.PubMedPubMedCentralCrossRef Liu S, Li S, Hai J, Wang X, Chen T, Quinn MM, et al. Targeting HER2 aberrations in non-small cell lung cancer with osimertinib. Clin Cancer Res. 2018;24:2594–604.PubMedPubMedCentralCrossRef
153.
Zurück zum Zitat La Monica S, Cretella D, Bonelli M, Fumarola C, Cavazzoni A, Digiacomo G, et al. Trastuzumab emtansine delays and overcomes resistance to the third-generation EGFR-TKI osimertinib in NSCLC EGFR mutated cell lines. J Exp Clin Cancer Res. 2017;36:174.PubMedPubMedCentralCrossRef La Monica S, Cretella D, Bonelli M, Fumarola C, Cavazzoni A, Digiacomo G, et al. Trastuzumab emtansine delays and overcomes resistance to the third-generation EGFR-TKI osimertinib in NSCLC EGFR mutated cell lines. J Exp Clin Cancer Res. 2017;36:174.PubMedPubMedCentralCrossRef
154.
Zurück zum Zitat Jani JP, Arcari J, Bernardo V, Bhattacharya SK, Briere D, Cohen BD, et al. PF-03814735, an orally bioavailable small molecule aurora kinase inhibitor for cancer therapy. Mol Cancer Ther. 2010;9:883–94.PubMedCrossRef Jani JP, Arcari J, Bernardo V, Bhattacharya SK, Briere D, Cohen BD, et al. PF-03814735, an orally bioavailable small molecule aurora kinase inhibitor for cancer therapy. Mol Cancer Ther. 2010;9:883–94.PubMedCrossRef
155.
Zurück zum Zitat Kim C, Giaccone G. MEK inhibitors under development for treatment of non-small-cell lung cancer. Expert Opin Investig Drugs. 2018;27:17–30.PubMedCrossRef Kim C, Giaccone G. MEK inhibitors under development for treatment of non-small-cell lung cancer. Expert Opin Investig Drugs. 2018;27:17–30.PubMedCrossRef
156.
Zurück zum Zitat Ortiz-Cuaran S, Scheffler M, Plenker D, Dahmen L, Scheel AH, Fernandez-Cuesta L, et al. Heterogeneous mechanisms of primary and acquired resistance to third-generation EGFR inhibitors. Clin Cancer Res. 2016;22:4837–47.PubMedCrossRef Ortiz-Cuaran S, Scheffler M, Plenker D, Dahmen L, Scheel AH, Fernandez-Cuesta L, et al. Heterogeneous mechanisms of primary and acquired resistance to third-generation EGFR inhibitors. Clin Cancer Res. 2016;22:4837–47.PubMedCrossRef
157.
Zurück zum Zitat Della Corte CM, Ciaramella V, Cardone C, La Monica S, Alfieri R, Petronini PG, et al. Antitumor efficacy of dual blockade of EGFR signaling by osimertinib in combination with selumetinib or cetuximab in activated EGFR human NCLC tumor models. J Thorac Oncol. 2018;13:810–20.PubMedCrossRef Della Corte CM, Ciaramella V, Cardone C, La Monica S, Alfieri R, Petronini PG, et al. Antitumor efficacy of dual blockade of EGFR signaling by osimertinib in combination with selumetinib or cetuximab in activated EGFR human NCLC tumor models. J Thorac Oncol. 2018;13:810–20.PubMedCrossRef
158.
Zurück zum Zitat Jacobsen K, Bertran-Alamillo J, Molina MA, Teixido C, Karachaliou N, Pedersen MH, et al. Convergent Akt activation drives acquired EGFR inhibitor resistance in lung cancer. Nat Commun. 2017;8:410.PubMedPubMedCentralCrossRef Jacobsen K, Bertran-Alamillo J, Molina MA, Teixido C, Karachaliou N, Pedersen MH, et al. Convergent Akt activation drives acquired EGFR inhibitor resistance in lung cancer. Nat Commun. 2017;8:410.PubMedPubMedCentralCrossRef
159.
Zurück zum Zitat Namba K, Shien K, Takahashi Y, Torigoe H, Sato H, Yoshioka T, 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. 2019;17:499–507.PubMedCrossRef Namba K, Shien K, Takahashi Y, Torigoe H, Sato H, Yoshioka T, 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. 2019;17:499–507.PubMedCrossRef
160.
Zurück zum Zitat Jimbo T, Hatanaka M, Komatsu T, Taira T, Kumazawa K, Maeda N, et al. DS-1205b, a novel selective inhibitor of AXL kinase, blocks resistance to EGFR-tyrosine kinase inhibitors in a non-small cell lung cancer xenograft model. Oncotarget. 2019;10:5152–67.PubMedPubMedCentralCrossRef Jimbo T, Hatanaka M, Komatsu T, Taira T, Kumazawa K, Maeda N, et al. DS-1205b, a novel selective inhibitor of AXL kinase, blocks resistance to EGFR-tyrosine kinase inhibitors in a non-small cell lung cancer xenograft model. Oncotarget. 2019;10:5152–67.PubMedPubMedCentralCrossRef
161.
Zurück zum Zitat Kim D, Bach DH, Fan YH, Luu TT, Hong JY, Park HJ, et al. AXL degradation in combination with EGFR-TKI can delay and overcome acquired resistance in human non-small cell lung cancer cells. Cell Death Dis. 2019;10:361.PubMedPubMedCentralCrossRef Kim D, Bach DH, Fan YH, Luu TT, Hong JY, Park HJ, et al. AXL degradation in combination with EGFR-TKI can delay and overcome acquired resistance in human non-small cell lung cancer cells. Cell Death Dis. 2019;10:361.PubMedPubMedCentralCrossRef
162.
Zurück zum Zitat Liu YN, Tsai MF, Wu SG, Chang TH, Tsai TH, Gow CH, et al. Acquired resistance to EGFR tyrosine kinase inhibitors is mediated by the reactivation of STC2/JUN/AXL signaling in lung cancer. Int J Cancer. 2019;145:1609–24.PubMedCrossRef Liu YN, Tsai MF, Wu SG, Chang TH, Tsai TH, Gow CH, et al. Acquired resistance to EGFR tyrosine kinase inhibitors is mediated by the reactivation of STC2/JUN/AXL signaling in lung cancer. Int J Cancer. 2019;145:1609–24.PubMedCrossRef
163.
Zurück zum Zitat Gu J, Qian L, Zhang G, Mahajan NP, Owonikoko TK, Ramalingam SS, et al. Inhibition of ACK1 delays and overcomes acquired resistance of EGFR mutant NSCLC cells to the third generation EGFR inhibitor, osimertinib. Lung Cancer. 2020;150:26–35.PubMedCrossRef Gu J, Qian L, Zhang G, Mahajan NP, Owonikoko TK, Ramalingam SS, et al. Inhibition of ACK1 delays and overcomes acquired resistance of EGFR mutant NSCLC cells to the third generation EGFR inhibitor, osimertinib. Lung Cancer. 2020;150:26–35.PubMedCrossRef
164.
Zurück zum Zitat Lawrence HR, Mahajan K, Luo Y, Zhang D, Tindall N, Huseyin M, et al. Development of novel ACK1/TNK2 inhibitors using a fragment-based approach. J Med Chem. 2015;58:2746–63.PubMedPubMedCentralCrossRef Lawrence HR, Mahajan K, Luo Y, Zhang D, Tindall N, Huseyin M, et al. Development of novel ACK1/TNK2 inhibitors using a fragment-based approach. J Med Chem. 2015;58:2746–63.PubMedPubMedCentralCrossRef
165.
Zurück zum Zitat Sequist LV, Lynch TJ. EGFR tyrosine kinase inhibitors in lung cancer: an evolving story. Annu Rev Med. 2008;59:429–42.PubMedCrossRef Sequist LV, Lynch TJ. EGFR tyrosine kinase inhibitors in lung cancer: an evolving story. Annu Rev Med. 2008;59:429–42.PubMedCrossRef
166.
Zurück zum Zitat Kummar S, Chen HX, Wright J, Holbeck S, Millin MD, Tomaszewski J, et al. Utilizing targeted cancer therapeutic agents in combination: novel approaches and urgent requirements. Nat Rev Drug Discov. 2010;9:843–56.PubMedCrossRef Kummar S, Chen HX, Wright J, Holbeck S, Millin MD, Tomaszewski J, et al. Utilizing targeted cancer therapeutic agents in combination: novel approaches and urgent requirements. Nat Rev Drug Discov. 2010;9:843–56.PubMedCrossRef
167.
Zurück zum Zitat Anighoro A, Bajorath J, Rastelli G. Polypharmacology: challenges and opportunities in drug discovery. J Med Chem. 2014;57:7874–87.PubMedCrossRef Anighoro A, Bajorath J, Rastelli G. Polypharmacology: challenges and opportunities in drug discovery. J Med Chem. 2014;57:7874–87.PubMedCrossRef
168.
Zurück zum Zitat Chen G, Bao Y, Weng Q, Zhao Y, Lu X, Fu L, et al. Compound 15c, a novel dual inhibitor of EGFR(L858R/T790M) and FGFR1, efficiently overcomes epidermal growth factor receptor-tyrosine kinase inhibitor resistance of non-small-cell lung cancers. Front Pharmacol. 2019;10:1533.PubMedCrossRef Chen G, Bao Y, Weng Q, Zhao Y, Lu X, Fu L, et al. Compound 15c, a novel dual inhibitor of EGFR(L858R/T790M) and FGFR1, efficiently overcomes epidermal growth factor receptor-tyrosine kinase inhibitor resistance of non-small-cell lung cancers. Front Pharmacol. 2019;10:1533.PubMedCrossRef
169.
Zurück zum Zitat Cui Z, Chen S, Wang Y, Gao C, Chen Y, Tan C, et al. Design, synthesis and evaluation of azaacridine derivatives as dual-target EGFR and Src kinase inhibitors for antitumor treatment. Eur J Med Chem. 2017;136:372–81.PubMedCrossRef Cui Z, Chen S, Wang Y, Gao C, Chen Y, Tan C, et al. Design, synthesis and evaluation of azaacridine derivatives as dual-target EGFR and Src kinase inhibitors for antitumor treatment. Eur J Med Chem. 2017;136:372–81.PubMedCrossRef
170.
Zurück zum Zitat Mansour TS, Pallepati RR, Basetti V. Potent dual EGFR/HER4 tyrosine kinase inhibitors containing novel (1,2-dithiolan-4-yl)acetamides. Bioorg Med Chem Lett. 2020;30:127288.PubMedCrossRef Mansour TS, Pallepati RR, Basetti V. Potent dual EGFR/HER4 tyrosine kinase inhibitors containing novel (1,2-dithiolan-4-yl)acetamides. Bioorg Med Chem Lett. 2020;30:127288.PubMedCrossRef
171.
Zurück zum Zitat El-Sayed NA, Nour MS, Salem MA, Arafa RK. New oxadiazoles with selective- COX-2 and EGFR dual inhibitory activity: design, synthesis, cytotoxicity evaluation and in silico studies. Eur J Med Chem. 2019;183:111693.PubMedCrossRef El-Sayed NA, Nour MS, Salem MA, Arafa RK. New oxadiazoles with selective- COX-2 and EGFR dual inhibitory activity: design, synthesis, cytotoxicity evaluation and in silico studies. Eur J Med Chem. 2019;183:111693.PubMedCrossRef
172.
Zurück zum Zitat Abdelatef SA, El-Saadi MT, Amin NH, Abdelazeem AH, Omar HA, Abdellatif KRA. Design, synthesis and anticancer evaluation of novel spirobenzo[h]chromene and spirochromane derivatives with dual EGFR and B-RAF inhibitory activities. Eur J Med Chem. 2018;150:567–78.PubMedCrossRef Abdelatef SA, El-Saadi MT, Amin NH, Abdelazeem AH, Omar HA, Abdellatif KRA. Design, synthesis and anticancer evaluation of novel spirobenzo[h]chromene and spirochromane derivatives with dual EGFR and B-RAF inhibitory activities. Eur J Med Chem. 2018;150:567–78.PubMedCrossRef
173.
Zurück zum Zitat Jang J, Son JB, To C, Bahcall M, Kim SY, Kang SY, et al. Discovery of a potent dual ALK and EGFR T790M inhibitor. Eur J Med Chem. 2017;136:497–510.PubMedPubMedCentralCrossRef Jang J, Son JB, To C, Bahcall M, Kim SY, Kang SY, et al. Discovery of a potent dual ALK and EGFR T790M inhibitor. Eur J Med Chem. 2017;136:497–510.PubMedPubMedCentralCrossRef
174.
Zurück zum Zitat Chen Y, Wu J, Wang A, Qi Z, Jiang T, Chen C, et al. Discovery of n-(5-((5-chloro-4-((2-(isopropylsulfonyl)phenyl)amino)pyrimidin-2-yl)amino)-4-met hoxy-2-(4-methyl-1,4-diazepan-1-yl)phenyl)acrylamide (chmfl-alk/EGFR-050) as a potent ALK/EGFR dual kinase inhibitor capable of overcoming a variety of ALK/EGFR associated drug resistant mutants in NSCLC. Eur J Med Chem. 2017;139:674–97.PubMedCrossRef Chen Y, Wu J, Wang A, Qi Z, Jiang T, Chen C, et al. Discovery of n-(5-((5-chloro-4-((2-(isopropylsulfonyl)phenyl)amino)pyrimidin-2-yl)amino)-4-met hoxy-2-(4-methyl-1,4-diazepan-1-yl)phenyl)acrylamide (chmfl-alk/EGFR-050) as a potent ALK/EGFR dual kinase inhibitor capable of overcoming a variety of ALK/EGFR associated drug resistant mutants in NSCLC. Eur J Med Chem. 2017;139:674–97.PubMedCrossRef
175.
Zurück zum Zitat Jing T, Miao X, Jiang F, Guo M, Xing L, Zhang J, et al. Discovery and optimization of tetrahydropyrido[4,3-d]pyrimidine derivatives as novel ATX and EGFR dual inhibitors. Bioorg Med Chem. 2018;26:1784–96.PubMedCrossRef Jing T, Miao X, Jiang F, Guo M, Xing L, Zhang J, et al. Discovery and optimization of tetrahydropyrido[4,3-d]pyrimidine derivatives as novel ATX and EGFR dual inhibitors. Bioorg Med Chem. 2018;26:1784–96.PubMedCrossRef
176.
Zurück zum Zitat Kurup S, McAllister B, Liskova P, Mistry T, Fanizza A, Stanford D, et al. Design, synthesis and biological activity of n(4)-phenylsubstituted-7h-pyrrolo[2,3-d]pyrimidin-4-amines as dual inhibitors of aurora kinase a and epidermal growth factor receptor kinase. J Enzyme Inhib Med Chem. 2018;33:74–84.PubMedCrossRef Kurup S, McAllister B, Liskova P, Mistry T, Fanizza A, Stanford D, et al. Design, synthesis and biological activity of n(4)-phenylsubstituted-7h-pyrrolo[2,3-d]pyrimidin-4-amines as dual inhibitors of aurora kinase a and epidermal growth factor receptor kinase. J Enzyme Inhib Med Chem. 2018;33:74–84.PubMedCrossRef
177.
Zurück zum Zitat Gadekar PK, Urunkar G, Roychowdhury A, Sharma R, Bose J, Khanna S, et al. Design, synthesis and biological evaluation of 2,3-dihydroimidazo[2,1-b]thiazoles as dual EGFR and IGF1R inhibitors. Bioorg Chem. 2021;115:105151.PubMedCrossRef Gadekar PK, Urunkar G, Roychowdhury A, Sharma R, Bose J, Khanna S, et al. Design, synthesis and biological evaluation of 2,3-dihydroimidazo[2,1-b]thiazoles as dual EGFR and IGF1R inhibitors. Bioorg Chem. 2021;115:105151.PubMedCrossRef
178.
Zurück zum Zitat Romagnoli R, Prencipe F, Oliva P, Baraldi S, Baraldi PG, Schiaffino Ortega S, et al. Design, synthesis, and biological evaluation of 6-substituted thieno[3,2-d]pyrimidine analogues as dual epidermal growth factor receptor kinase and microtubule inhibitors. J Med Chem. 2019;62:1274–90.PubMedCrossRef Romagnoli R, Prencipe F, Oliva P, Baraldi S, Baraldi PG, Schiaffino Ortega S, et al. Design, synthesis, and biological evaluation of 6-substituted thieno[3,2-d]pyrimidine analogues as dual epidermal growth factor receptor kinase and microtubule inhibitors. J Med Chem. 2019;62:1274–90.PubMedCrossRef
179.
Zurück zum Zitat Alswah M, Bayoumi AH, Elgamal K, Elmorsy A, Ihmaid S, Ahmed HEA. Design, synthesis and cytotoxic evaluation of novel chalcone derivatives bearing triazolo[4,3-a]-quinoxaline moieties as potent anticancer agents with dual EGFR kinase and tubulin polymerization inhibitory effects. Molecules. 2017;23:48.PubMedCentralCrossRef Alswah M, Bayoumi AH, Elgamal K, Elmorsy A, Ihmaid S, Ahmed HEA. Design, synthesis and cytotoxic evaluation of novel chalcone derivatives bearing triazolo[4,3-a]-quinoxaline moieties as potent anticancer agents with dual EGFR kinase and tubulin polymerization inhibitory effects. Molecules. 2017;23:48.PubMedCentralCrossRef
180.
Zurück zum Zitat Khan I, Garikapati KR, Setti A, Shaik AB, Kanth Makani VK, Shareef MA, et al. Design, synthesis, in silico pharmacokinetics prediction and biological evaluation of 1,4-dihydroindeno[1,2-c]pyrazole chalcone as EGFR/AKT pathway inhibitors. Eur J Med Chem. 2019;163:636–48.PubMedCrossRef Khan I, Garikapati KR, Setti A, Shaik AB, Kanth Makani VK, Shareef MA, et al. Design, synthesis, in silico pharmacokinetics prediction and biological evaluation of 1,4-dihydroindeno[1,2-c]pyrazole chalcone as EGFR/AKT pathway inhibitors. Eur J Med Chem. 2019;163:636–48.PubMedCrossRef
181.
Zurück zum Zitat Dong H, Yin H, Zhao C, Cao J, Xu W, Zhang Y. Design, synthesis and biological evaluation of novel osimertinib-based HDAC and EGFR dual inhibitors. Molecules. 2019;24:2407.PubMedCentralCrossRef Dong H, Yin H, Zhao C, Cao J, Xu W, Zhang Y. Design, synthesis and biological evaluation of novel osimertinib-based HDAC and EGFR dual inhibitors. Molecules. 2019;24:2407.PubMedCentralCrossRef
182.
Zurück zum Zitat Fischer T, Najjar A, Totzke F, Schachtele C, Sippl W, Ritter C, et al. Discovery of novel dual inhibitors of receptor tyrosine kinases EGFR and PDGFR-β related to anticancer drug resistance. J Enzyme Inhib Med Chem. 2018;33:1–8.PubMedCrossRef Fischer T, Najjar A, Totzke F, Schachtele C, Sippl W, Ritter C, et al. Discovery of novel dual inhibitors of receptor tyrosine kinases EGFR and PDGFR-β related to anticancer drug resistance. J Enzyme Inhib Med Chem. 2018;33:1–8.PubMedCrossRef
183.
Zurück zum Zitat Hamed MM, Darwish SS, Herrmann J, Abadi AH, Engel M. First bispecific inhibitors of the epidermal growth factor receptor kinase and the NF-κB activity as novel anticancer agents. J Med Chem. 2017;60:2853–68.PubMedCrossRef Hamed MM, Darwish SS, Herrmann J, Abadi AH, Engel M. First bispecific inhibitors of the epidermal growth factor receptor kinase and the NF-κB activity as novel anticancer agents. J Med Chem. 2017;60:2853–68.PubMedCrossRef
184.
Zurück zum Zitat Dokla EME, Fang CS, Abouzid KAM, Chen CS. 1,2,4-oxadiazole derivatives targeting EGFR and c-Met degradation in TKI resistant NSCLC. Eur J Med Chem. 2019;182:111607.PubMedCrossRef Dokla EME, Fang CS, Abouzid KAM, Chen CS. 1,2,4-oxadiazole derivatives targeting EGFR and c-Met degradation in TKI resistant NSCLC. Eur J Med Chem. 2019;182:111607.PubMedCrossRef
185.
Zurück zum Zitat Singh PK, Silakari O. Molecular dynamics guided development of indole based dual inhibitors of EGFR (T790M) and c-Met. Bioorg Chem. 2018;79:163–70.PubMedCrossRef Singh PK, Silakari O. Molecular dynamics guided development of indole based dual inhibitors of EGFR (T790M) and c-Met. Bioorg Chem. 2018;79:163–70.PubMedCrossRef
186.
Zurück zum Zitat Fischer T, Kruger T, Najjar A, Totzke F, Schachtele C, Sippl W, et al. Discovery of novel substituted benzo-anellated 4-benzylamino pyrrolopyrimidines as dual EGFR and vEGFR2 inhibitors. Bioorg Med Chem Lett. 2017;27:2708–12.PubMedCrossRef Fischer T, Kruger T, Najjar A, Totzke F, Schachtele C, Sippl W, et al. Discovery of novel substituted benzo-anellated 4-benzylamino pyrrolopyrimidines as dual EGFR and vEGFR2 inhibitors. Bioorg Med Chem Lett. 2017;27:2708–12.PubMedCrossRef
187.
Zurück zum Zitat Zhang HQ, Gong FH, Ye JQ, Zhang C, Yue XH, Li CG, et al. Design and discovery of 4-anilinoquinazoline-urea derivatives as dual TK inhibitors of EGFR and vEGFR-2. Eur J Med Chem. 2017;125:245–54.PubMedCrossRef Zhang HQ, Gong FH, Ye JQ, Zhang C, Yue XH, Li CG, et al. Design and discovery of 4-anilinoquinazoline-urea derivatives as dual TK inhibitors of EGFR and vEGFR-2. Eur J Med Chem. 2017;125:245–54.PubMedCrossRef
188.
Zurück zum Zitat Wei H, Duan Y, Gou W, Cui J, Ning H, Li D, et al. Design, synthesis and biological evaluation of novel 4-anilinoquinazoline derivatives as hypoxia-selective EGFR and vEGFR-2 dual inhibitors. Eur J Med Chem. 2019;181:111552.PubMedCrossRef Wei H, Duan Y, Gou W, Cui J, Ning H, Li D, et al. Design, synthesis and biological evaluation of novel 4-anilinoquinazoline derivatives as hypoxia-selective EGFR and vEGFR-2 dual inhibitors. Eur J Med Chem. 2019;181:111552.PubMedCrossRef
189.
Zurück zum Zitat Sun S, Zhang J, Wang N, Kong X, Fu F, Wang H, et al. Design and discovery of quinazoline- and thiourea-containing sorafenib analogs as EGFR and vEGFR-2 dual TK inhibitors. Molecules. 2017;23:24.PubMedCentralCrossRef Sun S, Zhang J, Wang N, Kong X, Fu F, Wang H, et al. Design and discovery of quinazoline- and thiourea-containing sorafenib analogs as EGFR and vEGFR-2 dual TK inhibitors. Molecules. 2017;23:24.PubMedCentralCrossRef
190.
Zurück zum Zitat Das D, Xie L, Wang J, Xu X, Zhang Z, Shi J, et al. Discovery of new quinazoline derivatives as irreversible dual EGFR/HER2 inhibitors and their anticancer activities: part 1. Bioorg Med Chem Lett. 2019;29:591–6.PubMedCrossRef Das D, Xie L, Wang J, Xu X, Zhang Z, Shi J, et al. Discovery of new quinazoline derivatives as irreversible dual EGFR/HER2 inhibitors and their anticancer activities: part 1. Bioorg Med Chem Lett. 2019;29:591–6.PubMedCrossRef
191.
Zurück zum Zitat Maher M, Kassab AE, Zaher AF, Mahmoud Z. Novel pyrazolo[3,4-d]pyrimidines: design, synthesis, anticancer activity, dual EGFR/ErbB2 receptor tyrosine kinases inhibitory activity, effects on cell cycle profile and caspase-3-mediated apoptosis. J Enzyme Inhib Med Chem. 2019;34:532–46.PubMedPubMedCentralCrossRef Maher M, Kassab AE, Zaher AF, Mahmoud Z. Novel pyrazolo[3,4-d]pyrimidines: design, synthesis, anticancer activity, dual EGFR/ErbB2 receptor tyrosine kinases inhibitory activity, effects on cell cycle profile and caspase-3-mediated apoptosis. J Enzyme Inhib Med Chem. 2019;34:532–46.PubMedPubMedCentralCrossRef
192.
Zurück zum Zitat Zou M, Li J, Jin B, Wang M, Chen H, Zhang Z, et al. Design, synthesis and anticancer evaluation of new 4-anilinoquinoline-3-carbonitrile derivatives as dual EGFR/HER2 inhibitors and apoptosis inducers. Bioorg Chem. 2021;114:105200.PubMedCrossRef Zou M, Li J, Jin B, Wang M, Chen H, Zhang Z, et al. Design, synthesis and anticancer evaluation of new 4-anilinoquinoline-3-carbonitrile derivatives as dual EGFR/HER2 inhibitors and apoptosis inducers. Bioorg Chem. 2021;114:105200.PubMedCrossRef
193.
Zurück zum Zitat Alsaid MS, Al-Mishari AA, Soliman AM, Ragab FA, Ghorab MM. Discovery of benzo[g]quinazolin benzenesulfonamide derivatives as dual EGFR/HER2 inhibitors. Eur J Med Chem. 2017;141:84–91.PubMedCrossRef Alsaid MS, Al-Mishari AA, Soliman AM, Ragab FA, Ghorab MM. Discovery of benzo[g]quinazolin benzenesulfonamide derivatives as dual EGFR/HER2 inhibitors. Eur J Med Chem. 2017;141:84–91.PubMedCrossRef
194.
Zurück zum Zitat Ghorab MM, Alsaid MS, Soliman AM. Dual EGFR/HER2 inhibitors and apoptosis inducers: new benzo[g]quinazoline derivatives bearing benzenesulfonamide as anticancer and radiosensitizers. Bioorg Chem. 2018;80:611–20.PubMedCrossRef Ghorab MM, Alsaid MS, Soliman AM. Dual EGFR/HER2 inhibitors and apoptosis inducers: new benzo[g]quinazoline derivatives bearing benzenesulfonamide as anticancer and radiosensitizers. Bioorg Chem. 2018;80:611–20.PubMedCrossRef
195.
Zurück zum Zitat Soliman AM, Alqahtani AS, Ghorab M. Novel sulphonamide benzoquinazolinones as dual EGFR/HER2 inhibitors, apoptosis inducers and radiosensitizers. J Enzyme Inhib Med Chem. 2019;34:1030–40.PubMedPubMedCentralCrossRef Soliman AM, Alqahtani AS, Ghorab M. Novel sulphonamide benzoquinazolinones as dual EGFR/HER2 inhibitors, apoptosis inducers and radiosensitizers. J Enzyme Inhib Med Chem. 2019;34:1030–40.PubMedPubMedCentralCrossRef
196.
Zurück zum Zitat Liu X, Du Q, Tian C, Tang M, Jiang Y, Wang Y, et al. Discovery of cape derivatives as dual EGFR and CSK inhibitors with anticancer activity in a murine model of hepatocellular carcinoma. Bioorg Chem. 2021;107:104536.PubMedCrossRef Liu X, Du Q, Tian C, Tang M, Jiang Y, Wang Y, et al. Discovery of cape derivatives as dual EGFR and CSK inhibitors with anticancer activity in a murine model of hepatocellular carcinoma. Bioorg Chem. 2021;107:104536.PubMedCrossRef
197.
Zurück zum Zitat Zhang B, Liu Z, Xia S, Liu Q, Gou S. Design, synthesis and biological evaluation of sulfamoylphenyl-quinazoline derivatives as potential EGFR/CAIX dual inhibitors. Eur J Med Chem. 2021;216:113300.PubMedCrossRef Zhang B, Liu Z, Xia S, Liu Q, Gou S. Design, synthesis and biological evaluation of sulfamoylphenyl-quinazoline derivatives as potential EGFR/CAIX dual inhibitors. Eur J Med Chem. 2021;216:113300.PubMedCrossRef
198.
Zurück zum Zitat Zang H, Qian G, Arbiser J, Owonikoko TK, Ramalingam SS, Fan S, et al. Overcoming acquired resistance of EGFR-mutant NSCLC cells to the third generation EGFR inhibitor, osimertinib, with the natural product honokiol. Mol Oncol. 2020;14:882–95.PubMedPubMedCentralCrossRef Zang H, Qian G, Arbiser J, Owonikoko TK, Ramalingam SS, Fan S, et al. Overcoming acquired resistance of EGFR-mutant NSCLC cells to the third generation EGFR inhibitor, osimertinib, with the natural product honokiol. Mol Oncol. 2020;14:882–95.PubMedPubMedCentralCrossRef
199.
Zurück zum Zitat Cao F, Gong YB, Kang XH, Lu ZH, Wang Y, Zhao KL, et al. Degradation of MCL-1 by bufalin reverses acquired resistance to osimertinib in EGFR-mutant lung cancer. Toxicol Appl Pharmacol. 2019;379:114662.PubMedCrossRef Cao F, Gong YB, Kang XH, Lu ZH, Wang Y, Zhao KL, et al. Degradation of MCL-1 by bufalin reverses acquired resistance to osimertinib in EGFR-mutant lung cancer. Toxicol Appl Pharmacol. 2019;379:114662.PubMedCrossRef
200.
Zurück zum Zitat Sun P, Qu Y, Wang Y, Wang J, Wang X, Sheng J. Wighteone exhibits an antitumor effect against EGFR L858R/T790M mutation non-small cell lung cancer. J Cancer. 2021;12:3900–8.PubMedPubMedCentralCrossRef Sun P, Qu Y, Wang Y, Wang J, Wang X, Sheng J. Wighteone exhibits an antitumor effect against EGFR L858R/T790M mutation non-small cell lung cancer. J Cancer. 2021;12:3900–8.PubMedPubMedCentralCrossRef
201.
Zurück zum Zitat Niu M, Xu J, Liu Y, Li Y, He T, Ding L, et al. FBXL 2 counteracts Grp94 to destabilize EGFR and inhibit EGFR-driven NSCLC growth. Nat Commun. 2021;12:5919.PubMedPubMedCentralCrossRef Niu M, Xu J, Liu Y, Li Y, He T, Ding L, et al. FBXL 2 counteracts Grp94 to destabilize EGFR and inhibit EGFR-driven NSCLC growth. Nat Commun. 2021;12:5919.PubMedPubMedCentralCrossRef
202.
Zurück zum Zitat Zhang KR, Zhang YF, Lei HM, Tang YB, Ma CS, Lv QM, et al. Targeting AKR1B1 inhibits glutathione de novo synthesis to overcome acquired resistance to EGFR-targeted therapy in lung cancer. Sci Transl Med. 2021;13:eabg6428.PubMedCrossRef Zhang KR, Zhang YF, Lei HM, Tang YB, Ma CS, Lv QM, et al. Targeting AKR1B1 inhibits glutathione de novo synthesis to overcome acquired resistance to EGFR-targeted therapy in lung cancer. Sci Transl Med. 2021;13:eabg6428.PubMedCrossRef
203.
Zurück zum Zitat Hitosugi T, Zhou L, Elf S, Fan J, Kang HB, Seo JH, et al. Phosphoglycerate mutase 1 coordinates glycolysis and biosynthesis to promote tumor growth. Cancer Cell. 2012;22:585–600.PubMedPubMedCentralCrossRef Hitosugi T, Zhou L, Elf S, Fan J, Kang HB, Seo JH, et al. Phosphoglycerate mutase 1 coordinates glycolysis and biosynthesis to promote tumor growth. Cancer Cell. 2012;22:585–600.PubMedPubMedCentralCrossRef
204.
Zurück zum Zitat Liang Q, Gu WM, Huang K, Luo MY, Zou JH, Zhuang GL, et al. HKB99, an allosteric inhibitor of phosphoglycerate mutase 1, suppresses invasive pseudopodia formation and upregulates plasminogen activator inhibitor-2 in erlotinib-resistant non-small cell lung cancer cells. Acta Pharmacol Sin. 2021;42:115–9.PubMedCrossRef Liang Q, Gu WM, Huang K, Luo MY, Zou JH, Zhuang GL, et al. HKB99, an allosteric inhibitor of phosphoglycerate mutase 1, suppresses invasive pseudopodia formation and upregulates plasminogen activator inhibitor-2 in erlotinib-resistant non-small cell lung cancer cells. Acta Pharmacol Sin. 2021;42:115–9.PubMedCrossRef
205.
Zurück zum Zitat Huang K, Liang Q, Zhou Y, Jiang LL, Gu WM, Luo MY, et al. A novel allosteric inhibitor of phosphoglycerate mutase 1 suppresses growth and metastasis of non-small-cell lung cancer. Cell Metab. 2021;33:223.PubMedCrossRef Huang K, Liang Q, Zhou Y, Jiang LL, Gu WM, Luo MY, et al. A novel allosteric inhibitor of phosphoglycerate mutase 1 suppresses growth and metastasis of non-small-cell lung cancer. Cell Metab. 2021;33:223.PubMedCrossRef
206.
Zurück zum Zitat Huang K, Liang Q, Zhou Y, Jiang L-l, Gu W-m, Luo M-y, et al. A novel allosteric inhibitor of phosphoglycerate mutase 1 suppresses growth and metastasis of non-small-cell lung cancer. Cell Metab. 2019;30:1107-19.e8.PubMedCrossRef Huang K, Liang Q, Zhou Y, Jiang L-l, Gu W-m, Luo M-y, et al. A novel allosteric inhibitor of phosphoglycerate mutase 1 suppresses growth and metastasis of non-small-cell lung cancer. Cell Metab. 2019;30:1107-19.e8.PubMedCrossRef
207.
Zurück zum Zitat Qiu Y, Yin X, Li X, Wang Y, Fu Q, Huang R, et al. Untangling dual-targeting therapeutic mechanism of epidermal growth factor receptor (EGFR) based on reversed allosteric communication. Pharmaceutics. 2021;13:747.PubMedPubMedCentralCrossRef Qiu Y, Yin X, Li X, Wang Y, Fu Q, Huang R, et al. Untangling dual-targeting therapeutic mechanism of epidermal growth factor receptor (EGFR) based on reversed allosteric communication. Pharmaceutics. 2021;13:747.PubMedPubMedCentralCrossRef
208.
Zurück zum Zitat Yin L, Zhang Y, Yin L, Ou Y, Lewis MS, Wang R, et al. Novel mitochondria-based targeting restores responsiveness in therapeutically resistant human lung cancer cells. Mol Cancer Ther. 2021;20(12):2527–38.PubMedCrossRef Yin L, Zhang Y, Yin L, Ou Y, Lewis MS, Wang R, et al. Novel mitochondria-based targeting restores responsiveness in therapeutically resistant human lung cancer cells. Mol Cancer Ther. 2021;20(12):2527–38.PubMedCrossRef
209.
Zurück zum Zitat He J, Huang Z, Han L, Gong Y, Xie C. Mechanisms and management of 3rd-generation EGFR-TKI resistance in advanced non-small cell lung cancer (Review). Int J Oncol. 2021;59:90.PubMedPubMedCentralCrossRef He J, Huang Z, Han L, Gong Y, Xie C. Mechanisms and management of 3rd-generation EGFR-TKI resistance in advanced non-small cell lung cancer (Review). Int J Oncol. 2021;59:90.PubMedPubMedCentralCrossRef
210.
Zurück zum Zitat Planchard D, Feng PH, Karaseva N, Kim SW, Kim TM, Lee CK, et al. Osimertinib plus platinum–pemetrexed in newly diagnosed epidermal growth factor receptor mutation-positive advanced/metastatic non-small-cell lung cancer: safety run-in results from the FLAURA2 study. ESMO Open. 2020;6:100271.CrossRef Planchard D, Feng PH, Karaseva N, Kim SW, Kim TM, Lee CK, et al. Osimertinib plus platinum–pemetrexed in newly diagnosed epidermal growth factor receptor mutation-positive advanced/metastatic non-small-cell lung cancer: safety run-in results from the FLAURA2 study. ESMO Open. 2020;6:100271.CrossRef
211.
Zurück zum Zitat Yi M, Zheng X, Niu M, Zhu S, Ge H, Wu K. Combination strategies with PD-1/PD-L1 blockade: current advances and future directions. Mol Cancer. 2022;21:28.PubMedPubMedCentralCrossRef Yi M, Zheng X, Niu M, Zhu S, Ge H, Wu K. Combination strategies with PD-1/PD-L1 blockade: current advances and future directions. Mol Cancer. 2022;21:28.PubMedPubMedCentralCrossRef
212.
Zurück zum Zitat Gandhi L, Rodriguez-Abreu D, Gadgeel S, Esteban E, Felip E, De Angelis F, et al. Pembrolizumab plus chemotherapy in metastatic non-small-cell lung cancer. N Engl J Med. 2018;378:2078–92.PubMedCrossRef Gandhi L, Rodriguez-Abreu D, Gadgeel S, Esteban E, Felip E, De Angelis F, et al. Pembrolizumab plus chemotherapy in metastatic non-small-cell lung cancer. N Engl J Med. 2018;378:2078–92.PubMedCrossRef
213.
Zurück zum Zitat Paz-Ares L, Luft A, Vicente D, Tafreshi A, Gumus M, Mazieres J, et al. Pembrolizumab plus chemotherapy for squamous non-small-cell lung cancer. N Engl J Med. 2018;379:2040–51.PubMedCrossRef Paz-Ares L, Luft A, Vicente D, Tafreshi A, Gumus M, Mazieres J, et al. Pembrolizumab plus chemotherapy for squamous non-small-cell lung cancer. N Engl J Med. 2018;379:2040–51.PubMedCrossRef
214.
Zurück zum Zitat Zhou C, Wu L, Fan Y, Wang Z, Liu L, Chen G, et al. Sintilimab plus platinum and gemcitabine as first-line treatment for advanced or metastatic squamous nsclc: results from a randomized, double-blind, phase 3 trial (ORIENT-12). J Thorac Oncol. 2021;16:1501–11.PubMedCrossRef Zhou C, Wu L, Fan Y, Wang Z, Liu L, Chen G, et al. Sintilimab plus platinum and gemcitabine as first-line treatment for advanced or metastatic squamous nsclc: results from a randomized, double-blind, phase 3 trial (ORIENT-12). J Thorac Oncol. 2021;16:1501–11.PubMedCrossRef
215.
Zurück zum Zitat Socinski MA, Jotte RM, Cappuzzo F, Orlandi F, Stroyakovskiy D, Nogami N, et al. Atezolizumab for first-line treatment of metastatic nonsquamous NSCLC. N Engl J Med. 2018;378:2288–301.PubMedCrossRef Socinski MA, Jotte RM, Cappuzzo F, Orlandi F, Stroyakovskiy D, Nogami N, et al. Atezolizumab for first-line treatment of metastatic nonsquamous NSCLC. N Engl J Med. 2018;378:2288–301.PubMedCrossRef
216.
Zurück zum Zitat Jabbour SK, Berman AT, Decker RH, Lin Y, Feigenberg SJ, Gettinger SN, et al. Phase 1 trial of pembrolizumab administered concurrently with chemoradiotherapy for locally advanced non-small cell lung cancer: a nonrandomized controlled trial. JAMA Oncol. 2020;6:848–55.PubMedCrossRef Jabbour SK, Berman AT, Decker RH, Lin Y, Feigenberg SJ, Gettinger SN, et al. Phase 1 trial of pembrolizumab administered concurrently with chemoradiotherapy for locally advanced non-small cell lung cancer: a nonrandomized controlled trial. JAMA Oncol. 2020;6:848–55.PubMedCrossRef
217.
Zurück zum Zitat Liu D, Gong J, Liu T, Li K, Yin X, Liu Y, et al. Phase 1 study of SHR-1701, a bifunctional fusion protein targeting PD-L1 and TGF-β, in patients with advanced solid tumors. J Clin Oncol. 2021;39:2503–2503.CrossRef Liu D, Gong J, Liu T, Li K, Yin X, Liu Y, et al. Phase 1 study of SHR-1701, a bifunctional fusion protein targeting PD-L1 and TGF-β, in patients with advanced solid tumors. J Clin Oncol. 2021;39:2503–2503.CrossRef
218.
Zurück zum Zitat Sequist LV, Waltman BA, Dias-Santagata D, Digumarthy S, Turke AB, Fidias P, et al. Genotypic and histological evolution of lung cancers acquiring resistance to EGFR inhibitors. Sci Trans Med. 2011;3:75ra26.CrossRef Sequist LV, Waltman BA, Dias-Santagata D, Digumarthy S, Turke AB, Fidias P, et al. Genotypic and histological evolution of lung cancers acquiring resistance to EGFR inhibitors. Sci Trans Med. 2011;3:75ra26.CrossRef
219.
Zurück zum Zitat 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;10:602762–862.PubMedPubMedCentralCrossRef 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;10:602762–862.PubMedPubMedCentralCrossRef
Metadaten
Titel
Emerging strategies to overcome resistance to third-generation EGFR inhibitors
verfasst von
Kunyu Shi
Guan Wang
Junping Pei
Jifa Zhang
Jiaxing Wang
Liang Ouyang
Yuxi Wang
Weimin Li
Publikationsdatum
01.12.2022
Verlag
BioMed Central
Erschienen in
Journal of Hematology & Oncology / Ausgabe 1/2022
Elektronische ISSN: 1756-8722
DOI
https://doi.org/10.1186/s13045-022-01311-6

Weitere Artikel der Ausgabe 1/2022

Journal of Hematology & Oncology 1/2022 Zur Ausgabe

Erhöhtes Risiko fürs Herz unter Checkpointhemmer-Therapie

28.05.2024 Nebenwirkungen der Krebstherapie Nachrichten

Kardiotoxische Nebenwirkungen einer Therapie mit Immuncheckpointhemmern mögen selten sein – wenn sie aber auftreten, wird es für Patienten oft lebensgefährlich. Voruntersuchung und Monitoring sind daher obligat.

Positiver FIT: Die Ursache liegt nicht immer im Dickdarm

27.05.2024 Blut im Stuhl Nachrichten

Immunchemischer Stuhltest positiv, Koloskopie negativ – in solchen Fällen kann die Blutungsquelle auch weiter proximal sitzen. Ein Forschungsteam hat nachgesehen, wie häufig und in welchen Lokalisationen das der Fall ist.

Mammakarzinom: Brustdichte beeinflusst rezidivfreies Überleben

26.05.2024 Mammakarzinom Nachrichten

Frauen, die zum Zeitpunkt der Brustkrebsdiagnose eine hohe mammografische Brustdichte aufweisen, haben ein erhöhtes Risiko für ein baldiges Rezidiv, legen neue Daten nahe.

Mehr Lebenszeit mit Abemaciclib bei fortgeschrittenem Brustkrebs?

24.05.2024 Mammakarzinom Nachrichten

In der MONARCHE-3-Studie lebten Frauen mit fortgeschrittenem Hormonrezeptor-positivem, HER2-negativem Brustkrebs länger, wenn sie zusätzlich zu einem nicht steroidalen Aromatasehemmer mit Abemaciclib behandelt wurden; allerdings verfehlte der numerische Zugewinn die statistische Signifikanz.

Update Onkologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.