Europe PMC
Do data resources managed by EMBL-EBI and our collaborators make a difference to your work?
If so, please take 10 minutes to fill in our survey, and help us make the case for why sustaining open data resources is critical for life sciences research.

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


The United States Food and Drug Administration (US FDA) has always been a forerunner in drug evaluation and supervision. Over the past 31 years, 1050 drugs (excluding vaccines, cell-based therapies, and gene therapy products) have been approved as new molecular entities (NMEs) or biologics license applications (BLAs). A total of 228 of these 1050 drugs were identified as cancer therapeutics or cancer-related drugs, and 120 of them were classified as therapeutic drugs for solid tumors according to their initial indications. These drugs have evolved from small molecules with broad-spectrum antitumor properties in the early stage to monoclonal antibodies (mAbs) and antibody‒drug conjugates (ADCs) with a more precise targeting effect during the most recent decade. These drugs have extended indications for other malignancies, constituting a cancer treatment system for monotherapy or combined therapy. However, the available targets are still mainly limited to receptor tyrosine kinases (RTKs), restricting the development of antitumor drugs. In this review, these 120 drugs are summarized and classified according to the initial indications, characteristics, or functions. Additionally, RTK-targeted therapies and immune checkpoint-based immunotherapies are also discussed. Our analysis of existing challenges and potential opportunities in drug development may advance solid tumor treatment in the future.

Free full text 


Logo of jhemoncLink to Publisher's site
J Hematol Oncol. 2022; 15: 143.
Published online 2022 Oct 8. https://doi.org/10.1186/s13045-022-01362-9
PMCID: PMC9548212
PMID: 36209184

Small-molecule inhibitors, immune checkpoint inhibitors, and more: FDA-approved novel therapeutic drugs for solid tumors from 1991 to 2021

Associated Data

Supplementary Materials
Data Availability Statement

Abstract

The United States Food and Drug Administration (US FDA) has always been a forerunner in drug evaluation and supervision. Over the past 31 years, 1050 drugs (excluding vaccines, cell-based therapies, and gene therapy products) have been approved as new molecular entities (NMEs) or biologics license applications (BLAs). A total of 228 of these 1050 drugs were identified as cancer therapeutics or cancer-related drugs, and 120 of them were classified as therapeutic drugs for solid tumors according to their initial indications. These drugs have evolved from small molecules with broad-spectrum antitumor properties in the early stage to monoclonal antibodies (mAbs) and antibody‒drug conjugates (ADCs) with a more precise targeting effect during the most recent decade. These drugs have extended indications for other malignancies, constituting a cancer treatment system for monotherapy or combined therapy. However, the available targets are still mainly limited to receptor tyrosine kinases (RTKs), restricting the development of antitumor drugs. In this review, these 120 drugs are summarized and classified according to the initial indications, characteristics, or functions. Additionally, RTK-targeted therapies and immune checkpoint-based immunotherapies are also discussed. Our analysis of existing challenges and potential opportunities in drug development may advance solid tumor treatment in the future.

Supplementary Information

The online version contains supplementary material available at 10.1186/s13045-022-01362-9.

Keywords: The United States Food and Drug Administration, Solid tumors, Receptor tyrosine kinase inhibitors, Immune checkpoint blockades

Background

Cancer is the first or second leading cause of premature death in all countries except Africa, second only to cardiovascular disease [1]. An estimated 19.3 million new cancer cases and almost 10 million cancer-related deaths occurred in 2020 worldwide [2]. Solid tumors represent more than 90% of human cancers and cancer-related mortalities [2]. For unresectable locally advanced or metastatic solid tumors, therapeutic drugs have always been the mainstream strategy. Profound changes have occurred in therapeutic drugs for solid tumors during the past 31 years. Both the number of solid tumor drugs and their proportion among all FDA-approved drugs increased in this period, especially in the most recent decade (Fig. 1a, b). More importantly, cytotoxic drugs have evolved into drugs with more precise targeting effects, including small-molecule targeted drugs, monoclonal antibodies (mAbs), and antibody–drug conjugates (ADCs), and the proportion of biological drugs has increased accordingly (Fig. 1c).

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig1_HTML.jpg

Statistics of FDA-approved drugs and cancer drugs. a Number of FDA-approved drugs (NMEs: New molecular entities, BLAs: Biologics license applications) over the past 31 years. b Number of FDA-approved cancer drugs over the past 31 years. c Number of FDA-approved therapeutic drugs for solid tumors during the past 31 years

During the past three decades, the FDA granted 120 approvals for novel solid tumor therapeutic drugs (Additional file 1: Table S1–S3), and these drugs treat the most high-incidence solid tumors, including lung cancer, breast cancer, prostate cancer, gastrointestinal cancers, etc. These drugs constitute the mainstay of the modern cancer treatment system for solid tumors and hematological malignancies. Despite extraordinary achievements, the effective application of these drugs is still limited by great challenges, such as drug resistance [3], adverse effects [4], and even hyperprogressive disease with programmed death receptor-1 (PD1)/programmed death-ligand 1 (PDL1)-based immunotherapy [5].

This review describes the properties of 120 therapeutic drugs for solid tumors, summarizes the main biological mechanisms of their antitumor activity, and analyzes the target distribution of these drugs. Additionally, we elaborate on the challenges and opportunities in developing solid tumor therapeutic drugs and provide constructive suggestions and helpful solutions for the further study of solid tumor treatment.

FDA-approved therapeutic drugs for lung cancers

Lung cancer accounted for 11.4% of cancer cases and 18.0% of cancer-related deaths worldwide in 2020. Although the incidence rate of lung cancer was surpassed by that of breast cancer in 2020, its mortality rate still far exceeded that of any other type of cancer [2]. Over the past 31 years, the FDA has granted approvals for 22 novel therapeutic drugs (including 20 small molecules and two mAbs) for lung cancer.

Non-small cell lung cancer

Non-small cell lung cancer (NSCLC) includes adenocarcinoma, squamous cell carcinoma (SCC), and large-cell carcinoma (LCC) and accounts for approximately 85% of all lung cancer cases [6]. The majority of diagnosed NSCLC cases present as locally advanced or metastatic diseases [7]. Twenty of the 22 therapeutic drugs are approved for NSCLC as the initial indication, and most of them are classified as epidermal growth factor receptor (EGFR) and anaplastic lymphoma kinase (ALK) inhibitors. Therefore, EGFR mutation and ALK rearrangement tests are recommended for NSCLC before EGFR- or ALK-directed therapies [8, 9] (Fig. 2a and Table Table11).

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig2_HTML.jpg

FDA-approved therapeutic drugs for lung cancers. a Distribution of therapeutic drugs for lung cancers during the past 31 years (adapted from [126]). b Microtubule inhibitor. c EGFR inhibitors and EGFR-directed mAb. d EGFR- and MET-bispecific antibody and EGFR inhibitor for NSCLC with EGFRex20ins mutations. e ALK inhibitors. f MET inhibitors. g RET inhibitors. h Multitarget TKI. i KRASG12C-targeted small-molecule inhibitor. j DNA topoisomerase inhibitor and DNA alkylating agent for SCLC

Table 1

FDA-approved therapeutic drugs for lung cancers

Drug (brand name)SponsorPropertiesIndicationApproval dateReview

Vinorelbine tartrate

(Navelbine)

Pierre FabreMicrotubule-destabilizing agentNSCLC12/23/1994P
Gefitinib (Iressa)AstraZenecaEGFR inhibitorNSCLC05/05/2003P
Erlotinib HCl (Tarceva)OSI PharmasEGFR inhibitorNSCLC11/18/2004P
Afatinib dimaleate (Gilotrif)Boehringer IngelheimEGFR, HER2, and HER4 inhibitorMetastatic NSCLC with EGFR exon 19 deletion or exon 21 (L858R) mutation07/12/2013P, O
Osimertinib mesylate (Tagrisso)AstraZenecaEGFR inhibitorNSCLC with EGFRT790M mutations11/13/2015P, O
Dacomitinib (Vizimpro)PfizerEGFR inhibitorEGFR-mutated NSCLC09/27/2018P, O
Necitumumab (Portrazza)Eli LillyEGFR-directed mAbNSCLC11/24/2015S, O
Amivantamab (Rybrevant)Janssen BiotechEGFR- and MET-bispecific antibodyEGFR exon 20-mutated NSCLC05/21/2021P
Mobocertinib succinate (Exkivity)TakedaEGFR inhibitorEGFR exon 20-mutated NSCLC09/15/2021P, O
Crizotinib (Xalkori)Merck & CoMultitarget TKI (ALK, ROS1, and MET)ALK-positive advanced or metastatic NSCLC08/26/2011P, O
Ceritinib (Zykadia)NovartisMultitarget TKI (ALK, IGF1R, INSR, and ROS1)ALK-positive metastatic NSCLC04/29/2014P, O
Alectinib HCl (Alecensa)RocheALK inhibitorNSCLC12/11/2015P, O
Brigatinib (Alunbrig)TakedaMultitarget TKI (ALK, EGFR, IGF1R, FLT3, and ROS1)ALK-positive NSCLC04/28/2017P, O
Lorlatinib (Lorbrena)PfizerMultitarget TKI (ALK and ROS1)ALK-positive NSCLC11/02/2018P, O
Capmatinib HCl (Tabrecta)NovartisMET inhibitorNSCLC05/06/2020P, O
Tepotinib HCl (Tepmetko)Emd SeronoMET inhibitorNSCLC02/03/2021P, O
Selpercatinib (Retevmo)Loxo OncologyRET inhibitorRET fusion-positive NSCLC and thyroid cancer05/08/2020P, O
Pralsetinib (Gavreto)GenentechRET inhibitorRET fusion-positive NSCLC09/04/2020P, O
Entrectinib (Rozlytrek)GenentechMultitarget TKI (TRKs, ROS1, and ALK)NTRK fusion-positive solid tumors and ROS1-positive NSCLC08/15/2019P, O
Sotorasib (Lumakras)AmgenKRASG12C inhibitorKRASG12C-mutated NSCLC05/28/2021P, O
Topotecan HCl (Hycamtin)NovartisDNA topoisomerases inhibitorRelapsed SCLC05/28/1996P
Lurbinectedin (Zepzelca)JazzDNA alkylating drugSCLC06/15/2020P, O

ALK Anaplastic lymphoma kinase; EGFR Epidermal growth factor receptor; HER2/4 Human epidermal growth factor receptor 2/4; IGF1R Insulin-like growth factor-1 receptor; INSR Insulin receptor; NSCLC Non-small-cell lung cancer; NTRK Neurotrophic tyrosine receptor kinase; O Orphan; P Priority; RET Rearranged during transfection; ROS1 ROS proto-oncogene 1; S Standard; SCLC Small-cell lung cancer; TKI Tyrosine kinase inhibitor; TRKs Tropomyosin receptor kinases

Vinorelbine is recommended as an ingredient of systemic therapy regimens for neoadjuvant and adjuvant therapy of NSCLC. As a derivative of vinca alkaloid, it binds to tubulin in a complex with the RB3 protein stathmin-like domain (RB3-SLD), heavily overlapping the binding site of vinblastine [10, 11], thereby destabilizing α/β-tubulin heterodimers and leading to mitotic arrest and cell death [12] (Fig. 2b).

EGFR mutations occur in approximately 50% of Asian patients and 11 ~ 16% of patients in European countries with NSCLC [1315]. Exon 19 deletion and exon 21 L858R point mutation make up the majority (> 90%) of all EGFR mutation-positive NSCLC [16, 17], which frequently leads to lung tumorigenesis and sensitivity to EGFR-targeted therapies [18]. The FDA has approved six EGFR tyrosine kinase inhibitors (TKIs) which have been the first-line standard of care for patients with NSCLC harboring EGFR mutations [19]. These TKIs include the first-generation reversible EGFR inhibitors (gefitinib [20] and erlotinib [21]), the second-generation irreversible EGFR inhibitors (afatinib [22] and dacomitinib [23]), and the third-generation irreversible EGFR inhibitor (osimertinib [24]). First-generation EGFR inhibitors exert their clinical efficacy by targeting the ATP-binding pocket of the kinase domain [20, 21]. However, despite the initial response, patients almost invariably develop primary resistance to gefitinib and erlotinib and relapse after several months [25, 26]. The most common resistance mechanism is associated with the T790M ‘gatekeeper’ mutation at exon 20 of EGFR [27], which blocks reversible ATP competitive inhibitors from binding and, in turn, increases ATP binding [28]. Second-generation irreversible EGFR TKIs are highly active against the T790M point mutation of EGFR [26, 29] and exert their effect by irreversibly alkylating Cys797 and forming a covalent bond with Cys797 at the ATP-binding pocket [30], thus avoiding the increased ATP affinity conferred by the T790M gatekeeper mutation. However, EGFR T790M shares a similar ATP affinity with wild-type (WT)-EGFR, which limits the ability to achieve plasma concentrations sufficient to inhibit EGFRT790M and results in skin rash and diarrhea in patients, thereby failing to overcome T790M-mediated resistance [31]. The third-generation irreversible EGFR inhibitor osimertinib shares a similar binding mechanism with second-generation irreversible EGFR inhibitors but exhibits lower activity against EGFRWT, thereby overcoming the T790M-mediated TKI resistance [32]. As expected, osimertinib significantly prolongs median progression-free survival (PFS) by almost nine months compared with first-generation EGFR inhibitors [24]. However, acquired EGFRC797S point mutation-induced impairment in the covalent binding between EGFRCys797 and osimertinib and acquired MET amplification induced activation of the bypass pathway [33] lead to resistance to osimertinib [34]. Additionally, necitumumab is a fully human anti-EGFR IgG1κ that binds specifically to EGFR domain III, which overlaps with the EGF binding site, thereby preventing EGF ligands from binding to EGFR [35]. Thus, necitumumab was approved for first-line treatment (in combination with gemcitabine and cisplatin) for patients with metastatic squamous NSCLC [36] (Fig. 2c). Notably, necitumumab binds to most cetuximab- and panitumumab-resistant EGFR variants, such as EGFRS440L and EGFRS468R [37].

EGFR exon 20 insertion (EGFRex20ins) is clustered between codons 762–775, such as A767_V769dup (V769_D770insASV) and S768_D770dup (D770_N771insSVD) [38, 39]; it represents approximately 6 ~ 12% of EGFR mutations in NSCLC cases [4043] and frequently leads to the constitutive activation of EGFR [38]. Most EGFRex20ins driver mutations in NSCLC are insensitive to first- and second-generation EGFR inhibitors [4446], except osimertinib, which exhibits partial activity against some EGFRex20ins driver mutations in preclinical studies [39, 45, 47]. However, the clinical trials of osimertinib are inadequate and yield contradictory results [48, 49].

Amivantamab is a bispecific IgG1 that targets both EGFR and MET produced from the two purified bivalent parental antibodies by controlled Fab-arm exchange, each containing single matched point mutations in the CH3 domains (K409R and F405L) [50, 51]. The amivantamab EGFR H-arm shares an epitope identical to that of zalutumumab and binds to EGFR domain III, which overlaps with the EGF binding site, while the MET arm of amivantamab binds to the MET Sema region, which overlaps with the hepatocyte growth factor (HGF) binding site [52]. Amivantamab exhibits antitumor efficiency through the Fc-dependent antibody-dependent cellular cytotoxicity (ADCC) mechanism, Fc-independent EGFR/MET inactivation/degradation and blockade of downstream signaling transduction, and increased interferon-γ (IFNγ) secretion [44, 53, 54]. It yielded robust and durable responses with tolerable safety in patients with EGFRex20ins mutations who progressed on or after platinum-based chemotherapy [55].

Designing a novel EGFR inhibitor is another strategy to address EGFRex20ins mutations. However, the conformation of EGFRex20ins mutants largely resembles that of EGFRWT proteins because there are no amino acid substitutions in the binding site [39, 56]. Mobocertinib is an irreversible EGFR inhibitor that is structurally similar to osimertinib. It targets potential structural nuances between the EGFRex20ins and EGFRWT proteins in the vicinity of the α C-helix to gain selectivity by binding to the portions of the binding site that are not exploited by osimertinib [39]. Mobocertinib demonstrates greater activity against EGFRex20ins mutants than EGFRWT and more potent efficacy than erlotinib, gefitinib, afatinib, or osimertinib against EGFRex20ins mutants, except EGFRC797S-containing triple mutants [39, 57]. In subsequent clinical trials, mobocertinib exhibited potent activity with manageable toxicity in patients with advanced previously treated EGFRex20ins NSCLC [58, 59] (Fig. 2d).

Both aberrant ALK expression caused by ALK rearrangements [60] and ALK amplification are oncogenic driving factors of NSCLC [61]; for example, gene fusion of EMAP-like protein 4 (EML4) and ALK induced by ALK rearrangements encodes a cytoplasmic chimeric protein with constitutive kinase activity, which accounts for 3 ~ 13% of NSCLC [62, 63]. The FDA has approved five ALK inhibitors, which have been the first-line standard of care for patients with NSCLC harboring ALK rearrangements [17], including the first-generation ALK inhibitor (crizotinib [64, 65]), the second-generation ALK inhibitors (ceritinib [66], alectinib [67], and brigatinib [68]), and the third-generation ALK inhibitor (lorlatinib [69]). As with EGFR inhibitors, acquired drug resistance inevitably occurs in most patients after treatment with ALK inhibitors [70, 71]. The mechanisms of ALK inhibitor resistance also involve on-target mechanisms (e.g., ALK mutations and amplification) and off-target mechanisms and are even more complicated [72]. Approximately 20 ~ 36% of crizotinib-resistant NSCLCs harbor ALK mutations, including 1151Tins, L1152R, C1156Y, I1171T/N/S, L1196M, G1202R, S1206C/Y, E1210K, and G1269A mutations [17, 7074]. Regarding second-generation ALK inhibitors, ALK mutations account for more than half of the instances of resistance [72]. Specifically, 1151Tins, L1152P, C1156Y, F1174C/L/V, and G1202R mutations confer resistance to ceritinib [17, 72, 74], while I1171T/N/S, V1180L, L1196M, and G1202R mutations confer resistance to alectinib [72]. In addition, G1202R, D1203N, S1206Y/C, and E1210K mutations are associated with resistance to brigatinib [17, 72]. Thus, the G1202R mutation is the most common mechanism of first- and second-generation ALK inhibitor resistance. Fortunately, G1202R mutation-induced resistance can be overcome by the third-generation ALK inhibitor lorlatinib [75], which is active against the EML4-ALKG1202R mutation [76]. Intriguingly, acquired C1156Y and L1198F mutations after lorlatinib treatment resensitize the tumor to crizotinib [69]. However, the off-target mechanisms of ALK inhibitor resistance are still under exploration [77] (Fig. 2e).

Mesenchymal–epithelial transition gene (MET) exon 14 skipping mutations and MET amplification occur in approximately 3 ~ 4% [7880] and 1 ~ 6% [8183] of patients with NSCLC, respectively [84]. MET exon 14 skipping mutations produce a truncated MET with a missing regulatory domain that disrupts ubiquitin-mediated degradation, resulting in increased MET levels, sustained MET activation, and oncogenesis [85]. Thus, MET exon 14 skipping mutations and MET amplification act as oncogenic-driven factors and confer EGFR inhibitor resistance to various cancers, including NSCLC, making it a promising therapeutic target [86]. Capmatinib is a highly selective, reversible type Ib MET inhibitor that targets MET and its mutants (M1250T and Y1235D) [87, 88]. It is more potent than other MET inhibitors (approximately 30 and five times more potent than crizotinib and tepotinib in vitro, respectively) [89]. Capmatinib directly binds to the phenol moiety of the METY1230 residue, while METD1228 forms a salt bridge with METK1110 to support the Y1230–capmatinib interaction, similar to crizotinib (Type Ia MET inhibitor) [88]. Capmatinib occupies the ATP-binding site of MET, blocks MET phosphorylation, and inhibits MET-mediated downstream signaling activation [88]. Capmatinib exhibits substantial antitumor activity in patients with advanced NSCLC harboring MET exon 14 skipping mutations and MET amplification [84, 85]. Additionally, capmatinib reverses MET-dependent EGFR inhibitor resistance and blocks the signaling pathway activation mediated by EGFR and HER3 [87]. Significant resistance was observed in cells and clinical NSCLC cases bearing METD1228 and METY1230 mutations due to the structural model of the MET–capmatinib interaction [88, 90]. Tepotinib is another selective, reversible type Ib MET inhibitor for a similar clinical setting to capmatinib. Tepotinib shares a similar mechanism with capmatinib in blocking MET [91]. Thus, they achieved equivalent clinical outcomes and adverse events [84, 92, 93]. In vitro, tepotinib overlaps the most MET mutation-induced resistance with capmatinib, especially METY1230 mutations, suggesting that tepotinib may not overcome capmatinib resistance [90]. Compared with standard chemotherapy, tepotinib plus gefitinib exhibits improved antitumor activity in patients with EGFR-mutant NSCLC with MET overexpression or MET amplification [94] (Fig. 2f).

The rearranged during transfection (RET) gene rearrangements occur in approximately 1 ~ 2% of patients with NSCLC [95], which is frequently associated with brain metastases [96]. Two selective RET inhibitors (selpercatinib and pralsetinib) were approved as first-line treatments for patients with NSCLC harboring RET rearrangements [97, 98]. Selpercatinib and pralsetinib are designed to penetrate the central nervous system (CNS), thereby achieving poor CNS concentrations sufficient to maintain antitumor activity [99]. Both selpercatinib and pralsetinib exhibit activity against acquired RETV804M/L gatekeeper resistance mutations [100, 101]. However, RETG810C/S solvent front mutations (on-target) and MET amplification (off-target) were observed in selpercatinib- and pralsetinib-resistant cases [102105]. Selpercatinib and pralsetinib bind to the RET kinase in a similar mode that occupies both front and back pockets in the active site clefts without passing through the gate between V804 and K758 into the BP-I pocket [106]. This novel binding mode avoids gatekeeper V804M/L mutation-induced resistance but fails to overcome RET mutations in G810 and V738 [106]. New-generation RET inhibitors are needed for this clinical dilemma. Fortunately, selpercatinib plus crizotinib therapy may be an available strategy to overcome selpercatinib resistance in RET fusion-positive NSCLC with MET amplification [104] (Fig. 2g).

ROS proto-oncogene 1 (ROS1) rearrangements occur in approximately 1% of patients with NSCLC [107, 108]. Crizotinib has been the first-line therapy for patients with metastatic ROS1 fusion-positive NSCLC since 2016 [108, 109]. However, 47% of patients with ROS1-positive NSCLC develop brain metastases upon crizotinib treatment because of crizotinib’s poor CNS penetration due to P-glycoprotein-mediated efflux [110112]. In addition, the ROS1G2032R mutation is frequently observed in NSCLC with acquired resistance to crizotinib [113]. Entrectinib is a multitarget TKI that targets ROS1, tropomyosin receptor kinases (TRKs) (encoded by neurotrophic tyrosine receptor kinase (NTRK) genes), and ALK [114]. Compared with crizotinib, entrectinib is a weak substrate of P-glycoprotein that is 30 times more potent against ROS1, thereby overcoming P-glycoprotein-mediated efflux and achieving high CNS concentrations [112, 114, 115]. However, ROS1G2032R and ROS1F2004C/I mutations are also found in NSCLC with acquired resistance to entrectinib [116]. In addition, more mechanisms of entrectinib resistance are being identified; these include NTRK1G595R and NTRK1G667C mutations in colorectal cancer [117], NTRK3G623R mutation in mammary analog secretory carcinoma (MASC) [118], and insulin-like growth factor-1 receptor (IGF1R) activation and increased P75 expression in neuroblastoma [119]. These findings present new clinical challenges (Fig. 2h).

KRAS, one of the most frequently mutated oncogenes in various cancers, was once considered an undruggable protein due to its small size, relatively smooth surface, and rapid and tight binding properties to GTP in its active state [120]. KRASG12C is an oncogenic driver mutation that occurs in approximately 13% of patients with NSCLC [121]. Sotorasib is the first and only KRASG12C inhibitor that binds to KRASG12C via the cysteine residue mutated from the glycine residue, locking KRAS in an inactive state [120, 122, 123]. Sotorasib provides durable clinical benefits in previously treated patients with NSCLC, making it a milestone in cancer therapy [121]. Nevertheless, acquired resistance to sotorasib inevitably occurs via both on- and off-target mechanisms in most patients [124, 125]. G12C/R68S and G12C/Y96C/A double mutants and the G12D mutant of KRAS confer on-target resistance to sotorasib [124]. MET amplification is detected in sotorasib-resistant subclonal NSCLC cells with KRASG12C mutation in vitro. Thus, sotorasib plus crizotinib therapy may be a potential strategy to combat off-target resistance [125] (Fig. 2i).

Small-cell lung cancer

Small-cell lung cancer (SCLC) is a high-grade neuroendocrine carcinoma with an abysmal prognosis that accounts for approximately 15% of all lung cancer cases [126]. However, only two therapeutic drugs for SCLC have been approved by the FDA over the past 31 years. Both topotecan and lurbinectedin are approved as second-line treatments for patients with recurrent metastatic SCLC. Topotecan and irinotecan are topoisomerase I (TOP1) inhibitors and belong to alkaloid camptothecin derivatives [127]. Topotecan targets TOP1 cleavage complexes (TOP1CCs) by forming a network of hydrogen bonds with Asn722, Arg364, and Asp533 residues of TOP1 at the interface of TOP1CCs [128], thereby forming a physical impediment and blocking transcription elongation [129]. Lurbinectedin is a DNA minor groove covalent binder that binds to selected DNA triplets harboring central guanine (e.g., AGC, CGG, AGG, and TGG), resulting in the formation of a covalent adduct and inhibition of oncogenic transcription [130, 131] (Fig. 2j).

In general, genetic alterations that predict response to treatment account for approximately 30% of patients with NSCLC, including the mutations and/or rearrangements of EGFR, MET, BRAFV600E, ALK, ROS1, RET, and NTRK [109]. These approved therapeutic drugs, especially the various TKIs, provide significant clinical benefits for patients with lung cancer and other malignancies. However, overcoming the multiple mutations that induced TKI resistance and the off-target effects that induced disease progression remains challenging. As to SCLC, although the comprehensive genomic profiles have been elucidated, the majority of potential targets are undruggable. Seeking efficacious therapeutic targets and novel therapeutic strategies are still the focus of current research on this most deadly human cancer.

FDA-approved therapeutic drugs for breast cancers

Breast cancer is common in females (males only account for approximately 1% of breast cancer patients [132]). Breast cancer alone accounted for 24.5% of cancer cases and 15.5% of cancer-related deaths in women and surpassed lung cancer as one of the most commonly diagnosed cancers in 2020 [2]. Over the past 31 years, the FDA granted approvals for 24 new therapeutic drugs (including 18 small molecules, three mAbs, and three ADCs) for breast cancer, more than any other type of solid tumor [133135] (Fig. 3a and Table Table22).

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig3_HTML.jpg

FDA-approved therapeutic drugs for breast cancers. a Distribution of therapeutic drugs for breast cancers during the past 31 years (adapted from [863]). b Microtubule inhibitors. c Antimetabolite. d DNA topoisomerase inhibitor. e Aromatase inhibitors. f ER inhibitors. g CDK4/6 inhibitors. h HER2-directed mAbs. i HER2-directed ADCs. j HER2 inhibitors. k PARP inhibitor. l PI3Kα inhibitor. m Trop-2-directed ADC

Table 2

FDA-approved therapeutic drugs for breast cancers

Drug (brand name)SponsorPropertiesIndicationApproval dateReview
Docetaxel (Taxotere)SanofiMicrotubule-stabilizing agentLocally advanced or metastatic breast cancer05/14/1996P
Ixabepilone (Ixempra)R-Pharm USMicrotubule-stabilizing agentLocally advanced or metastatic breast cancer10/16/2007P
Eribulin mesylate (Halaven)EisaiMicrotubule-destabilizing agentLocally advanced or metastatic breast cancer11/15/2010P
Capecitabine (Xeloda)RocheA prodrug of 5-FUMetastatic breast cancer04/30/1998P
Epirubicin HCl (Ellence)PfizerDNA topoisomerase II inhibitorPrimary breast cancer with axillary node tumor involvement09/15/1999P, O
Anastrozole (Arimidex)Ani pharmsAromatase inhibitorAdvanced breast cancer in postmenopausal women12/27/1995S
Letrozole (Femara)NovartisAromatase inhibitorAdvanced breast cancer07/25/1997S
Exemestane (Aromasin)PfizerAromatase inhibitorER-positive early breast cancer10/21/1999S, O
Toremifene citrate (Fareston)Kyowa KirinER inhibitorER-positive metastatic breast cancer05/29/1997S, O
Fulvestrant (Faslodex)AstraZenecaER antagonistHR-positive metastatic breast cancer04/25/2002S
Palbociclib (Ibrance)PfizerCDK4/6 inhibitorHER2-negative and HR-positive advanced or metastatic breast cancer02/03/2015P
Ribociclib succinate (Kisqali)NovartisCDK4/6 inhibitorHR-positive, HER2-negative breast cancer03/13/2017P
Abemaciclib (Verzenio)Eli LillyCDK4/6 inhibitorHR-positive, HER2-negative breast cancer09/28/2017P
Trastuzumab (Herceptin)GenentechHER2-directed mAbHER2-positive breast cancer09/25/1998P
Pertuzumab (Perjeta)GenentechHER2-directed mAbHER2-positive metastatic breast cancer06/08/2012P
Margetuximab (Margenza)MacroGenicsHER2-directed mAbHER2-positive breast cancer12/16/2020S
Ado-trastuzumab emtansine (Kadcyla)GenentechHER2-directed ADCHER2-positive metastatic breast cancer02/22/2013P
Trastuzumab deruxtecan (Enhertu)Daiichi SankyoHER2-directed ADCHER2-positive breast cancer12/20/2019P
Lapatinib ditosylate (Tykerb)NovartisHER2 and EGFR inhibitorAdvanced or metastatic breast cancer03/13/2007P
Neratinib maleate (Nerlynx)Puma BiotechEGFR, HER2, and HER4 inhibitorHER2-overexpressed breast cancer07/17/2017S
Tucatinib (Tukysa)SeagenHER2 inhibitorHER2-positive breast cancer04/17/2020P, O
Talazoparib tosylate (Talzenna)PfizerPARP inhibitorBRCA-mutated HER2-negative breast cancer10/16/2018P
Alpelisib (Piqray)NovartisPI3Kα inhibitorPIK3CA-altered, HR-positive, and HER2-negative breast cancer05/24/2019P
Sacituzumab govitecan (Trodelvy)ImmunomedicsTrop-2-directed ADCTriple-negative breast cancer (TNBC)04/22/2020P

BRCA: Breast cancer susceptibility gene; CDK4/6 Cyclin-dependent kinases 4/6; ER Estrogen receptor; HER2/4 Human epidermal growth factor receptor 2/4; HR Hormone receptor; O Orphan; P Priority; PARP Poly (ADP-ribose) polymerase; PI3Kα Phosphatidylinositol 3-kinase α; S Standard; TKI Tyrosine kinase inhibitor

Cytotoxic drugs for breast cancer

Cytotoxic drugs are still widely used in clinical practice, especially in systemic chemotherapy for recurrent unresectable (local or regional) human epidermal growth factor receptor 2 (HER2)-negative breast cancer and other malignancies [133]. Among these cytotoxic drugs, docetaxel represents one of the most notable microtubule-stabilizing agents. Docetaxel shares the same taxane binding site of β-tubulin with its analog paclitaxel [136] but shows more potent antitumor activity [137]. It exerts its activity by binding to free β-tubulin and inducing microtubule polymerization, resulting in cell cycle arrest and death [138, 139]. Ixabepilone is a β-lactam analog of epothilone B and is also classified as a microtubule-stabilizing agent. It binds tubulin in a similar but not identical manner to that of paclitaxel and exhibits potent cytotoxic activity in paclitaxel-resistant cells harboring P-glycoprotein expression or mutant tubulin [140]. In contrast, eribulin is a microtubule-destabilizing agent that terminates protofilament elongation by binding predominantly to the vinca domain on β-tubulin, resulting in microtubule catastrophes [141] (Fig. 3b). Capecitabine, a prodrug of 5-fluorouracil (5-FU), is first metabolized to 5′-deoxy-5-fluorouridine (5′DFUR) by carboxylesterase and cytidine deaminase in the liver. 5′DFUR is converted to 5-FU by thymidine phosphorylase (TP) and/or uridine phosphorylase (UP). Given the significantly higher concentrations of both TP and UP in tumor tissues than in normal tissues [142144], the formation of 5-FU and the subsequent production of active metabolites, including fluorodeoxyuridine monophosphate (FdUMP), fluorouridine triphosphate (FUTP), and fluorodeoxyuridine triphosphate (FdUTP), preferentially occur in tumor tissues [145]. These metabolites finally lead to cell injury by attenuating thymidylate synthase activity (by FdUMP) and incorporating fraudulent bases into RNA (via FUTP) and DNA (via FdUTP) [146] (Fig. 3c). Epirubicin is a 4′-epimer of anthracycline antibiotic doxorubicin that exhibits at least equipotent cytotoxicity but is less myelotoxic than doxorubicin [147]. It binds to topoisomerase IIα (TOP2A), which interferes with helicase activity and TOP2A-DNA cleavable complex formation, resulting in irreversible DNA double-stranded breaks (DSBs) and gene transcription inhibition [148, 149] (Fig. 3d).

ER- or HR-positive breast cancer

Hormone receptor (HR)-positive breast cancers, including estrogen receptor (ER)- and/or progesterone receptor (PR)-positive breast cancers, account for more than 70% of all breast cancer cases [150, 151] and lead to approximately 50% of breast cancer-induced deaths [152]. Selective ER modulators (SERMs), such as tamoxifen (brand name: Nolvadex, approved on Nov. 30, 1977, by the FDA), have been the standard of care for patients with ER-positive breast cancer for over 40 years. At present, aromatase inhibitors/inactivators, SERMs, selective ER degrader/down-regulator (SERD), and cyclin-dependent kinases 4/6 (CDK4/6) inhibitors are the first-line standard of care for patients with HR-positive and HER2-negative breast cancers [153].

In premenopausal women, estrogens are mainly synthesized in the ovaries. In postmenopausal women, however, estrogens are synthesized in adipose tissue, breast, and skin, and this process is mediated by aromatase [154]. As a member of the P450 superfamily, aromatase (encoded by CYP19) is expressed at extragonadal sites, such as adipose tissue, breast, vascular tissue, bone, brain, and skin, in postmenopausal women [154, 155]. It converts androstenedione and testosterone released from ovaries and adrenal glands to estrone (E1) and E2, respectively [156]. Based on this principle, three third-generation aromatase inhibitors have been developed and approved for postmenopausal women with ER-positive breast cancer [153]. The reversible nonsteroidal aromatase inhibitors anastrozole and letrozole are triazole derivatives that exert clinical efficacy by binding to the heme prosthetic group of aromatase [157, 158]. In contrast, the irreversible aromatase inactivator exemestane binds to the substrate-binding pocket of aromatase, leading to its degradation [159161]. Among the third-generation aromatase inhibitors, letrozole exhibits the most potent inhibitory effect on aromatase enzyme activity in vivo [159, 162, 163]. It is consistently 10 ~ 30 times more potent than anastrozole in inhibiting intracellular aromatase [164]. Nevertheless, conflicting results exist in various independent studies on clinical efficacy [165167]. These contradictory results are potentially correlated with the mutation status of GATA binding protein 3 (GATA3) [168] or the saturation effect (all third-generation aromatase inhibitors reproducibly cause ~ 98% aromatase inhibition in humans) [161] (Fig. 3e).

Toremifene is a SERM structurally similar to tamoxifen, differing only by a single chlorine atom [169]. Like tamoxifen, toremifene exerts pharmacological activity by competitively inhibiting estradiol (E2) binding to the ER. It thus cannot be used as second-line therapy after tamoxifen failure due to similar pharmacological mechanisms [170]. In contrast, fulvestrant is a full ER antagonist approved as a SERD that overcomes the agonistic effects of tamoxifen and toremifene [171, 172]. However, because of its poor physicochemical features, fulvestrant must be administered monthly intramuscular injections, limiting its clinical application [173]. Mechanistically, it has recently been proven to exert its properties by markedly impairing the intranuclear mobility of the ER [152] (Fig. 3f).

The formation of the cyclin D–cyclin-dependent kinases 4/6 (CDK4/6) complex (also known as G1-CDK) and CDK4/6-induced retinoblastoma (RB) phosphorylation are core events of the G1-S transition in the cell cycle [174]. Inhibition of CDK4/6 induces RB hypophosphorylation and reactivation, resulting in stable cell cycle arrest in the G1 phase [175]. Three CDK4/6 inhibitors (palbociclib, ribociclib, and abemaciclib) have been approved for the first-line therapy of patients with HR-positive and HER2-negative breast cancers in combination with nonsteroidal aromatase inhibitors [176179] or SERD (fulvestrant) [180, 181] (Fig. 3g), thereby delaying or overcoming endocrine resistance [182]. Although three CDK4/6 inhibitors share multiple similarities, unique characteristics exist in each of them [182]. Palbociclib primarily targets CDK4 monomers instead of endogenous CDK4 trimer complexes or CDK6 but promotes the formation of inactive CDK2 complexes [183]. Palbociclib and ribociclib are more selective for CDK4/6 than abemaciclib, probably due to the greater lipophilicity and larger binding site side chains than abemaciclib, which may reduce the probability of interaction with off-target kinase ATP-binding pockets [184, 185]. Ribociclib is less potent than palbociclib and abemaciclib in inhibiting RB phosphorylation [184]. In contrast, abemaciclib binds to the ATP cleft more readily and forms a hydrogen bond with the conserved catalytic residue (Lys43) of CDKs, which decreases its selectivity [184, 185].

HER2-positive breast cancer

HER2-positive breast cancer (including some luminal B subtype cancers) accounts for 13 ~ 15% of all breast cancer cases [186] and is associated with aggressive and metastatic behavior [187]. As the first mAb to be approved to treat solid tumors, trastuzumab is a landmark in tailored therapies [133]. Trastuzumab binds to the extracellular region of HER2 on the C-terminal portion of domain IV and exerts its function via several mechanisms, including ADCC, inhibition of HER2 shedding, and disruption of ligand-independent downstream cascades [188191]. However, trastuzumab is insufficient to block ligand-induced HER2/HER3 dimerization [191]. In contrast, pertuzumab binds to the extracellular domain II of HER2 and blocks both ligand-dependent and ligand-independent HER2/HER3 dimerization and activation [191193]. The addition of pertuzumab to the combination of trastuzumab plus docetaxel significantly improves median PFS, and overall survival (OS) compared to that with a pertuzumab-free regimen [194, 195]. Margetuximab, as the latest approved HER2 mAb, improves the ADCC effect in HER2-low tumors with enhanced targeting activity and overcomes trastuzumab resistance [196] (Fig. 3h). Compared with trastuzumab plus chemotherapy, margetuximab plus chemotherapy significantly improves PFS in HER2-positive patients who have received two or more prior anti-HER2 therapies [197].

Trastuzumab significantly improves the clinical outcomes of patients with HER2-positive breast cancer [198]. In the metastatic setting, however, resistance to trastuzumab and disease progression occurs in most patients treated with trastuzumab within one year [199, 200]. The general mechanisms of trastuzumab resistance refer to obstacles for trastuzumab–HER2 interaction, reactivation of HER2 downstream signaling pathways, initiation of bypass signaling pathways, and failure to trigger immune-mediated mechanisms [201]. For this reason, two HER2-based ADCs (ado-trastuzumab emtansine and trastuzumab deruxtecan) were introduced (Fig. 3i). Ado-trastuzumab emtansine is composed of trastuzumab and DM1, linked with a non-cleavable thioether linker, N-succinimidyl-4-(N-maleimidomethyl) cyclohexane-1-carboxylate (SMCC, designated MCC after conjugation) [202, 203]. DM1 is a derivative of maytansine isolated from various Maytenus species [204] that exerts antitumor activity by destabilizing microtubules [205]. Ado-trastuzumab emtansine retains all the antitumor efficiency of trastuzumab and is active against lapatinib-resistant breast cancer cells and lapatinib-insensitive tumors [203]. Ado-trastuzumab emtansine shows significant clinical advantages over lapatinib plus capecitabine [206] and trastuzumab plus docetaxel [207]. Despite these therapeutic advances, most patients treated with ado-trastuzumab emtansine eventually experience disease progression [208, 209]. The resistance mechanisms of ado-trastuzumab emtansine partially overlap with those of trastuzumab but also include P-glycoprotein overexpression and receptor-mediated endocytosis defects [210, 211]. Trastuzumab deruxtecan is composed of trastuzumab and TOP1 inhibitor payload (Dxd, an exatecan derivative) linked with a protease-cleavable maleimide tetrapeptide linker [212]. Trastuzumab deruxtecan exhibits durable antitumor activity in patients previously treated with ado-trastuzumab emtansine [213]. Recently, the phase 3 DESTINY-Breast03 trial (NCT03529110) demonstrated that trastuzumab deruxtecan exhibits superiority over trastuzumab emtansine in patients previously treated with the trastuzumab plus taxane regimen [214].

Moreover, three HER2 inhibitors (lapatinib, neratinib, and tucatinib) were approved as third-line regimens for the treatment of HER2-positive breast cancer in combination with trastuzumab and/or capecitabine [215] (Fig. 3j). In contrast to HER2-directed mAbs, HER2 inhibitors bind to the cytoplasmic tyrosine kinase domain instead of the extracellular region of HER2. Lapatinib is a potent dual inhibitor of both EGFR and HER2 [216, 217] that exerts antitumor activity by reversibly binding to the cytoplasmic ATP-binding sites of EGFR and HER2, leading to the impediment of tyrosine kinase phosphorylation, which dampens or abrogates the activation of HER2-mediated downstream pathways [218]. Intriguingly, lapatinib also reverses P-glycoprotein- and ABCG2-mediated multidrug resistance (MDR) by directly attenuating their transport activity [219]. However, HER2T798M/I gatekeeper mutations and bypass signaling pathway initiation inevitably confer resistance to lapatinib [220222]. Neratinib is an irreversible inhibitor of EGFR, HER2, and HER4 that binds to the conserved Cys773 of EGFR and Cys805 of HER2, which forms a covalent bond with the HER family at the cleft of the ATP-binding site [223, 224]. Neratinib exhibited substantial clinical activity in patients with and without prior trastuzumab treatment [225], while the neratinib plus paclitaxel regimen yielded higher complete pathological response rates than the trastuzumab plus paclitaxel regimen in patients with HER2-positive, HR-negative breast cancer [226]. However, the HER2T798I gatekeeper mutation also confers resistance to neratinib [227], suggesting that neratinib may not overcome lapatinib resistance, although it displays nanomolar antiproliferative activity against this mutant in vitro [228]. Tucatinib is another reversible HER2 inhibitor that shares a similar binding mechanism with lapatinib but exhibits the highest selectivity to HER2 among these HER2 inhibitors [229]. Compared to placebo, tucatinib’s addition to the trastuzumab plus capecitabine regimen exhibited acceptable toxicity [230], improved survival outcomes, improved objective response rate (ORR), and reduced the risk of death [231, 232]. Among these HER2 inhibitors, neratinib exhibits the most potent activity against HER2 kinase, followed by tucatinib and lapatinib [228].

BRCA-mutated breast cancer

Germline mutations of BRCA1 and/or BRCA2 are observed in more than 5% of all breast cancer cases and approximately 13% of basal-like breast cancer (BLBC) cases [233]. BRCA1/2 mutations frequently indicate a deficiency in repairing DNA DSBs by homologous recombination [234] and predispose patients to breast, ovarian, and other cancers [235237]. Poly (ADP-ribose) polymerases (PARPs) are essential for DNA single-strand break (SSB) repair by base excision repair (BER) [238]. The N-terminal zinc finger motifs of PARPs bind to damaged DNA, which activates its catalytic C-terminal to hydrolyze nicotinamide adenine dinucleotide (NAD+) and produce ADP-ribose units, thereby yielding linear and branched poly (ADP-ribose) (PAR) for the resealing of DNA SSBs during BER [239, 240]. PARP inhibitors are designed to inhibit auto-PARylation by competitively binding to PARPs at the NAD+ binding site [241, 242], leading to cell death in BRCA1/2-mutated cancer cells through a synthetic lethality mechanism [243]. Breast cancers harboring germline mutations in either BRCA1 or BRCA2 are highly sensitive to PARP inhibitors [244, 245], and thus, inhibiting PARPs has become a therapeutic strategy for targeting BRCA1/2-mutated cancer cells [246]. Talazoparib is the fourth (also the latest) PARP inhibitor approved by the FDA (after olaparib, rucaparib, and niraparib) [247]. Through hydrogen-bonding and π-stacking interactions, including those mediated by active site water molecules, talazoparib is anchored to the nicotinamide-binding pocket [248], leading to a noticeable displacement of the bound ligand within the NAD+ site [249]. Compared with olaparib, rucaparib, and niraparib (IC50 values 1.94, 1.98, and 3.8 nM, respectively, for the inhibition of PARP1), talazoparib is three times more potent, with an IC50 of 0.57 nM [250, 251]. Therefore, talazoparib exhibits superiority over olaparib or rucaparib in trapping PARP–DNA at the site of DNA damage [239]. Talazoparib monotherapy demonstrates a tolerable safety profile and preliminary clinical activity in patients with sporadic cancers harboring germline BRCA1/2 mutations [252]. It also exhibits a significant benefit over standard chemotherapy (capecitabine, eribulin, gemcitabine, or vinorelbine) among patients with germline BRCA1/2-mutated breast cancer [253] (Fig. 3k).

PIK3CA-altered breast cancer

Phosphatidylinositol 3-kinase catalytic subunit A (PIK3CA) gene mutation is observed in approximately 40% of HR-positive and HER2-negative breast cancers [233, 254, 255]. PIK3CA mutation induces phosphatidylinositol 3-kinase (PI3K) activation, leading to cell proliferation and apoptosis evasion [233]. Alpelisib is a PI3Kα inhibitor that binds to PI3Kα and forms multiple hydrogen bonds with PI3Kα at the ATP-binding pocket, thereby inhibiting the enzymatic activity of PI3Kα and PI3Kα-mediated downstream pathways [256, 257]. Alpelisib demonstrates tolerable safety and favorable clinical efficiency in patients with PIK3CA-altered, HR-positive, HER2-negative breast cancer in combination with fulvestrant [255, 258260] or letrozole [261] (Fig. 3l).

Triple-negative breast cancer

Triple-negative breast cancer (TNBC) is defined as breast cancer lacking expression of ER, PR, and HER2, which accounts for 10 ~ 15% of all breast cancer cases [186]. It represents the subtype with the worst prognostic outcome among breast cancers [262]. Before 2019, single-agent taxanes or anthracyclines were the first-line regimens for unresectable locally advanced or metastatic TNBC [263]. However, the median OS remains at approximately 18 months or even less [264, 265]. On March 8, 2019, the FDA-approved atezolizumab plus albumin-bound nab-paclitaxel (brand name: Abraxane, approved on January 7, 2005) as a first-line regimen for unresectable locally advanced or metastatic TNBC with PDL1 expression [265, 266]. Trophoblastic cell surface antigen-2 (Trop-2, also known as EGP-1, encoded by TACSTD2) is a transmembrane glycoprotein overexpressed in 83% of breast cancer cases [267] and 85% of TNBCs [268]. It is considered a key driver of human cancers, making it an attractive target for TNBC treatment [267]. Sacituzumab govitecan-hziy is a Trop-2-directed ADC composed of sacituzumab and SN-38 covalently linked with a hydrolyzable CL2A linker [269, 270]. Safituzumab is a humanized Trop-2-directed mAb developed from murine RS7-3G11 [271, 272], while SN-38 is an active metabolite of irinotecan, a TOP1 inhibitor [273]. Sacituzumab govitecan-hziy exhibits acceptable toxicity and preliminary clinical activity in previously treated patients with refractory metastatic solid tumors [274], especially with metastatic TNBC [275, 276]. Compared with standard chemotherapy, it demonstrates durable objective responses and significant superiorities in heavily treated patients with metastatic TNBC [277]. However, the clinical benefits of sacituzumab govitecan-hziy are highly dependent on Trop-2 expression; definitive conclusions are difficult to draw in the low Trop-2 expression subgroup [278]. In addition, canonical TOP1E418K resistance mutation, TOP1p.-122 fs (frameshift mutation), and TACSTD2T256R missense mutation confer resistance to sacituzumab govitecan-hziy [268]. These findings pose new challenges regarding sacituzumab govitecan-hziy application (Fig. 3m).

Breast cancer drugs are frequently at the forefront of advances in cancer treatment and diagnosis, especially in CDK4/6 inhibitors, HER2 inhibitors, and HER2-directed mAbs and ADCs [133]. Meanwhile, the progress of breast cancer drugs provides an essential basis for other malignancies in drug research and development. Cytotoxic drugs and selective ER antagonists dominated the early decades until 2010. However, these two types of medicines have been overshadowed by targeted drugs, which have accounted for the majority of the newly approved breast cancer drugs since 2010. Besides, the approved drugs mainly focused on targeting HER2 in recent years, limiting the breakthrough in drug development, especially for TNBCs.

FDA-approved therapeutic drugs for gynecologic cancers

Gynecologic cancers include cervical, ovarian, uterine, vaginal, vulvar, and fallopian tube cancers, accounting for 15.2% of all malignancies among females and 15.3% of cancer-related deaths worldwide in 2020 [2]. However, only six therapeutic drugs have been approved by the FDA for gynecologic cancers as initial indications since 1991 (Fig. 4a and Table Table33).

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig4_HTML.jpg

FDA-approved therapeutic drugs for gynecologic cancers. a Distribution of therapeutic drugs for gynecologic cancers during the past 31 years (adapted from [864]). b Microtubule inhibitor. c PARP inhibitors. d PD1-directed mAb. e TF-targeted ADC

Table 3

FDA-approved therapeutic drugs for gynecologic cancers

Drug (brand name)SponsorPropertiesIndicationApproval dateReview
Paclitaxel (Taxol)HQ SpcltMicrotubule-stabilizing agentAdvanced ovarian cancer12/29/1992P, O
Olaparib (Lynparza)AstraZenecaPARP inhibitorAdvanced BRCA-mutated ovarian cancer12/19/2014P, O
Rucaparib camsylate (Rubraca)Clovis OncologyPARP inhibitorBRCA-positive ovarian cancer12/19/2016P, O
Niraparib tosylate (Zejula)GlaxoSmithKlinePARP inhibitorEpithelial ovarian, fallopian tube, or primary peritoneal cancer03/27/2017P, O
Dostarlimab (Jemperli)GlaxoSmithKlinePD1-directed mAbEndometrial cancer04/22/2021P
Tisotumab vedotin (Tivdak)SeagenTF-targeted ADCCervical cancer09/20/2021P, O

O Orphan; P Priority; PARP Poly (ADP-ribose) polymerase; PD1 Programmed death receptor-1; S Standard; TF Tissue factor

Ovarian cancer

Ovarian cancer is the third most common gynecologic cancer, accounting for 3.4% of all female malignancies and 4.7% of cancer-related deaths in females worldwide in 2020 [2]. FDA has granted four new therapeutic drug approvals for ovarian cancer. Paclitaxel is undoubtedly a milestone in the history of cancer drugs. It was isolated by Wall and Wani from the bark of Taxus brevifolia in 1971 [279] (Fig. 4b). Currently, paclitaxel (including nab-paclitaxel albumin-bound) and its analog docetaxel are widely used to treat various malignancies [280]. It covalently binds to β-tubulin at amino acid residues 1–31 [281], 217–233 [282], and Arg282 [283] and enhances microtubule polymerization, thereby suppressing microtubule dynamics and blocking cell mitosis [284].

Germline mutations of BRCA1 and/or BRCA/2 are present in approximately 14.1% of all ovarian cancer cases [285]. Based on the same principle described in the breast cancer section above, three PARP inhibitors (olaparib [286], rucaparib [287], and niraparib [250]) were approved as maintenance therapies for BRCA1/2-mutated ovarian cancer [288290] (Fig. 4c). These PARP inhibitors are designed to competitively bind to the NAD+ binding site of the PARP enzyme [237, 238]. Platinum and PARP inhibitor sensitivity commonly coexist in BRCA1/2-mutated ovarian cancer due to homologous recombination deficiency (HRD) [291]; however, nucleotide excision repair (NER) alterations confer enhanced platinum sensitivity but not PARP inhibitor sensitivity [292]. There is no significant efficacy difference between these PARP inhibitors as maintenance therapies in patients with BRCA-mutated, platinum-sensitive relapsed ovarian cancer [293]. Additionally, PARP inhibitors yield similar response and survival rates in patients harboring either somatic or germline BRCA mutations [294]. Of note, olaparib represents the most cost-effective [295] PARP inhibitor, and the olaparib plus bevacizumab regimen achieved a dramatic improvement in PFS in ovarian cancer patients with BRCA mutations (37.2 months) and without BRCA mutations (28.1 months) compared to that with placebo plus bevacizumab (17.7 and 16.6 months, respectively) [296]. Thus, the olaparib plus bevacizumab regimen was approved for first‐line maintenance treatment of HRD-positive advanced ovarian cancer [297]. Clinical trials of PARP inhibitors (rucaparib and niraparib) combined with bevacizumab for ovarian cancer maintenance therapy are still ongoing [298, 299].

Endometrial cancer

Endometrial cancer is the second most common gynecologic cancer and originates in the inner epithelial lining of the uterus [300]. It accounted for 4.5% of all malignancies among females and 2.2% of cancer-related deaths in females worldwide in 2020 [2]. Mismatch repair deficiency (dMMR) is a consequence of germline mutations or epigenetic silencing in MMR genes, resulting in the accumulation of errors introduced during DNA replication [301]. Therefore, dMMR leads to genome-wide instability, especially in regions of simple repetitive DNA sequences (known as microsatellite instability—high (MSI-H)), resulting in tumorigenesis [302]. MSI-H/dMMR is observed in 18 ~ 28% of endometrial cancer cases [303, 304] and confers sensitivity to PD1 blockade [304]. A higher number of CD3+ and CD8+ TILs and increased PD1 expression (but not PDL1) are observed in the hypermutated subgroups (POLE mutations or MSI-H/dMMR) of endometrial cancer than in the hypomutated microsatellite-stable subgroup [305], explaining why MSI-H/dMMR-positive endometrial cancer is sensitive to PD1 blockade. Dostarlimab is the fourth (also the latest) FDA-approved PD1-directed mAb after pembrolizumab, nivolumab, and cemiplimab [306]. It exhibits a high affinity for both human and cynomolgus monkey PD1, preventing PDL1 and PDL2 from interacting with PD1 [307]. Dostarlimab demonstrates a manageable safety profile equivalent to that of other PD1-directed mAbs and robust clinical activity in previously treated patients with recurrent or advanced MSI-H/dMMR or MMR proficient/stable (MMRp/MSS) endometrial cancer [308, 309] (Fig. 4d). Of note, dostarlimab achieved a complete response in 100% of patients with dMMR-positive locally advanced rectal cancer [310].

Cervical cancer

Cervical cancer is the most common gynecologic cancer, accounting for 6.6% of all malignancies among females and 7.8% of cancer-related deaths in females worldwide in 2020 [2]. Cervical cancer is strongly linked with human-papillomavirus (HPV) infection [311], especially HPV-16 and HPV-18 subtypes [312]. Tissue factor (TF, also known as thromboplastin, factor III, or CD142) is overexpressed in various cancers [313], especially cervical cancer [314]. TF promotes tumor progression by initiating the coagulation pathway with its procoagulant activity and protease-activated receptor 2 (PAR-2)-mediated signaling, making it an attractive target [315]. Tisotumab vedotin is a TF-directed ADC composed of tisotumab and microtubule-destabilizing agent monomethyl auristatin E (MMAE), linked with protease-cleavable maleimidocaproyl valine-citrulline p-aminobenzyl alcohol carbamate (MC-vc-PAB) linker [316]. Tisotumab is a TF-directed mAb generated by immunization of HuMAb mice [316], while vedotin refers to MMAE plus the MC-vc-PAB linker. Tisotumab vedotin demonstrates a manageable safety profile and durable antitumor activity in previously treated (e.g., bevacizumab plus doublet chemotherapy) patients with recurrent or metastatic cervical cancer [317, 318] (Fig. 4e).

FDA-approved therapeutic drugs for gastrointestinal cancers

Gastrointestinal cancers include esophageal, gastric, colorectal, pancreatic, gallbladder, and liver cancer (including cholangiocarcinoma), accounting for 26.4% of cancer cases and 36.3% of cancer-related mortalities worldwide in 2020 [2]. Over the past 31 years, the FDA granted approvals for 17 new therapeutic drugs (including 12 small molecules, four mAbs, and one recombinant fusion protein) for gastrointestinal cancers (Fig. 5a and Table Table44).

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig5_HTML.jpg

FDA-approved therapeutic drugs for gastrointestinal cancers. a Distribution of therapeutic drugs for gastrointestinal cancers during the past 31 years (adapted from [865, 866]). b Photosensitizer. c VEGFR2-directed mAb. d Multitarget TKI and PDGFR inhibitors. e Somatostatin receptor-targeted radiopharmaceutical. f FGFR inhibitors. g DNA synthesis inhibitor. h DNA topoisomerase inhibitor. i Organoplatinum alkylating agent. j EGFR-directed mAb. k VEGF-A-directed mAb. l Soluble receptor decoy that binds VEGF-A, VEGF-B, and PlGF. m Multitarget TKI. n Thymidine phosphorylase inhibitor plus nucleoside metabolic inhibitor

Table 4

FDA-approved therapeutic drugs for gastrointestinal cancers

Drug (brand name)SponsorPropertiesIndicationApproval dateReview
Porfimer sodium (Photofrin)Pinnacle BiolgsA photosensitizer used for photodynamic therapyObstructing esophageal cancer12/27/1995P, O
Ramucirumab (Cyramza)Eli LillyVEGFR2-directed mAbGastric cancer04/21/2014P, O
Sunitinib malate (Sutent)CPPI CVMultitarget TKI (VEGFRs, PDGFRα/β, CSF1R, KIT, and FLT3)Imatinib-resistant GIST and advanced RCC01/26/2006P
Avapritinib (Ayvakit)BlueprintPDGFRα, PDGFRα mutants, and KIT inhibitorGIST with PDGFRA exon 18 mutations01/09/2020P, O
Ripretinib (Qinlock)DecipheraPDGFRα, PDGFRα mutants, and KIT inhibitorAdvanced GIST05/15/2020P, O
Lutetium Lu-177 dotatate (Lutathera)AAA USASomatostatin receptor-targeted radiopharmaceuticalGEP-NETs01/26/2018P, O
Pemigatinib (Pemazyre)IncyteFGFR1-3 inhibitorAdvanced cholangiocarcinoma with FGFR2 fusions/rearrangements04/17/2020P, O
Infigratinib phosphate (Truseltiq)Helsinn HlthcareFGFR1-3 inhibitorAdvanced cholangiocarcinoma with FGFR2 fusions/rearrangements05/28/2021P, O
Gemcitabine HCl (Gemzar)Eli LillyDNA synthesis inhibitorLocally advanced or metastatic pancreatic cancer05/15/1996P
Irinotecan HCl (Camptosar)PfizerDNA topoisomerase I inhibitorMetastatic colorectal cancer06/14/1996P
Oxaliplatin (Eloxatin)SanofiOrganoplatinum alkylating agentColorectal cancer (in combination with 5-FU and leucovorin)08/09/2002P
Cetuximab (Erbitux)ImCloneEGFR-directed mAbColorectal cancer02/12/2004N/A
Panitumumab (Vectibix)AmgenEGFR-directed mAbColorectal cancer09/27/2006N/A
Bevacizumab (Avastin)GenentechVEGF-A-directed mAbColorectal cancer02/26/2004O
Ziv-aflibercept (Zaltrap)SanofiSoluble receptor decoy that binds VEGF-A, VEGF-B, and PlGFMetastatic colorectal cancer08/03/2012N/A
Regorafenib (Stivarga)BayerMultitarget TKI (RET, VEGFRs, KIT, PDGFRα/β, FGFR1/2, RAF1, BRAF, and BRAFV600E)Metastatic colorectal cancer09/27/2012P
Tipiracil HCl; Trifluridine (Lonsurf)TaihoThymidine phosphorylase inhibitor plus a nucleoside metabolic inhibitorColorectal cancer09/22/2015S

EGFR Epidermal growth factor receptor; FGFR1-3 Fibroblast growth factor receptor-1–3; GEP-NET Gastroenteropancreatic neuroendocrine tumor; GIST Gastrointestinal stromal tumor; O Orphan; P Priority; PDGFRα/β Platelet-derived growth factor receptor α/β; PlGF Placenta growth factor; RCC Renal cell carcinoma; RET Rearranged during transfection; S Standard; VEGF Vascular endothelial growth factor; VEGFR Vascular endothelial growth factor receptor

Esophageal cancer

Esophageal cancer accounted for 3.1% of cancer cases and 5.5% of cancer-related mortalities worldwide in 2020 [2]. Porfimer sodium was approved by the FDA as a photosensitizer for photodynamic therapy of obstructing esophageal cancer [319] (Fig. 5b). In the presence of oxygen, this approach utilizes light to activate the porfimer sodium, which is relatively selectively concentrated in cancer cells, leading to cell death [320]. However, photodynamic therapy with porfimer sodium as an endoscopic therapy for esophageal cancer is losing popularity due to the potential for long-term complications [321]. Fluoropyrimidine plus platinum-based chemotherapies are frequently used as first-line therapy for advanced esophageal cancer [322]. Compared to chemotherapy alone, pembrolizumab plus 5-FU and cisplatin (chemotherapy) significantly improve clinical outcomes in the first-line treatment of advanced esophageal cancer [323].

Gastric cancer

Gastric cancer accounted for 5.6% of cancer cases and 7.7% of cancer-related mortalities worldwide in 2020 [2]. Vascular endothelial growth factor receptor 2 (VEGFR2), the principal receptor of VEGF-induced angiogenesis, is expressed in most solid tumors, including gastric cancer [324]. Ramucirumab is a VEGFR2-directed mAb [325] that binds selectively to the g-like extracellular domain III of VEGFR2, which prevents VEGF ligands from binding to VEGFR2 [326], thereby inhibiting VEGF ligand-induced cell proliferation, migration, and angiogenesis [327] (Fig. 5c). Ramucirumab monotherapy exhibits significant survival benefits in patients with advanced gastric or gastroesophageal junction adenocarcinoma who have disease progression after first-line chemotherapy compared to placebo [328]. The ramucirumab plus paclitaxel regimen also demonstrated superiority over placebo plus paclitaxel therapy in the same clinical setting; thus, ramucirumab plus paclitaxel could be regarded as a new second-line treatment for advanced gastric cancer [329].

Gastrointestinal stromal tumors

Gastrointestinal stromal tumors (GISTs) constitute the largest subset of mesenchymal tumors that arise from precursors of the connective tissue cells of the gastrointestinal tract [330, 331]. They occur predominantly (60%) in the stomach, with 30% of cases in the small intestine and 10% of cases in other sites of the gastrointestinal tract; 10 ~ 30% are malignant and exhibit intra-abdominal spread or liver metastases [332]. RTKs, such as VEGFR2, platelet-derived growth factor receptor α/β (PDGFRα/β), and KIT, are frequently overexpressed or mutated in GISTs, leading to constitutive activation of these kinases [333, 334]. Approximately 75 ~ 80% of GISTs harbor KIT mutations, and 5 ~ 8% of GISTs harbor PDGFRA mutations [334]. Therefore, the FDA-approved three multitarget TKIs for GIST treatment (Fig. 5d).

Imatinib (Additional file 1: Table S1, page 13; Table S2, page 44) is still the first-line treatment for advanced GISTs [335, 336]. However, approximately 50% of patients develop resistance within two years [336, 337]. Sunitinib is a potent inhibitor of multiple RTKs, including PDGFRα/β, VEGFR2, and KIT [338], and has been approved as second-line therapy for imatinib-resistant GISTs [339, 340]. The ATP-binding-pocket mutants KITV654A, KITT670I, and PDGFRαD842V are the most common in imatinib-resistant GISTs, whereas certain mutant-induced resistance can be overcome by sunitinib, except PDGFRαD842V [337, 340, 341]. Given the failures in overcoming the PDGFRAD842V-induced resistance, avapritinib was approved as a first-line regimen for GISTs harboring PDGFRA exon 18 (including D842V) mutation [342]. Avapritinib is a potent TKI that targets KIT exon 17 (including D816V) and PDGFRA exon 18 (including D842V) mutations. In contrast, imatinib, sunitinib, and regorafenib exhibit weak potency in blocking mutation-induced constitutive kinase activity [343, 344]. Given the heterogeneity of KIT and PDGFRA mutants in GISTs, broader spectrum drugs are needed to overcome the multiple mutations of KIT and PDGFRA, as well as other RTKs. Ripretinib was designed to overcome the drug resistance of GISTs harboring broad KIT and PDGFRA mutations [345]. As a ‘switch control’ kinase inhibitor, ripretinib forces the activation loop of KIT or PDGFRα into an inactive conformation through a switch control mechanism that prevents switches from adopting a type I active state and stabilizes switches in type II inactive state [345, 346]. Therefore, the FDA-approved ripretinib for the fourth-line treatment of patients with advanced GIST who have received prior treatment with three or more TKIs [345]. Notably, the common PDGFRαD842V mutant is sensitive to avapritinib and crenolanib but resistant to ripretinib, and secondary resistance mutations after imatinib or avapritinib treatment, such as the triple mutant PDGFRαD842V/V658A/G652E, can be overcome by the heat shock protein 90 (HSP90) inhibitor tanespimycin [347].

Gastroenteropancreatic neuroendocrine tumors

Gastroenteropancreatic neuroendocrine tumors (GEP-NETs) account for more than 60% of NETs that arise from neuroendocrine cells of the digestive tract [348]. Regarding prevalence, GEP-NETs have been the second most common gastrointestinal cancer [349]. Somatostatin receptors (SSTRs) are G-protein-coupled receptors frequently expressed in GEP-NETs [350]. Somatostatin is the ligand of SSTRs that inhibits the release of pituitary and gastrointestinal hormones [351]. Octreotide (brand name: Sandostatin, approved by the FDA on Oct. 21, 1988), a synthetic octapeptide (D-Phe-c[Cys-Phe-D-Trp-Lys-Thr-Cys]-Thr-ol), is a somatostatin analog with long-acting pharmacologic properties mimicking natural somatostatin [352]. Therefore, it has been approved for metastatic carcinoid and vasoactive intestinal peptide-secreting tumors [353]. While it does not affect tumor progression, it can improve symptoms. On the other hand, Lutetium-177 (177Lu) is a medium-energy β- and low-energy γ-emitting radionuclide with a maximal tissue penetration of 2 mm [354] and a half-life of 160 h [355], allowing detection by scintigraphy and subsequent dosimetry. Combining the properties of 177Lu and octreotate (differs from octreotide only in that the C-terminal threoninol is replaced with threonine but exhibits a higher affinity for SSTR2 than octreotide [356]), [177Lu-DOTA0,Tyr3]-octreotate (Lutetium Lu-177 dotatate) was approved by the FDA for peptide receptor radionuclide therapy (PRRT) of SSTR-positive advanced GEP-NETs [349] (Fig. 5e).

Cholangiocarcinoma

Hepatocellular carcinoma (HCC, comprising 75% ~ 85% of liver cancer cases) and intrahepatic cholangiocarcinoma (ICC, comprising 10 ~ 15% of liver cancer cases) are the most frequent types of primary liver cancer, which accounted for 4.7% of cancer cases and 8.3% of cancer-related mortalities worldwide in 2020 [2]. Compared to HCC, ICC has a poorer prognosis in terms of both mOS (HCC 71.7 months vs. ICC 21.5 months) and disease-free survival (DFS) (HCC 68.2 months vs. ICC 15.5 months) [357]. Genomic alterations (including mutation, fusion, and rearrangement) that activate fibroblast growth factor receptor 2 (FGFR2) are almost exclusively found in patients with ICC, making it a promising therapeutic target [358, 359].

Pemigatinib is a potent, selective inhibitor of FGFR1-3 that binds the ATP-binding pocket of FGFR at the hinge region, thereby inhibiting FGFR-mediated cell proliferation, differentiation, and angiogenesis [360, 361]. Pemigatinib exhibits a manageable safety profile and durable antitumor activity in previously treated patients with cholangiocarcinoma harboring FGFR2 fusions/rearrangements [362]. A phase 3 FIGHT-302 clinical trial of first-line pemigatinib vs. gemcitabine plus cisplatin for advanced cholangiocarcinoma harboring FGFR2 fusions/rearrangements is still ongoing [363]. Similar to all other TKIs, acquired resistance mutations in FGFR2 (N549K/H, E565A, L617V, K641R, and K659M) are observed in patients with progressive disease and may confer resistance to pemigatinib [359]. Infigratinib is another FGFR1-3 inhibitor [364] that binds to FGFR at a hinge region, similar to pemigatinib [365]. It shows manageable toxicity and meaningful clinical activity against chemotherapy-refractory cholangiocarcinoma FGFR2 fusions/rearrangements [366, 367]. Strikingly, 5 of 6 FGFR2 mutations observed in infigratinib-resistant patients completely overlapped with the five FGFR2 mutations observed in pemigatinib-resistant cases, except for the FGFR2V564F gatekeeper resistance mutation, which exclusively exists in infigratinib-resistant patients [359, 368, 369]. Thus, pemigatinib may theoretically overcome FGFR2V564F mutation-induced resistance to infigratinib (Fig. 5f).

Pancreatic cancer

Pancreatic cancer has the highest mortality-to-incidence ratio (1.808) among all malignancies, accounting for 2.6% of cancer cases and 4.7% of cancer-related mortalities worldwide in 2020 [2]. Although genomic and microenvironment alterations of pancreatic cancer have been elucidated [370, 371], most alterations (e.g., KRAS and TP53 mutations) are not druggable. Given the lack of effective targets, systemic chemotherapy is still the first-line regimen. Gemcitabine is an analog of deoxycytidine that acts as a DNA synthesis inhibitor. It is phosphorylated by deoxycytidine kinase to form its active products (including gemcitabine diphosphate and gemcitabine triphosphate), which are incorporated into the DNA, leading to the inhibition of the DNA synthesis process [372] (Fig. 5g). Gemcitabine exhibits significant superiority over 5-FU in patients with advanced pancreatic cancer [373]. Currently, systemic chemotherapy combinations, including FOLFIRINOX (5-FU, leucovorin, irinotecan, and oxaliplatin) [374] and gemcitabine plus nab-paclitaxel [375], have become the first-line treatment for patients with advanced pancreatic cancer [376]. There was no significant difference in the treatment efficacy between the FOLFIRINOX and gemcitabine plus nab-paclitaxel regimens [377].

Colorectal cancer

Colorectal cancer accounted for 9.8% of cancer cases and 9.2% of cancer-related mortalities worldwide in 2020 [2]. The FDA-approved eight therapeutic drugs in the past 31 years (Fig. 5h–n). Similar to topotecan, irinotecan is also a TOP1 inhibitor (Fig. 5h). However, compared to topotecan, irinotecan is a prodrug. It is hydrolyzed by uridine diphosphate glucuronosyltransferase 1A1 (UGT1A1) to its active metabolite SN-38 by carboxylesterases, which are abundant in plasma, liver, and cancer cells [273, 378, 379]. A study indicated that SN-38 is 1000 times more potent than irinotecan in inducing DNA SSBs [378]. Thus, SN-38 was adopted as a cytotoxic agent in a Trop-2-directed ADC. Oxaliplatin is a third-generation platinum-based drug that impairs normal DNA functions by generating mono-adducts and DNA crosslinks, similar to the first- (cisplatin) and second-generation (carboplatin) platinum drugs [380] (Fig. 5i). In contrast to cisplatin and carboplatin, oxaliplatin has a unique indication for colorectal cancer, as it facilitates organic cation transporter (OCT)-mediated uptake [381]. In addition, oxaliplatin shows different drug resistance mechanisms from cisplatin and carboplatin. Specifically, dMMR and replicative bypass increases that confer cisplatin resistance do not contribute to resistance to oxaliplatin [382]. On the other hand, multidrug resistance-associated protein 2 (MRP2)-mediated drug efflux limits both cisplatin and oxaliplatin accumulation [383, 384], rendering gastrointestinal cancer cells resistant to oxaliplatin [385] but not to cisplatin [381, 386]. Given the broad-spectrum antitumor activity of irinotecan and oxaliplatin, they have become essential ingredients in some classical regimens for colorectal cancer treatment, such as FOLFOXIRI (5-FU, leucovorin, oxaliplatin, and irinotecan), FOLFIRI (5-FU, leucovorin, and irinotecan) [387], FOLFOX (5-FU, leucovorin, oxaliplatin) [388], and CAPEOX (capecitabine and oxaliplatin) [389].

Similar to NSCLC, EGFR is overexpressed in approximately 50 ~ 80% of colorectal cancers [390, 391]. However, somatic mutations of EGFR occur at a very low frequency in colorectal cancer [392]. Thus, two EGFR-directed mAbs (cetuximab and panitumumab) were approved by the FDA for EGFR-positive metastatic colorectal cancer (Fig. 5j). Cetuximab interacts exclusively with the soluble extracellular region of EGFR and occludes the ligand-binding region on domain III of EGFR partially, which sterically prevents EGFR from adopting the extended conformation required for dimerization, thereby inhibiting the activation of EGFR [393]. Colorectal cancer harboring EGFRS492R, EGFRK467T, and EGFRR451C mutations confer cetuximab resistance but respond to panitumumab [394, 395]. These mutations may directly block cetuximab binding to domain III of EGFR but are permissive for panitumumab binding, which is attributed to a central cavity located between the heavy and light chain of panitumumab accommodating these mutations [396]. Given the low incidence of EGFR mutations, cetuximab and panitumumab are considered equivalent treatments in most clinical circumstances due to a shared epitope [397].

Compelling evidence indicates that VEGFR1 and VEGFR2 are the primary mediators of tumor angiogenesis and vascular permeability [398, 399]. Accordingly, VEGFR1/2-related ligands, vascular endothelial growth factors (VEGFs), have become promising targets in malignancies. The VEGF family consists of five glycoproteins, VEGF-A, -B, -C, -D, and placenta growth factor (PlGF). Each VEGF exerts its activity by binding to the corresponding receptors. Specifically, VEGF-A binds to VEGFR1 and VEGFR2, VEGF-B and PlGF bind exclusively to VEGFR1 [400], whereas VEGF-C and VEGF-D bind to VEGFR2 and VEGFR3 [401, 402]. Based on this principle, the VEGF-A-directed mAb bevacizumab was approved by the FDA as first-line therapy for metastatic colorectal cancer in combination with FOLFOXIRI (5-FU, leucovorin, oxaliplatin, and irinotecan) [403] (Fig. 5k). Bevacizumab binds to soluble VEGF-A and prevents VEGF-A from binding to its receptors (VEGFR1 and VEGFR2) by steric hindrance, thereby reducing blood vessel density, vascular permeability, and liver metastases of colorectal cancer mediated by VEGFR1 and VEGFR2 [404]. In contrast, ziv-aflibercept adopts a new strategy to antagonize VEGFs by utilizing the high binding affinity between VEGFRs and VEGFs (Fig. 5l). Specifically, ziv-aflibercept is constructed as a soluble receptor decoy that fuses the second immunoglobulin (Ig)-like domain of VEGFR1 and the third Ig-like domain of VEGFR2 to the Fc portion of human IgG1 [405]. Therefore, ziv-aflibercept acts as a VEGF trap that antagonizes multiple VEGFs, including VEGF-A, VEGF-B, and PlGF [406]. Similar to bevacizumab, ziv-aflibercept was also approved by the FDA as first-line therapy for metastatic colorectal cancer in combination with FOLFIRI (5-FU, leucovorin, and irinotecan) [407]. However, almost half of patients develop metastases, and most have unresectable tumors [408].

Increasing evidence indicates that the overactivation of RTKs and their downstream signaling cascades contribute to the development, progression, and acquired drug resistance of colorectal cancer [409, 410]. Regorafenib is a potent multitarget TKI that blocks angiogenic kinases (VEGFR1/2/3, PDGFRα/β, and FGFR1/2) and oncogenic kinases (KIT, RET, RAF1, BRAFWT, and BRAFV600E) [411] (Fig. 5m). CYP3A4 and UGT1A9 metabolize Regorafenib into two main circulating metabolites, M-2 (N-oxide) and M-5 (N-oxide/N-desmethyl) [412]. Both metabolites exhibit similar pharmacological activity to regorafenib. However, regorafenib primarily seems to induce stabilization of the disease rather than tumor regression in metastatic colorectal cancer because few patients achieve an objective tumor response [413]. Thus, regorafenib was approved by the FDA for patients with metastatic colorectal cancer who had received previous standard therapies [414].

Nevertheless, the OS benefit of regorafenib is 1.4 months, and over 50% of patients with colorectal cancer eventually develop resistance and progressive disease after a transient response to the standard therapy [413, 415]. Additional treatment options are needed for patients with metastatic colorectal cancer who have exhausted all standard therapies [416]. Trifluridine/tipiracil (known as TAS-102) is an antimetabolite agent that comprises a trifluridine (thymidine-based nucleoside analog) and a tipiracil (thymidine phosphorylase inhibitor) [417] (Fig. 5n). Like 5-FU, trifluridine inhibits thymidylate synthase (a central enzyme in DNA synthesis) and incorporates itself into DNA, leading to cell death [418]. Of note, trifluridine exhibits higher activity than 5-FU because it does not elicit an autophagic survival response as 5-FU [419]. Tipiracil attenuates thymidine phosphorylase-mediated catabolism of trifluridine, which increases the bioavailability and potentiates the in vivo efficacy of trifluridine [417]. Intriguingly, tipiracil/trifluridine exhibits pharmacological activity in both 5-FU-sensitive and 5-FU-resistant cancer cells [420, 421]. Thus, tipiracil/trifluridine was approved for the treatment of patients with metastatic colorectal cancer who are refractory to or are not considered candidates for current standard chemotherapy and biological therapy [422].

FDA-approved therapeutic drugs for prostate cancers

Prostate cancer accounted for 14.1% of cancer cases and 6.8% of cancer-related mortalities among males worldwide in 2020, second only to lung cancer [2]. Over the past 31 years, the FDA granted 12 new therapeutic drug approvals for prostate cancer (Fig. 6a and Table Table55).

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig6_HTML.jpg

FDA-approved therapeutic drugs for prostate cancers. a Distribution of therapeutic drugs for prostate cancers during the past 31 years (adapted from [423]). b AR antagonists. c GnRH agonist. d GnRH receptor antagonists. e Microtubule inhibitor. f CYP17A1 inhibitor. g α-particle-emitting radiopharmaceutical

Table 5

FDA-approved therapeutic drugs for prostate cancers

Drug (brand name)SponsorPropertiesIndicationApproval dateReview
Bicalutamide (Casodex)Ani pharmsAR antagonistAdvanced prostate cancer10/04/1995S
Nilutamide (Nilandron)ConcordiaAR antagonistMetastatic prostate cancer09/19/1996S
Enzalutamide (Xtandi)AstellasAR antagonistmCRPC08/31/2012P
Apalutamide (Erleada)JanssenAR antagonistProstate cancer02/14/2018P
Darolutamide (Nubeqa)BayerAR antagonistProstate cancer7/30/2019P
Triptorelin pamoate (Trelstar)VerityGnRH agonistAdvanced prostate cancer06/15/2000S
Abarelix (Plenaxis)SpecialityGnRH receptor antagonistAdvanced prostate cancer11/25/2003P
Degarelix acetate (Firmagon)FerringGnRH receptor antagonistAdvanced prostate cancer12/24/2008S
Relugolix (Orgovyx)MyovantGnRH receptor antagonistProstate cancer12/18/2020P
Cabazitaxel (Jevtana kit)SanofiMicrotubule-stabilizing agentProstate cancer06/17/2010P
Abiraterone acetate (Zytiga)JanssenCYP17A1 inhibitormCRPC04/28/2011P
Radium RA 223 dichloride (Xofigo)Bayerα-particle-emitting radiopharmaceuticalmCRPC05/15/2013P

AR Androgen receptor; CYP17A1 Cytochrome P450 17A1; GnRH Gonadotropin-releasing hormone; mCRPC Metastatic castration-resistant prostate cancer; O Orphan; P Priority; S Standard

The progression of prostate cancer is frequently accompanied by rising androgen receptor (AR) overexpression owing to the proliferation of luminal epithelial cells of the prostate caused by the accumulation of somatic mutations or AR amplification [423, 424]. Overexpression of AR enhances the binding activity to androgens [425], such as dihydrotestosterone (DHT), which initiates the translocation of AR from the cytoplasm to the nucleus [426], where AR binds to specific DNA sequences, namely, androgen response elements (AREs), thereby initiating the transcription of its target genes, including prostate-specific antigen (PSA) [427429]. As a result, PSA is frequently elevated in patients with prostate cancer and has become a classic biomarker for disease diagnosis [430], whereas an AR-mediated transcription program increases cell proliferation [431] and changes central metabolism and biosynthesis [432], leading to disease progression [433].

Based on this principle, five AR antagonists have been approved by the FDA for advanced or metastatic prostate cancer; these include two first-generation antiandrogens (nilutamide and bicalutamide) (another antiandrogen, flutamide, brand name: Eulexin, approved on January 27, 1989, by the FDA) and three second-generation antiandrogens (enzalutamide, apalutamide, and darolutamide) (Fig. 6b). Mechanistically, all these AR antagonists competitively bind to the ligand-binding domain (LBD) of AR and prevent androgens from binding to AR. Compared to first-generation antiandrogens, second-generation antiandrogens improve the pharmacologic properties by which AR translocation and AR-mediated transcription are blocked [434]. However, cross-resistance widely exists throughout antiandrogens due to mutations or deletions in the LBD of AR [435, 436]. Specifically, mutations of ARH874Y, ARV715M, and ART877A/S confer the conversion of flutamide and nilutamide from AR antagonists to agonists [437]. Fortunately, these AR mutants are sensitive to bicalutamide [437]. However, the ARW741C/L mutations convert bicalutamide from an AR antagonist to an agonist [438, 439] but are sensitive to nilutamide [440]. The ARF876L mutation also switches the second-generation antiandrogens enzalutamide and apalutamide from AR antagonists to agonists [441443] but is sensitive to the most novel antiandrogen darolutamide [444]. Moreover, darolutamide exhibits antagonistic effects on mutations of ARW741L and ART877A [444]. Nevertheless, it still cannot overcome the resistance of AR LBD-deletion variants [435]. Therefore, the N-terminus of AR should be considered the target domain for the next generation of antiandrogens.

Androgen deprivation therapy (ADT) is another strategy for the treatment of prostate cancer that suppresses serum testosterone to castration levels, thereby preventing AR activation and blocking AR-mediated transcription. Both gonadotropin-releasing hormone (GnRH, also known as luteinizing hormone-releasing hormone (LH-RH)) agonists and GnRH antagonists are used for ADT, although they have different pharmacological mechanisms [445].

GnRH agonists stimulate the pituitary gland, which causes a flare phenomenon by which testosterone levels are initially increased for 5 ~ 12 days [446]. However, sustained overstimulation leads to the downregulation and desensitization of GnRH receptors located in gonadotroph cells [447], thereby reducing the luteinizing hormone (LH) and follicle-stimulating hormone (FSH) levels, which eventually decreases serum testosterone and achieves castration levels [446, 447]. In contrast, GnRH antagonists induce a rapid decrease in LH and FSH by competitively binding to the GnRH receptors [448] and decrease serum testosterone to castration levels without causing a flare phenomenon [449]. Based on this principle, the GnRH agonist triptorelin (Fig. 6c) and three GnRH antagonists (abarelix, degarelix, and relugolix) (Fig. 6d) were approved by the FDA. Another GnRH agonist, histrelin, was approved by the FDA for prostate cancer under the brand name Vantas on Oct. 12, 2004 (Additional file 1: Table S1, page 1).

Similar to the previously approved GnRH agonists goserelin (brand name: Zoladex, approved by the FDA on Dec. 29, 1989) and leuprolide (brand name: Lupron Depot, approved by the FDA on January 26, 1989), both triptorelin (decapeptide) and histrelin (nonapeptide) are synthetic, polypeptide GnRH analogs. Compared to the endogenous GnRH (Glp-His-Trp-Ser-Tyr-Gly-Leu-Arg-Pro-Gly-NH2), both triptorelin and histrelin preserve the N-terminal five amino acid residues (Glp-His-Trp-Ser-Tyr) and C-terminal three amino acid residues (Leu-Arg-Pro) [450452]. Endogenous GnRH is rapidly degraded in blood by enzymatic cleavage at the Gly residue in position 6 [453]; Gly6 is replaced by D-Trp and D-His (Bzl) in triptorelin and histrelin, respectively, to increase resistance to degradation and thereby prolong the half-life time in vivo [452, 454]. On the other hand, the Gly10 of endogenous GnRH is replaced by AzaGly-NH2 in goserelin and Pro-NHEt in leuprolide and histrelin to increase the binding affinity between GnRH agonists and the GnRH receptor [452, 455]. Histrelin is a GnRH agonist administered once yearly that exhibits long-term efficacy and tolerability as a subcutaneous implant [456]. Although leuprolide, a GnRH agonist administered twice yearly [457], is comparable to histrelin in the drug administration schedule, 10% of patients treated with leuprolide failed to achieve medical castration [458].

Compared to triptorelin, histrelin may reduce the flare phenomenon and testosterone microsurges upon repeated administration to a certain extent. However, GnRH agonists cannot eliminate these adverse effects due to their natural pharmacological mechanism [452]. Therefore, GnRH antagonists have been developed for the treatment of advanced prostate cancer. However, first- and second-generation GnRH antagonists are unsuitable for clinical use due to solubility limitations and systemic allergic reactions caused by histamine release [452, 459, 460]. Abarelix and degarelix are third-generation GnRH antagonists derived from endogenous GnRH. Compared to endogenous GnRH, the N-terminal three amino acid residues (crucial for biological activity) Tyr5-Gly6, Arg8, and Gly10 are substituted in abarelix and degarelix to eliminate the biological activity of GnRH but increase the stability and binding affinity to the GnRH receptor [460].

As expected, abarelix induces a rapid suppression of serum testosterone and PSA levels and achieves medical castration without a testosterone surge [461463]. However, it also causes inevitable adverse effects, such as severe allergic reactions upon long-term administration [462, 464], and exhibits more frequent and shorter time intervals in escape from castration than complete ADT [446, 465]. Consequently, abarelix was withdrawn from the market in 2005 [466].

In contrast, degarelix is generally well tolerated, with most adverse events being mild to moderate in severity [467]. Additionally, the long-term clinical efficacy in suppressing testosterone and PSA levels are comparable to that of leuprolide over a one-year treatment period [468], and PSA-PFS is significantly improved upon degarelix treatment compared to leuprolide [469]. Thus, degarelix can be an alternative to GnRH agonists. Relugolix is a nonpeptidic drug and the first orally administered GnRH antagonist [470] that exhibits significantly superior clinical efficacy and a lower incidence of major adverse cardiovascular events than leuprolide [471]. Given the easier administration, relugolix is likely to become the new standard of care, although whether it is superior to surgical or established chemical castration treatments remains to be proven [472].

Cabazitaxel is a microtubule-stabilizing agent that binds to the N-terminus of the β-tubulin subunit, which promotes the assembly of tubulin into microtubules and stabilizes the mitotic spindle [473, 474]. It is synthesized from 10-deacetyl baccatin III, a compound extracted from the needles of yew trees (Taxus spp.) [475]. Compared to previous taxanes, such as paclitaxel and docetaxel, cabazitaxel exhibits favorable pharmacological efficacy, including increased cytotoxic activity in multidrug- and docetaxel-resistant cancer cells, probably due to a lower affinity for P-glycoprotein than docetaxel [476]. As expected, cabazitaxel exhibited an encouraging clinical advantage for the treatment of metastatic castration-resistant prostate cancer (mCRPC) compared to docetaxel [477, 478] and was approved as a second-line chemotherapy option for mCRPC [479] (Fig. 6e).

Cytochrome P450 17A1 (CYP17A1) is critical for producing androgenic and osteogenic sex steroids with its hydroxylase and 17, 20-lyase activities [480]. CYP17A1 is significantly elevated in mCRPC, making it an essential target for the treatment of mCRPC [481]. Abiraterone is a potent, selective, irreversible CYP17A1 inhibitor that binds to haem iron and occupies the majority of the enclosed active site of CYP17A1 [482], thereby attenuating the enzymatic activity of CYP17A1 and preventing androgen biosynthesis [483]. Abiraterone exhibits favorable clinical efficacy, making it an essential first-line option for the treatment of mCRPC [484487] (Fig. 6f).

Radium RA 223 dichloride (223RaCl2) is a radiopharmaceutical that has been approved for patients with prostate cancer-derived symptomatic bone metastases [488] (Fig. 6g). 223RaCl2 exerts its pharmacological effect through Radium-223 (223Ra), an α-particle-emitting radioisotope. 223Ra is also a calcium mimetic that binds preferentially to the newly formed bone in areas of bone metastases, with a half-life of 11.4 days and maximal tissue penetration of fewer than 100 μm [489, 490]. Each atom of 223Ra emits four high linear energy α-particles (composed of two protons and two neutrons), which exert pharmacological actions by inducing DNA DSBs in directly irradiated cells and adjacent cells [490] or by producing extracellular reactive oxygen species (ROS) in directly irradiated cells and then inducing DNA DSBs in adjacent cells with a bystander effect [491].

FDA-approved therapeutic drugs for urologic cancers

Aside from prostate cancer, kidney and bladder cancers are the most common urologic cancers [2]. Over the past 31 years, the FDA granted 12 new therapeutic drug approvals for urologic cancers (Fig. 7a and Table Table66).

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig7_HTML.jpg

FDA-approved therapeutic drugs for urologic cancers. a Distribution of therapeutic drugs for urologic cancers during the past 31 years (adapted from [867]). b Recombinant human IL-2 (obtained from www.rcsb.org and go.drugbank.com). c Multitarget TKI and VEGFR inhibitors. d mTOR inhibitors. e PDL1-directed mAbs. f Nectin-4-directed ADC. g DNA topoisomerase inhibitor. h FGFR inhibitor

Table 6

FDA-approved therapeutic drugs for urologic cancers

Drug (brand name)SponsorPropertiesIndicationApproval dateReview
Aldesleukin (Proleukin)ChironRecombinant human interleukin-2Metastatic RCC05/05/1992O
Sorafenib toylate (Nexavar)BayerMultitarget TKI (VEGFR2/3, PDGFRβ, FLT3, KIT, RAF1, and BRAF)Advanced RCC12/01/2005P, O
Pazopanib HCl (Votrient)NovartisMultitarget TKI (VEGFRs, PDGFRα/β, FGFR1/2, KIT)Metastatic RCC10/19/2009S
Axitinib (Inlyta)PfizerVEGFRs inhibitorAdvanced RCC01/27/2012S
Tivozanib HCl (Fotivda)Aveo PharmsVEGFRs inhibitorAdvanced RCC03/10/2021S
Temsirolimus (Torisel)Merck & ComTOR inhibitorAdvanced RCC05/30/2007P
Everolimus (Afinitor)NovartismTOR inhibitorAdvanced RCC03/30/2009P
Atezolizumab (Tecentriq)GenentechPDL1-directed mAbUrothelial carcinoma05/18/2016N/A
Durvalumab (Imfinzi)AstraZenecaPDL1-directed mAbUrothelial carcinoma05/01/2017N/A
Enfortumab vedotin (Padcev)AstellasNectin-4-directed ADCUrothelial carcinoma12/18/2019P
Valrubicin (Valstar)Endo PharmA semisynthetic analog of anthracycline doxorubicinBCG-refractory CIS of urinary bladder cancer09/25/1998P, O
Erdafitinib (Balversa)JanssenFGFR inhibitorBladder cancer04/12/2019P

BCG Bacillus Calmette-Guérin; CIS Carcinoma in situ; FGFR Fibroblast growth factor receptor; mTOR Mammalian target of rapamycin; O Orphan; P Priority; PDGFR Platelet-derived growth factor receptor; PDL1 Programmed death-ligand 1; RCC Renal cell carcinoma; S Standard; VEGFR Vascular endothelial growth factor receptor

Renal cell carcinoma

Kidney cancer accounted for 2.2% of cancer cases and 1.8% of cancer-related mortalities worldwide in 2020 [2]. Renal cell carcinoma (RCC) is the most common subtype (~ 70%) of kidney cancer [492]. One recombinant human interleukin-2 (IL-2, aldesleukin), four TKIs (sorafenib, pazopanib, axitinib, and tivozanib), and two mammalian targets of rapamycin (mTOR) inhibitors (temsirolimus and everolimus) have been approved for the treatment of RCC in the past three decades (Fig. 7b-d).

IL-2 was first discovered as a T cell growth factor in 1976 [493] and cloned in 1983 [494]. Over the ensuing years, IL-2 was proven to be a pivotal cytokine produced primarily by CD4+ T cells [495, 496]. As a pleiotropic mediator within the immune system, it interacts with IL-2 receptors (IL-2Rα, IL-2Rβ, and IL-2Rγ) and induces the proliferation and differentiation of immune cells, thereby regulating a range of diseases involving infection, autoimmune disease, and cancer [496]. Aldesleukin is a nonglycosylated, modified form of human endogenous IL-2 that exerts its antitumor activity by enhancing the cytotoxicity of T lymphocytes and the activity of natural killer and lymphokine-activated killer (LAK) cells [497] (Fig. 7b). Aldesleukin monotherapy achieves an ORR of 14 ~ 25% and exhibits durable antitumor activity in patients with metastatic RCC [498, 499].

Sorafenib is an oral first-generation multitarget TKI that targets several RTKs, including RAF1, BRAFWT, BRAFV600E, VEGFRs, PDGFR-β, FGFR1, FMS-like tyrosine kinase 3 (FLT3), KIT, and RET [500, 501]. It occupies the ATP adenine binding pocket of these RTKs with its distal 4-pyridyl ring, which blocks the autophosphorylation of these RTKs, thereby attenuating the mitogen-activated protein kinase (MAPK)/extracellular signal-regulated kinase (ERK) signaling pathway [500, 502]. Given its potent antitumor effects, sorafenib is applied to various malignancies in addition to RCC [501]. Sorafenib resistance inevitably occurs and mainly involves mutations in RTKs and activation of the bypass pathway [503]. Regorafenib, as a fluoro‐sorafenib, provides a nearly 3-month improvement in OS in HCC patients progressing on sorafenib treatment [504]. However, regorafenib cannot overcome sorafenib resistance because they share a similar structure [505]. Sunitinib is another first-generation multitarget TKI with a similar target profile to sorafenib. Intriguingly, sequential sorafenib-sunitinib and vice versa provide similar clinical benefits in metastatic RCC [506]. In contrast, pazopanib is an oral second-generation multitarget TKI that preferentially targets VEGFRs, PDGFRα/β, and KIT [507509]. It competes with ATP for binding to the cytoplasmic domain of these RTKs and prevents ATP-induced activation [510]. Pazopanib retains clinical activity in patients with advanced clear-cell RCC after failure of sunitinib or bevacizumab [511]. Thus, pazopanib is non-inferior to sunitinib as first-line therapy in clinical efficacy and exhibits advantages in the safety profile [512]. Axitinib and tivozanib are selective second-generation VEGFR inhibitors that exhibit greater selectivity for VEGFRs than other TKIs (e.g., sorafenib, sunitinib, pazopanib) [513, 514]. Axitinib is a substituted indazole derivative produced from a structure-based drug design [515]. It exhibits antitumor activity and a manageable safety profile in sorafenib-refractory metastatic RCC [516] but has no significant superiority over sorafenib as first-line therapy [517]. Compared with sorafenib, axitinib significantly prolongs the median PFS by two months and can be an option for second-line therapy [518, 519]. Nevertheless, axitinib plus avelumab or pembrolizumab therapies exhibit significant clinical benefits in advanced RCC as first-line treatment compared to the standard of care of sunitinib [520, 521]. Tivozanib is a quinoline urea derivative that interacts with the ATP-binding site and the allosteric-binding site consisting of the DFG motif within the activation loop of VEGFR, similar to sorafenib [522]. It inhibits VEGF-induced VEGFR phosphorylation and blocks VEGF-dependent but not VEGF-independent MAPK activation [523]. Tivozanib improves PFS and is better tolerated as third- or fourth-line therapy than sorafenib [524] (Fig. 7c).

mTOR is a serine/threonine-protein kinase that governs a diverse set of biological events by joining with other components to form two distinct complexes known as mTOR complex 1 (mTORC1) and mTOR complex 2 (mTORC2) [525]. mTORC1 is composed of three core components: mTOR, mammalian lethal with SEC13 protein 8 (mLST8, also known as GβL) [526], and its unique defining subunit, regulatory-associated protein of mTOR (RAPTOR) [527]. In lieu of RAPTOR, mTORC2 contains rapamycin-insensitive companion of mTOR (RICTOR) [528, 529]. mTORC1 governs glucose metabolism, cell cycle progression, cell survival, and the biosynthesis of proteins, lipids, and nucleotides, while mTORC2 governs cytoskeletal rearrangement and prosurvival pathways [525]. Temsirolimus and everolimus are derivatives of sirolimus (also known as rapamycin) (Additional file 1: Table S1, page 11), a compound extracted from a Streptomyces hygroscopicus soil bacterium [530]. Mechanistically, rapamycin, temsirolimus, and everolimus are inhibitors of mTORC1. Similar to rapamycin, both temsirolimus and everolimus bind to FK506-binding protein 12 (FKBP12) and form a gain-of-function complex, which subsequently prohibits the activation of mTOR, resulting in cell cycle arrest and suppression of hypoxia-inducible factor-1α (HIF1α) and VEGF expression [531, 532].

Temsirolimus is a water-soluble ester of rapamycin with improved pharmaceutical properties, including stability and solubility, making it suitable for intravenous administration [533]. Intriguingly, temsirolimus is hydrolyzed by CYP3A4 to its major metabolite rapamycin in vivo [534]. Compared with interferon-α (IFNα) monotherapy, temsirolimus improves OS among patients with metastatic RCC [535]. Everolimus is a hydroxyethyl ether derivative of rapamycin with superior pharmaceutical characteristics, making it suitable for oral administration. Unlike temsirolimus, everolimus is not converted to rapamycin in vivo [532]. Compared with placebo, everolimus prolongs PFS in patients with metastatic RCC previously treated with sunitinib and/or sorafenib [536]. Compared to temsirolimus, everolimus exhibits superior clinical efficacy in metastatic RCC in terms of prolonging OS and PFS and decreasing the risk of death [537, 538] (Fig. 7d).

RCC is a highly vascularized tumor prone to distant metastasis [503]. Mechanistically, clear-cell RCC accounts for approximately 80 ~ 85% of metastatic RCC cases [539], whereas 60% of clear-cell RCC harbors loss-of-function of the von Hippel–Lindau (VHL) tumor suppressor gene, which leads to the accumulation of HIF1α and activation of its target genes, including VEGF and PDGF [540]. It explains why all these TKIs target VEGFRs and/or PDGFRs, although they have different target profiles. In addition, two mTOR inhibitors can reduce the expression of HIF1α and VEGF. Before 2005, nonspecific immune cytokines, such as IL-2 and IFNα, were previously the mainstays of therapy for advanced RCC [540]. Currently, PD1-directed mAbs, such as pembrolizumab and nivolumab, plus TKIs, such as axitinib [521], lenvatinib [541], and cabozantinib [542], have become the first-line regimens for patients with advanced or metastatic RCC.

Bladder cancer

Bladder cancer accounted for 3.0% of cancer cases and 2.1% of cancer-related mortalities worldwide in 2020 [2]. Urothelial carcinoma accounts for approximately 90% of bladder cancers [543]. Over the past 31 years, the FDA has approved five therapeutic drugs for bladder cancer treatment (Fig. 7e–h).

PDL1 is an immune checkpoint expressed in 20% of tumor cells and 40% of tumor-infiltrating mononuclear cells (TIMCs) in urothelial carcinoma [544]. It binds to its receptor PD1 on the surface of T lymphocytes and negatively regulates the antitumor function of T lymphocytes [545]. PDL1-directed mAbs bind exclusively to PDL1 and block the interaction between PDL1 and PD1, which reactivates the antitumor immunity of T lymphocytes [546]. Mechanistically, atezolizumab binds to the CC′, C′C″, and FG loops of PDL1, while durvalumab binds to the CC′ loop and N-terminal region of PDL1 [547]. The Fc fragments of both mAbs are engineered to eliminate the ADCC effect and complement-dependent cytotoxicity (CDC) and prevent the depletion of activated T lymphocytes [548, 549]. Atezolizumab is the first PDL1-directed mAb [550] that exhibits durable activity, good tolerability, and superior clinical efficacy compared with chemotherapy in patients with platinum-treated locally advanced or metastatic urothelial carcinoma [551553]. Notably, an increase in the mutation load increases the response to atezolizumab [551]. Likewise, durvalumab exhibits at least equivalent clinical efficacy to atezolizumab [554556]. Currently, other PDL1 (avelumab) [557, 558] and PD1 (pembrolizumab [559, 560] and [561, 562])-directed mAbs are also used as first- or second-line treatments for urothelial carcinoma (Fig. 7e).

Nectin-4 is a type I transmembrane protein expressed at low levels in normal human tissues and is also known as poliovirus receptor-like 4 (PVRL4) [563, 564]. It acts as an oncoprotein [565] that promotes cancer cell proliferation and metastasis by activating Wnt/β-catenin [566] and the HER2-mediated PI3K/AKT signaling pathway [567569]. Nectin-4 is expressed in 69% of solid tumors [563] and overexpressed in more than 60% of bladder cancers (or urothelial carcinomas) [563, 570], making it an attractive target for urothelial carcinoma treatment. Enfortumab vedotin is a nectin-4-directed ADC composed of enfortumab and MMAE with a protease-cleavable MC-vc-PAB linker [563], similar to tisotumab vedotin. Enfortumab is a nectin-4-directed mAb that binds to the extracellular domain of human nectin-4 [563] (Fig. 7f). Enfortumab vedotin significantly improves the median OS and PFS compared with that with chemotherapy (docetaxel, paclitaxel, or vinflunine) in patients with locally advanced or metastatic urothelial carcinoma who were previously treated with platinum-based treatment and PD1/PDL1 blockade, providing a new option for this patient population [571, 572].

Valrubicin is a semisynthetic analog of the anthracycline doxorubicin that binds weakly to DNA [573] (Fig. 7g). Compared to doxorubicin and epirubicin, valrubicin and its metabolites exhibit lower potency and less toxicity and exert antitumor effects by interfering with TOP2A-mediated cleavage and resealing of DNA, leading to the inhibition of DNA elongation and RNA biosynthesis [573575]. Therefore, the S-G2 transition of the cell cycle is blocked, and chromosome stability is disrupted. Valrubicin is effective and well tolerated in patients with bacillus Calmette-Guérin (BCG)-refractory carcinoma in situ (CIS) of the bladder [576].

In addition, genetic alterations of FGFRs, including amplification, mutation, and rearrangement, occur in approximately one-third of urothelial carcinomas, making FGFRs promising therapeutic targets [577]. Erdafitinib is an oral pan-FGFR inhibitor that binds to the inactive DGF-Din conformation, which prevents the FGF ligand-induced dimerization, phosphorylation, and activation of FGFRs [578580] (Fig. 7h). Compared to other inhibitors (rogaratinib [581], pemigatinib [582], and infigratinib [583]) with an ORR of approximately 25%, erdafitinib exhibits superior clinical efficacy with an ORR of 40% [584]. Thus, erdafitinib was approved for patients with locally advanced or metastatic urothelial carcinoma harboring FGFR2 or FGFR3 genetic alterations [579].

FDA-approved therapeutic drugs for melanoma and other skin cancers

Melanoma

Melanoma accounted for 1.7% of cancer cases and 0.6% of cancer-related mortalities worldwide in 2020 [2]. Nine therapeutic drugs (including BRAF and MAPK/ERK kinase (MEK) inhibitors and cytotoxic T lymphocyte antigen 4 (CTLA4)- and PD1-directed mAbs) have been approved by the FDA for melanoma in the past three decades (Fig. 8a and Table Table77).

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig8_HTML.jpg

FDA-approved therapeutic drugs for melanoma and other skin cancers. a Distribution of therapeutic drugs for melanoma and other skin cancers during the past 31 years (adapted from [868]). b BRAF inhibitors. c MEK inhibitors. d CTLA4-directed mAb. e PD1-directed mAbs. f Smoothened inhibitors. g PDL1-directed mAb. h PD1-directed mAb

Table 7

FDA-approved therapeutic drugs for melanoma and other skin cancers

Drug (brand name)SponsorPropertiesIndicationApproval dateReview
Vemurafenib (Zelboraf)RocheBRAF inhibitorBRAF-positive unresectable or metastatic melanoma08/17/2011P, O
Dabrafenib mesylate (Tafinlar)NovartisKinase inhibitor with activity against BRAFV600E/D/K, wild-type BRAF and other kinasesUnresectable or metastatic melanoma with BRAFV600E mutation05/29/2013S, O
Encorafenib (Braftovi)Array BioPharmaBRAF inhibitorBRAF-mutated melanoma06/27/2018S, O
Trametinib DMSO (Mekinist)NovartisMEK1/2 inhibitorUnresectable or metastatic melanoma with BRAFV600E mutation05/29/2013S, O
Cobimetinib fumarate (Cotellic)GenentechMEK inhibitorMelanoma with BRAFV600E/K mutations11/10/2015P, O
Binimetinib (Mektovi)Array BioPharmaMEK inhibitorBRAF-mutated melanoma06/27/2018S, O
Ipilimumab (Yervoy)Bristol-Myers SquibbCTLA4-directed mAbUnresectable or metastatic melanoma03/25/2011P, O
Pembrolizumab (Keytruda)MerckPD1-directed mAbMetastatic melanoma09/04/2014P, O
Nivolumab (Opdivo)Bristol-Myers SquibbPD1-directed mAbUnresectable or metastatic melanoma12/22/2014P, O
Vismodegib (Erivedge)GenentechSmoothened inhibitorAdvanced basal cell carcinoma01/30/2012P
Sonidegib phosphate (Odomzo)Sun PharmaSmoothened inhibitorBasal cell carcinoma07/24/2015S
Avelumab (Bavencio)Emd SeronoPDL1-directed mAbMerkel cell carcinoma03/23/2017P, O
Cemiplimab (Libtayo)RegeneronPD1-directed mAbCutaneous squamous cell carcinoma09/28/2018P

CTLA4 Cytotoxic T lymphocyte antigen 4; MEK MAPK/ERK kinase; O Orphan; P Priority; PD1 Programmed death receptor-1; PDL1 Programmed death-ligand 1; S Standard

BRAF forms a tight heterodimer with CRAF under the induction of active RAS and acts as a critical effector in the RAS–RAF–MEK–MAPK/ERK pathway [585]. However, approximately 70% of melanomas harbor BRAF mutations, whereas BRAFV600E and BRAFV600K account for 80 ~ 90% and 10 ~ 20% of all BRAF mutations, respectively [586, 587], making it a therapeutic target in melanoma [588]. Oncogenic BRAF constitutively activates the MAPK/ERK pathway, resulting in uncontrolled cell proliferation [589]. Therefore, three BRAF inhibitors (vemurafenib, dabrafenib, and encorafenib) (Fig. 8b) were approved for melanoma treatment.

Vemurafenib is a second-generation inhibitor with a mild selectivity for BRAFV600E over BRAFWT. It occupies the ATP-binding pocket in the ‘αC-OUT/DFG-in’ (active) conformation of BRAF and inhibits BRAF phosphorylation and activation, thereby attenuating downstream MEK–MAPK/ERK signaling transduction [590]. As expected, vemurafenib significantly improved OS and PFS in patients with previously untreated melanoma harboring the BRAFV600E mutation [591]. In parallel, dabrafenib, another highly potent and specific inhibitor, exhibits a virtually identical clinical outcome to vemurafenib [592] and a BRAF binding mechanism similar to that of vemurafenib [593]. Dabrafenib plus trametinib (a MEK inhibitor) therapy adds a clear benefit over vemurafenib monotherapy in patients with unresectable or metastatic melanoma harboring BRAFV600E/K mutations [594]. Unfortunately, most patients treated with vemurafenib or dabrafenib will develop disease progression following tumor regression within 6 ~ 8 months [595]. Encorafenib is still an αC-OUT inhibitor of BRAF and is used in combination with binimetinib (another MEK inhibitor) in clinical practice [596]. It showed a longer residence time and lower off-rate than vemurafenib and dabrafenib in a preclinical study [597, 598]. Compared to vemurafenib or dabrafenib monotherapy, encorafenib plus binimetinib combination therapy significantly improves clinical efficacy and tolerability [599, 600].

MEK is a direct downstream target of BRAF, and BRAF mutations that cause overactivation of the RAS–RAF–MEK–MAPK/ERK pathway highly depend on MEK activity [601]. Thus, three MEK inhibitors (trametinib, cobimetinib, and binimetinib) have been approved and are frequently used in combination with BRAF inhibitors [602] (Fig. 8c). Trametinib stably binds to unphosphorylated MEK1/2 with high affinity and maintains MEK in an unphosphorylated state [603]. However, trametinib shows a low affinity for phosphorylated MEK1/2 [604]. In contrast, cobimetinib not only inhibits ERK1/2 phosphorylation but also retains the inhibitory effect of phosphorylated MEK1/2 [605], whereas binimetinib exhibits clinical activity in both BRAF-mutated and NRAS-mutated melanoma [606]. In clinical practice, trametinib is used in combination with dabrafenib [607, 608], cobimetinib is used with vemurafenib [609, 610], and binimetinib is used with encorafenib [599, 600]. All three combined therapies exhibit equivalent clinical outcomes, such as objective/complete response rate, median PFS, toxic effects, and two-year survival rate. Of note, the mOS is 33.6 months with encorafenib plus binimetinib, 22.3 months with vemurafenib plus cobimetinib, and 25.1 months with dabrafenib plus trametinib [602].

CTLA4 is a second counterreceptor for the B7 family of costimulatory molecules that functions as a negative regulator of T lymphocyte activation [611]. Blocking CTLA4 significantly enhances antitumor immunity [612]. Ipilimumab is the first CTLA4-directed mAb that binds to CTLA4 on the cell surface, thereby blocking the interaction between CTLA4 and B7.1/B7.2 and restoring the activation of T lymphocytes [613] (Fig. 8d). Compared with gp100 monotherapy, ipilimumab monotherapy or plus glycoprotein 100 (gp100) significantly improved the median OS of patients with advanced or metastatic melanoma [614].

PD1 is another negative regulator of T lymphocytes that confers tumor immune evasion by interacting with its ligands PDL1 and PDL2 [615, 616]. PD1 is also expressed in melanoma cells and contributes to tumor growth [617]. Pembrolizumab is a PD1-directed mAb that binds to the C’D loop of PD1 [618620]. Pembrolizumab monotherapy is significantly superior to ipilimumab monotherapy in clinical trials [621, 622] and can be an effective treatment option for patients with ipilimumab-refractory advanced melanoma [623]. Nivolumab is another PD1-directed mAb that binds to the N-loop of PD1 [618, 620, 624] (Fig. 6e). It is frequently used in combination with ipilimumab for patients with advanced or metastatic melanoma [625, 626]. Compared with ipilimumab monotherapy, nivolumab monotherapy or ipilimumab plus ipilimumab significantly extends OS and five-year survival [627, 628] (Fig. 8e).

Nonmelanoma skin cancers (NMSCs) mainly encompass basal cell carcinoma, squamous cell carcinoma, and neuroendocrine skin carcinoma (also known as Merkel cell carcinoma) [629]. NMSCs are the most commonly diagnosed cancers, accounting for up to 30% of all human tumors [629632] and 0.6% of cancer-related mortalities [2].

Basal cell carcinoma

Basal cell carcinoma constitutes approximately 80% of all NMSCs, and more than five million new cases are diagnosed each year worldwide [633, 634]. However, the absolute incidence and mortality are difficult to determine since basal cell carcinoma is usually excluded from cancer registry statistics [633]. In part, basal cell carcinoma is the most frequent human cancer subtype [632]. Loss-of-function mutations of tumor suppressor gene patched homolog 1 (PTCH1) occur in 30 ~ 40% of basal cell carcinomas [635]. Dysfunctional PTCH1 causes constitutive activation of smoothened (SMO), resulting in continuous activation of hedgehog signaling and its target genes in basal cell carcinoma [174, 634], making SMO a promising target. Therefore, two SMO inhibitors (vismodegib and sonidegib) were approved for the treatment of basal cell carcinoma [636, 637] (Fig. 8f).

Vismodegib is the first SMO inhibitor that occupies the transmembrane domain core and forms hydrophobic interactions with SMO by a network of hydrogen bonds [638, 639], thereby inhibiting SMO activity and downstream signaling, regardless of PTCH1 [640]. However, approximately 21% of patients develop resistance within a year while undergoing continuous vismodegib treatment [641]. Various mutations in SMO are located in the drug-binding pocket of SMO and confer resistance to vismodegib by abrogating or impairing vismodegib binding to SMO, such as D473H, D473G, W281C, V321M, I408V, C469Y, and Q477E [642646]. Sonidegib is another SMO inhibitor that binds to a drug-binding pocket of SMO similar to vismodegib. It exerts antitumor effects by inhibiting the transcriptional activity of glioma-associated oncogene (GLI) and inducing the expression of caspase-3 and the cleavage of PARP, resulting in cell cycle arrest and apoptosis [647]. However, sonidegib has an SMO binding pattern similar to that of vismodegib. As expected, sonidegib cannot overcome the vismodegib resistance induced by SMO mutations [648].

Merkel cell carcinoma

Merkel cell carcinoma is a rare but highly aggressive NMSC with neuroendocrine features [649] frequently associated with Merkel cell polyomavirus infection and accumulation of ultraviolet-induced mutations [650]. Approximately 50% of tumor cells and 55% of tumor-infiltrating immune cells (TIICs) express PDL1 in Merkel cell carcinoma [651]. Avelumab is a PDL1-directed mAb that binds to the CC’ loop of PDL1 [547, 652], thereby blocking the interaction between PDL1 and PD1, which reactivates the antitumor immunity of T lymphocytes, similar to atezolizumab and durvalumab (Fig. 8g). Avelumab is well tolerated with durable responses [653] and has become the standard-of-care treatment for metastatic and advanced Merkel cell carcinoma [654], similar to other anti-PD1 immunotherapies, including pembrolizumab [655, 656] and nivolumab [657].

Cutaneous squamous cell carcinoma

Cutaneous squamous cell carcinoma (CSCC) accounts for approximately 20% of NMSCs, second only to basal cell carcinoma in NMSCs [658]. In contrast to most basal cell carcinomas, CSCC is highly aggressive, prone to metastasis, and correlated with ultraviolet radiation [658, 659]. More than half of CSCC TIICs express PD1, especially CD4+ and CD8+ TILs, which show PD1 positivity rates of 73% and 80%, respectively [660]. Cemiplimab is a PD1-directed mAb that binds to the BC loop, C’D loop, and FG loop of PD1 with its heavy chain variable domain (VH). In contrast, the light chain variable domain (VL) of cemiplimab sterically inhibits the interaction between PDL1 and PD1 [661] (Fig. 8h). Given the considerable antitumor activity and acceptable safety, cemiplimab was approved for patients with metastatic or locally advanced CSCC [662664].

FDA-approved therapeutic drugs for thyroid cancer and other solid tumors

Thyroid cancer

Thyroid cancer accounted for 3.0% of cancer cases and 0.6% of cancer-related mortalities worldwide in 2020 [2]. Contrary to pancreatic cancer, thyroid cancer has the lowest mortality-to-incidence ratio (0.133) among all malignancies [2].

Differentiated thyroid cancer (DTC) is derived from the follicular epithelial cells of the thyroid, accounting for approximately 95% of all cases, whereas surgery followed by either radioiodine therapy or observation is the standard treatment for most patients [665]. It is crucial to stimulate iodine uptake by elevating thyroid-stimulating hormone (TSH) or depleting thyroid hormone prior to radioiodine (iodine-131) administration [666]. Thyrotropin alfa is a recombinant human TSH (rhTSH) synthesized in a genetically modified Chinese hamster ovary cell line (Fig. 9a). It stimulates the thyroid gland to produce thyroxine (T4), and its more bioactive form triiodothyronine (T3), which increases iodine uptake. Clinically, thyrotropin alfa is used for radioiodine ablation of DTC [667] and radioiodine scanning of poorly differentiated thyroid cancer [668]. However, 9% of DTCs recur after thyroid hormone plus radioiodine therapy [669], and approximately 30% of patients with advanced, metastatic DTCs have the radioiodine-refractory disease [670]. Loss or low expression of sodium–iodide symporter (NIS) is associated with radioiodine refractoriness [669, 670]. Genetic and epigenetic alterations induced activation of RTKs and their downstream RAS–RAF–MEK–MAPK/ERK and PI3K–AKT–mTOR pathways contribute to the dysfunction of NIS [671673]. Lenvatinib is a TKI that targets VEGFRs, FGFRs, PDGFRα, RET, and KIT [674, 675] (Fig. 9b). It binds to the ATP-binding site and the neighboring region of RTKs, adopting a DFG-in conformation, compared to the DFG-out conformation of sorafenib [676]. As expected, lenvatinib significantly improves PFS with a high response rate in patients with radioiodine-refractory thyroid cancer [677].

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig9_HTML.jpg

FDA-approved therapeutic drugs for thyroid cancer and other solid tumors. a Recombinant human thyroid-stimulating hormone (obtained from www.rcsb.org and go.drugbank.com). b & c Multitargeted TKIs. d 9-cis-retinoic acid. e DNA alkylating agent. f PDGFRα-directed mAb. g CSF1R, KIT, and FLT3 inhibitor. h EZH2 inhibitor. i DNA alkylating agent. j GD2-directed mAbs. k Folate analog. l TRKs inhibitor

Medullary thyroid cancer (MTC) originates in the parafollicular neuroendocrine cells of the thyroid and accounts for 1 ~ 2% of all cases [665]. Acquired somatic RET mutations and germline RET mutations are observed in 35 ~ 50% and 6.5% of sporadic MTCs, respectively [678, 679], and are considered secondary events rather than initiators that drive the tumorigenesis of MTC [680]. Of note, RETM918T mutation is associated with a more aggressive disease and a poorer prognosis [681]. EGFR is overexpressed in 13% of MTCs, VEGFR2 expression is significantly higher in metastases than in the primary tumors of MTCs [682], and MET is overexpressed in thyroid epithelial cells [683]. These RTKs contribute to the tumorigenesis and angiogenesis of MTC [684]. Vandetanib is an inhibitor of VEGFR2, EGFR, and RET [678] that binds to the ATP-binding site of RTKs [685], leading to cell apoptosis rather than cell cycle arrest [686, 687]. However, RETV804M/L gatekeeper mutations and RET-S904F mutation confer resistance to vandetanib, mainly by increasing the ATP affinity and autophosphorylation activity of RET kinase [688, 689]. Cabozantinib is an inhibitor of VEGFR2, MET, and RET [684] that also binds to the ATP-binding site of RTKs [690], thereby inhibiting autophosphorylation of RTKs, which leads to tumor hypoxia and apoptosis and suppresses metastasis, angiogenesis, and tumor growth [684, 691]. Cabozantinib significantly prolongs PFS in patients with unresectable, locally advanced, or metastatic MTC [684, 692, 693]. Intriguingly, cabozantinib potently inhibits native ROS1 and the crizotinib-resistant ROS1G2032R mutation [694] and overcomes crizotinib resistance in CD74-ROS1D2033N-rearranged lung cancer [695] (Fig. 9c).

Soft tissue sarcomas (STSs) are rare tumors that account for 1% of all adult malignancies; these include Kaposi’s sarcoma, adipocytic tumors (e.g., liposarcoma), smooth muscle tumors (e.g., leiomyosarcoma), fibrohistiocytic tumors (e.g., tenosynovial giant cell tumor), tumors of uncertain differentiation (e.g., epithelioid sarcoma) [696], etc.

Kaposi’s sarcoma

Kaposi’s sarcoma accounted for 0.2% of cancer cases and 0.2% of cancer-related mortalities worldwide in 2020 [2]. It is a relatively rare cancer caused by Kaposi’s sarcoma-associated herpesvirus (KSHV, also known as human herpesvirus 8 (HHV8)) infection [697]. The skin and superficial mucosae are the most common sites of Kaposi’s sarcoma lesions [698]. Retinoid X receptor α (RXRα) and retinoic acid receptor γ (RARγ) control cell differentiation, proliferation, and apoptosis and are predominantly expressed in the skin [699, 700]. Alitretinoin is a 9-cis-retinoic acid that acts as a pan-agonist of RARs and RXRs [700]. It modulates cell differentiation and apoptosis in a variety of sarcomas by potentially inducing the formation of a homodimer of RXRs [701703] (Fig. 9d). Alitretinoin gel demonstrates durable responses with tolerable safety in patients with acquired immunodeficiency syndrome (AIDS)-related Kaposi’s sarcoma [704, 705].

Liposarcoma

Liposarcoma is a rare malignant tumor of adipocytic differentiation that accounts for 15 ~ 20% of STS cases [706]. It is characterized by recurrent amplifications within chromosome 12, which leads to the overexpression of disease-driving genes [706]. Leiomyosarcoma is a malignant mesenchymal tumor that accounts for 10 ~ 20% of STS cases [707]. Leiomyosarcoma also exhibits complex genomic alterations involving DNA copy number changes and gene mutations [708]. Trabectedin is a tetrahydroisoquinoline alkaloid derived from the Caribbean marine tunicate Ecteinascidia turbinata [709] (Fig. 9e). It binds to the minor groove of DNA that bends DNA toward the major groove by forming trabectedin-DNA adducts, which block the G2/M phase transition of the cell cycle and inhibit cell proliferation [709711]. The FDA-approved Trabectedin for patients with unresectable or metastatic liposarcoma or leiomyosarcoma who received a prior anthracycline-containing regimen [712].

Soft tissue sarcoma

PDGFRα expression in STSs is sevenfold higher than in normal tissues [713]. Olaratumab is a PDGFRα-directed mAb that selectively binds to the extracellular domain of PDGFRα, which prevents PDGF-AA, PDGF-BB, and PDGF-CC ligands from binding to PDGFRα [714, 715] (Fig. 9f). It inhibits the ligand-induced autophosphorylation of PDGFRα and downstream signaling, thereby blocking PDGFRα-mediated cell mitogenesis [716]. Olaratumab plus doxorubicin combination therapy significantly improves the PFS and OS compared to doxorubicin alone in patients with advanced STSs [717].

Tenosynovial giant cell tumor

Tenosynovial giant cell tumor (TGCT) is a rare, locally aggressive neoplasm mainly characterized by colony-stimulating factor-1 (CSF1) translocations and CSF1 receptor (CSF1R) overexpression [718, 719]. CSF1 translocations result in local overexpression of CSF1, which attracts histiocytoid and CSF1R-expressing inflammatory cells [718720]. Moreover, CSF1 promotes the differentiation of monocytes into tumor-associated macrophages (TAMs), which in turn facilitates tumor survival, growth, and metastases with their immunosuppressive effects [721723]. Thus, the CSF1/CSF1R axis is critical for the tumorigenesis and progression of TGCTs. Pexidartinib is an inhibitor of CSF1R, KIT, and FLT3 that accesses the autoinhibited state of CSF1R through direct interactions with juxtamembrane residues embedded in the ATP-binding pocket, thereby blocking the CSF1/CSF1R axis [724] (Fig. 9g). Notably, pexidartinib retains activity against the quizartinib-resistant FLT3 gatekeeper F691L mutation [725]. As the first systemic therapy of TGCT, pexidartinib exhibits a robust tumor response with improved clinical outcomes [718, 723].

Epithelioid sarcoma

Epithelioid sarcoma is an ultrarare high-grade soft tissue sarcoma with clinicopathological complexities that predisposes patients to locoregional recurrence [726], accounting for 1.2 ~ 1.5% of STS cases [727]. INI1 (encoded by the SMARCB1 tumor suppressor gene) is a core subunit of the switch/sucrose nonfermentable (SWI/SNF) chromatin remodeling complex frequently inactivated in epithelioid sarcomas [728]. The SWI/SNF complex is a crucial regulator of nucleosome positioning, frequently located at sites marked by acetylated histone H3 lysine 27 (H3K27ac), which establishes an open chromatin state with other transcription factors for transcriptional activation [729]. Enhancer of zeste homolog 2 (EZH2) is an enzymatic subunit of polycomb repressor complex 2 (PRC2) that negatively regulates the activity of the SWI/SNF complex by placing the repressive trimethylated histone H3 lysine 27 (H3K27me3) mark [729]. EZH2 is expressed in approximately one-third of epithelioid sarcomas [730], making EZH2 a promising target. Tazemetostat is an inhibitor of EZH2 that blocks the lysine methyltransferase activity of EZH2 by selectively binding to the S-adenosyl methionine (SAM) binding site of EZH2 [731] (Fig. 9h). Tazemetostat exhibits clinical activity with favorable safety in patients with advanced epithelioid sarcoma harboring INI1 loss [732, 733].

Brain and other CNS tumors accounted for 1.6% of cancer cases and 2.5% of cancer-related mortalities worldwide in 2020 [2]. These tumors comprise over 100 histologically distinct subtypes with varying clinical characteristics, treatments, and outcomes, mainly including tumors of neuroepithelial tissue, cranial and spinal nerves, meninges, etc. [734].

Glioblastoma

Glioblastoma multiforme (GBM) is the most frequent and lethal subtype of brain cancer that originates in the CNS [735]. Compared with surrounding healthy tissue, brain cancers possess a more alkaline pH [736]. Temozolomide is a DNA alkylating prodrug stable at acidic pH values but labile at alkaline pH values [737] (Fig. 9i). The alkaline microenvironment within brain cancer preferentially facilitates the activation of temozolomide [736, 737]. It adds a methyl group to the O6 position of guanine (G), resulting in a methyl-guanine (meG)-to-thymine (T) mismatch during DNA replication instead of a G-to-cytosine (C) match [738], which leads to DNA damage and ultimately cell apoptosis [739, 740]. Temozolomide exhibits an acceptable safety profile and improves PFS compared with procarbazine in patients with GBM at first relapse [741]. Currently, temozolomide-containing regimens are still the first-line therapy for GBM [742, 743]. However, O6-meG methyltransferase (MGMT) removes alkyl groups from the O6 position of G, conferring resistance to temozolomide [744, 745].

Neuroblastoma

Neuroblastoma is a malignant embryonal tumor derived from primitive cells of the sympathetic nervous system [746, 747]. It is the most frequent and lethal solid tumor in children and is commonly associated with a poor overall prognosis [747]. Disialoganglioside GD2 is expressed almost uniformly on the surface of neuroblastoma cells and induces cell proliferation, invasion, and motility by activating RTK-mediated signal transduction [746, 748, 749], making it an effective and tractable target of neuroblastoma [750, 751]. Dinutuximab is a human/mouse chimeric GD2-directed mAb that recognizes and binds to the sugar moiety of GD2 exposed to the extracellular milieu (similar to 14G2a antibody [752]), thereby inducing cell lysis through ADCC and CDC [751753]. Compared with standard therapy (six cycles of 13-cis-retinoic acid), dinutuximab significantly improved clinical outcomes in combination with alternating granulocyte-macrophage colony-stimulating factor (GM-CSF) and IL-2 after standard therapy in patients with high-risk neuroblastoma [754]. However, anti-drug antibodies, including human anti-mouse or -chimeric antibodies, may cause treatment delays, terminations, or even abrogate the antitumor effects [746]. Naxitamab is designed to reduce the effects of anti-drug antibodies but enhance ADCC through humanized IgG1-Fc and retain complement-mediated cytotoxicity potency through its high affinity for GD2 [746, 755]. As expected, naxitamab exhibits modest toxic effects, low immunogenicity, and substantial anti-neuroblastoma activity in combination with GM-CSF in patients with relapsed or refractory high-risk neuroblastoma [756758] (Fig. 9j).

Malignant pleural mesothelioma

Malignant pleural mesothelioma (MPM) is a rare but highly aggressive and lethal cancer that originates in the serosal outer linings of the lungs (pleurae), heart, abdomen, and testes, with a 5-year OS rate of ~ 5% [759, 760]. Folate receptors (FRα, FRβ, and FRγ) are cysteine-rich cell surface glycoproteins that mediate the cellular uptake of folate, commonly expressed at low levels in most normal tissues [761]. Folate-dependent one-carbon metabolism is required for the de novo synthesis of purines, thymidylate, and S-adenosyl methionine and is thus critical to DNA synthesis [762]. Nevertheless, FRα is highly activated and overexpressed in MPM tissues compared with normal adjacent tissues, making the folate–FRα delivery and metabolism system an attractive target for MPM treatment [763]. Pemetrexed is a multitargeted anti-folate agent that inhibits at least three enzymes (thymidylate synthase, dihydrofolate reductase, and glycinamide ribonucleotide formyltransferase) involved in folate metabolism and DNA synthesis [764] (Fig. 9k). Compared to cisplatin monotherapy, pemetrexed plus cisplatin therapy improved the mOS (12.1 months vs. 9.3 months) and was thus approved by the FDA for unresectable MPM treatment [765]. Most recently, compared with cisplatin-pemetrexed chemotherapy, durvalumab plus platinum-pemetrexed chemotherapy therapy significantly improved the mOS (20.4 months vs. 12.1 months) in patients with unresectable MPM [766].

NTRK-positive solid tumors

TRK proteins (including TRKA, TRKB, and TRKC) are encoded by the NTRK gene family (NTRK1-3), which are frequently fusion-positive in a broad range of solid tumors, including glioblastoma, NSCLC, and STSs. [767]. NTRK fusion leads to the constitutive activation of TRK protein, which acts as an oncogenic driver, making it a potential therapeutic target [768]. Larotrectinib is an oral, highly selective inhibitor of TRKs that binds to and competitively inhibits the ATP-binding site of TRKs [769] (Fig. 9l). Larotrectinib induces both cell apoptosis and inhibition of cell growth in TRK-overexpressed tumors [770] and exhibits encouraging antitumor activity with good tolerance in patients with tumors harboring NTRK gene fusions [771773]. However, NTRK1G595R and NTRK1G667C mutations located in the catalytic domain confer resistance to both entrectinib and larotrectinib [117, 774]. Ponatinib and nintedanib (a PDGFR, FGFR, and VEGFR inhibitor used for idiopathic pulmonary fibrosis treatment) potentially overcome NTRK1G667C mutation-induced resistance but not NTRK1G595R mutation-induced resistance [774]. Moreover, the next-generation TRK inhibitors repotrectinib and LOXO-195 exhibit encouraging activity to overcome TRK mutation-induced resistance [775] (Table (Table88).

Table 8

FDA-approved therapeutic drugs for thyroid cancer and other solid tumors

Drug (brand name)SponsorPropertiesIndicationApproval dateReview
Thyrotropin alfa (Thyrogen)GenzymeRecombinant human thyroid-stimulating hormoneThyroid cancer11/30/1998O
Lenvatinib mesylate (Lenvima)EisaiMultitarget TKI (VEGFRs, FGFRs, PDGFRα, RET, and KIT)Thyroid cancer02/13/2015P, O
Vandetanib (Caprelsa)GenzymeMultitarget TKI (VEGFR2/3, EGFR, and RET)Unresectable or metastatic medullary thyroid cancer04/06/2011P, O
Cabozantinib S-malate (Cometriq)ExelixisMultitarget TKI (VEGFRs, MET, RET, FLT3, KIT, TIE2, and AXL)Progressive, metastatic medullary thyroid cancer11/29/2012P, O
Alitretinoin (Panretin)Concordia9-cis-retinoic acid, a form of vitamin AAIDS-related Kaposi’s sarcoma02/02/1999P, O
Trabectedin (Yondelis)JanssenAlkylating drugLiposarcoma or leiomyosarcoma10/23/2015P, O
Olaratumab (Lartruvo)Eli LillyPDGFRα-directed mAbSoft tissue sarcoma10/19/2016P, O
Pexidartinib HCl (Turalio)Daiichi SankyoCSF1R, KIT, and FLT3 inhibitorTenosynovial giant cell tumor08/02/2019P, O
Tazemetostat (Tazverik)EpizymeEZH2 inhibitorEpithelioid sarcoma01/23/2020P, O
Temozolomide (Temodar)MerckDNA alkylating agentGlioblastoma08/11/1999P, O
Dinutuximab (Unituxin)United TherapGD2-directed mAbHigh-risk neuroblastoma03/10/2015P, O
Naxitamab (Danyelza)Y-mAbsGD2-directed mAbHigh-risk neuroblastoma11/25/2020P, O
Pemetrexed disodium (Alimta)Eli LillyFolate analogMalignant pleural mesothelioma02/04/2004P, O
Larotrectinib sulfate (Vitrakvi)BayerTRKs inhibitorNTRK-positive solid tumors11/26/2018P, O

AIDS Acquired immunodeficiency syndrome; CSF1R Colony-stimulating factor-1 receptor; EGFR Epidermal growth factor receptor; EZH2 Enhancer of zeste homolog 2; FLT3 FMS-like tyrosine kinase 3; MEK1/2 MAPK/ERK kinase 1/2; MET Mesenchymal–epithelial transition gene; NTRK Neurotrophic receptor tyrosine kinase; O Orphan; P Priority; PDGFRα Platelet-derived growth factor receptor α; RET: rearranged during transfection gene; S Standard; TRKs Tropomyosin receptor kinases; VEGFR Vascular endothelial growth factor receptor

The success and dilemma of current antitumor strategies

RTK inhibitors and immune checkpoint blockades (ICBs) have undoubtedly been the most successful antitumor drugs in the past 31 years. The human RTK family comprises 58 RTK proteins, which fall into 20 subfamilies [776]. These RTKs share a similar structure, mainly with ligand-binding domains in the extracellular region, a single transmembrane helix, and a tyrosine kinase domain in the cytoplasmic region [776] (Fig. 10). Aberrant overexpression and oncogenic gain-of-function mutation-induced ligand-independent activation of RTKs frequently leads to the activation of downstream pathways, resulting in various diseases involving cancers, diabetes, inflammation, etc. RTK-targeted therapies can occur at three levels: blocking the ligand–RTK interaction in the extracellular region, inhibiting the tyrosine kinase domain in the intracellular region, and inhibiting the constitutive components of RTK-mediated downstream pathways.

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig10_HTML.jpg

RTK families (adapted from [776])

The advent of trastuzumab is undoubtedly a milestone. It inhibited the RTK pathway from the first level and was the first RTK-targeted therapy. However, obstacles to trastuzumab–HER2 interaction [201] and reactivation of HER2 downstream pathways, whether induced by bypass pathway switching or mutations of downstream components (e.g., PIK3CA mutation [777]), confer resistance to trastuzumab. Regarding other RTKs, such as MET, oncogenic mutations lead to MET self-activation in a ligand-independent manner [778]. These biological mechanisms inevitably lead to the failure of the first-level RTK-targeted strategy. Gefitinib is a small-molecule inhibitor that targets the intracellular tyrosine kinase domain of RTK at the second level because it is difficult for antibodies to target intracellular antigens [779]. This strategy addresses the ligand-dependent activation and specific mutation-induced self-activation of RTK to a certain extent. However, it cannot overcome the bypass pathway switch, secondary mutations within the tyrosine kinase domain, and downstream component mutations, even if multitarget TKIs (e.g., sorafenib) are used. Inhibitors of the RAS–RAF–MEK–MAPK/ERK (e.g., sotorasib) and PI3K–AKT–mTOR (e.g., alpelisib) pathways block the RTK pathway at the third level. This strategy blocks the RTK pathway regardless of upstream RTK activation and may address the bypass pathway switch to a certain extent. However, the secondary mutations of targets and loss-of-function PTEN mutations still confer resistance [124, 780]. Nevertheless, RTK-targeted drugs have been the mainstay for the treatment of solid tumors. Over the past 31 years, 48 RTK inhibitors and 13 RTK downstream component inhibitors were approved by the FDA, and these drugs account for more than half of all therapeutic drugs for solid tumors (Fig. 11a).

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig11_HTML.jpg

Distribution of therapeutic drugs according to targets and approval years. a Distribution of therapeutic drugs according to targets. b Distribution of therapeutic drugs according to approval years

ICBs adopt a novel strategy that reinvigorates a range of CD4+ and CD8+ tumor-infiltrating T lymphocytes [781], enabling the possibility of long-term survival in patients with metastatic or advanced cancers [782]. The clinical application of ICBs heralds a new era of cancer treatment, as they are the most successful strategy in the recent decade [782, 783]. The FDA has approved nine ICBs in the USA since the first approval of ipilimumab in 2011 (Additional file 1: Table S4, page 56). Despite the clinical success, only a minority of people exhibit durable responses to ICBs [784]. The mutated proteins of cancer cells produced by nonsynonymous mutations and other genetic alterations need to be processed and then presented as neoantigens by major histocompatibility complex (MHC) molecules of antigen-presenting cells (APCs) and recognized by T cells [785]. However, neoantigens do not always bind to MHC molecules with high affinity or contain mutant amino acids at the appropriate position, making it difficult for T cells to recognize them [784]. Melanoma has the highest frequency of somatic mutations among human cancers and may produce the largest available neoantigen repertoire [785]. It explains why most ICBs are approved for the treatment of melanoma. In addition, preexisting PD1/PDL1-positive CD4+ and CD8+ T cells positioned in proximity to the cancer cells inside tumors are critical to clinical responses [786, 787]. In some tumors with an ‘immune-excluded’ phenotype, the T cells locate the stroma surrounding the tumor nest instead of penetrating the parenchyma of the tumor. Tumors with the ‘immune-desert’ phenotype lack T cells in either the parenchyma or stroma of the tumor [784]. Thus, tumors with immune-excluded and immune-desert phenotypes are often associated with unfavorable responses to ICBs [783]. In addition, immune-related adverse events (irAEs) [788] and hyperprogressive disease [5, 789] are of great concern. It is clear that further work is needed to reliably regulate the immune system in the clinic.

Breast and prostate cancers are associated with sex hormones and accounted for approximately one-fifth of cancer cases and more than 10% of cancer-related mortalities worldwide in 2020 [2]. Breast cancer drugs are frequently at the forefront of advances in cancer treatment and diagnosis [133]. Therapeutic drugs for breast cancer have begun to diversify, and no new drugs targeting ER (two SERDs bazedoxifene and ospemifene, are not indicated for breast cancer, Additional file 1: Table S1, page 26) or aromatase have been approved since the approval of fulvestrant in 2002. ER-positive breast cancer accounts for 80% of all breast cancer cases and half of breast cancer-related mortalities [152, 790]. Given the superiority of fulvestrant, newer-generation ER antagonists are needed to improve the poor physicochemical properties and administration mode of fulvestrant for this large group of patients [152].

In contrast, therapeutic drugs for prostate cancer are still limited to antiandrogens, even in recent years. The progression of mCRPC is the major cause of death in patients with prostate cancer [791], although OS is significantly improved with cabazitaxel [477, 478], abiraterone [486], and enzalutamide [792]. Bipolar androgen therapy (BAT) is a new strategy that induces rapid cycling between high and low serum testosterone concentrations, resulting in tumor responses and resensitization of mCRPC to enzalutamide. This strategy is more effective than abiraterone [793, 794]. Distinct strategies have been developed for the two sex hormone-related cancers; specifically, breast cancer treatment adopts strategies referring to multiple targets and mechanisms, while prostate cancer treatment emphasizes the refinement of antiandrogen strategies.

Therapeutic drugs for solid tumors have ushered in a new period of prosperity. Seventy-four therapeutic drugs and 61 RTK or RTK pathway inhibitors were approved in the last decade, accounting for 61.7% and 75.4% of all therapeutic drugs and RTK or RTK pathway inhibitors of solid tumors approved in the past 31 years, respectively (Fig. 11b). Quite a few drugs have been exquisitely designed. For instance, ziv-aflibercept utilizes the binding affinity between VEGFRs and VEGFs to capture VEGFs. In addition, ADCs retain all the antitumor efficiency of mAbs and add cytotoxic payloads, allowing for the targeted delivery of chemotherapeutic agents. The application of ADCs has dramatically expanded the clinical application of mAbs. ADCs and the first bispecific antibody, amivantamab, have started a new era of engineered antibodies. The approval of SMO, PARP, and EZH2 inhibitors was based on research progress on hedgehog signaling, synthetic lethality, and epigenetics in cancers. It is believed that there will be more drugs based on new mechanisms in the future alongside the exploration of new targets and vulnerabilities of tumors.

Future perspectives

Target identification and drug design have been the core drivers throughout antitumor history in recent decades, and antitumor strategies for solid tumors have profoundly changed over the past 30 years. During the first decade, pharmacologists were devoted to developing anti-endocrine agents, microtubule inhibitors, DNA alkylating agents, and DNA topoisomerase inhibitors. Overall, this stage did not focus much on targeting drugs, although the advent of trastuzumab began a new era of RTK-targeted therapy. During the second decade, pharmacologists extended RTK-targeted inhibitor studies to include RTK downstream component inhibitors, which enriched the TKI library and shifted the focus toward targeted drug development. In addition, the advent of ipilimumab, which converts immunotherapy from positive stimulation (e.g., IL-2 and INFα) to immune checkpoint blockade, started a true paradigm shift for metastatic or advanced solid tumors. During the third decade, RTK and RTK pathway inhibitors and ICBs were extensively developed. Drugs targeting novel targets and tumor vulnerabilities, such as PARP and SMO inhibitors, were added to the list for solid tumor treatment. KRASG12C, once considered an undruggable target, was blocked successfully by sotorasib. The treatment of solid tumors ushered in the precise targeting stage (Fig. 12a).

An external file that holds a picture, illustration, etc.
Object name is 13045_2022_1362_Fig12_HTML.jpg

Achievements in the past 30 years and future perspectives. a The achievements in the past 30 years (obtained from www.rcsb.org). b Future perspectives of oncology drug development (compositional element was obtained from https://www.16pic.com)

RTK and RTK pathway inhibitors, ADCs and ICBs, are still the mainstay. A new ICB relatlimab-rmbw (lymphocyte activation gene-3 (LAG-3)-directed mAb) was approved by the FDA in combination with nivolumab for unresectable or metastatic melanoma on March 18, 2022 [795]. Next, proteolysis-targeting chimeras (PROTACs) [796] and small interfering RNA (siRNA) technologies [797] degrade targets at the protein and RNA levels, respectively [798]. Indeed, the first RNA-targeted drug, inclisiran (Additional file 1: Table S1, page 40), has been approved by the FDA [799]. Increasing clinical trials of PROTAC-based drugs are ongoing, making PROTACs the gold rush [800]. The advent of PROTAC technology makes it possible to selectively degrade proteins that are typically difficult to target (e.g., transcription factors). Similar technologies, such as chaperone-mediated autophagy [801], Trim-Away [802], degradation tag (dTAG) [803], and lysosome-targeting chimeras (LYTACs) [804], are also of great concern. Multispecific antibodies (msAbs) bind two or more epitopes, which greatly extends the function of mAbs. With the approval of bispecific T-cell engagers (BiTEs) blinatumomab (Additional file 1: Table S1, page 28; Table S2, page 47) and amivantamab, msAbs will become a critical antitumor strategy in the coming decades [805].

Vaccines, cell-based therapies, and gene therapy products represent another essential pillar of cancer treatment, although they are not discussed in the text. Chimeric antigen receptor (CAR)-T cells have achieved great success in patients with hematological malignancies, especially CD19-directed CAR-T cells [806]. The clinical application of CAR-T cells in solid tumors has been limited by setbacks due to substantive biological barriers and risks [807]. Efforts to seek suitable targets [808] to overcome the immunosuppressive tumor microenvironment (TME) [809] and combat cytokine release syndrome (CRS) and immune effector cell-associated neurotoxicity syndrome (ICANS) [810, 811] are still ongoing. In recent years, CAR-T-cell clinical trials against solid tumors have exhibited acceptable safety and encouraging clinical outcomes [812815]. Oncolytic viruses are naturally or genetically engineered viruses that preferentially infect, lyse, and replicate in cancer cells relative to normal cells [816, 817]. Oncolytic viruses provide a platform for monotherapy [818] or in combination with chemotherapies [819, 820] and immunotherapies [821823] by delivering defined factors. With the first approval of talimogene laherparepvec (T-Vec) in 2015 [824], many clinical trials are ongoing [816]. Oncolytic viruses are also attractive carriers for cancer vaccines [825] and gene editing [826].

Cancer vaccines can be simply divided into preventive and therapeutic vaccines [827]. Therapeutic vaccines directly utilize APCs (e.g., dendritic cells (DCs)), viruses, liposomes, and nanoparticles as vesicles to deliver tumor-specific antigens (including neoantigens), inducing immune recognition and activation of T cells [828]. Preventive vaccines are confined to specific virus-induced cancers, such as HPV-related cancers [829] and hepatitis B virus (HBV)-related HCC [830]. Like preventive vaccines, chemoprevention is also a preventive strategy to reverse, suppress, or prevent carcinogenic progression to invasive cancer using chemical agents [831]. For instance, familial adenomatous polyposis (FAP) is a precancerous state of colorectal cancer [832] caused by germline mutations in the adenomatous polyposis coli (APC) gene [833, 834]. Almost all of the mutations of APC, both germline and somatic, produce a truncated APC protein, leading to APC dysfunction [835837]. Dysfunctional APC fails to form a destruction complex, resulting in β-Catenin stabilization and canonical Wnt/β-Catenin signaling activation [838]. Cyclooxygenase-2 (COX-2) is a crucial enzyme of prostaglandin E2 (PGE2) biosynthesis that plays an essential role in colorectal tumorigenesis [839]. PGE2 is a potent proinflammatory factor that serves as a ligand for the G protein-coupled receptor (GPCR) EP2. It promotes colon cancer cell growth through the Gαs-Axin-β-Catenin axis [840]. Celecoxib is a potent COX-2 inhibitor approved by the FDA in 1998 for treating FAP (Additional file 1: Table S1, page 10); thus, it is also an agent for the chemoprevention of colorectal cancer [841, 842]. From the cost-effectiveness perspective, preventive vaccines and chemoprevention have absolute superiorities, both economically and physiologically.

Cell death inducers have always been an important research field in cancer treatment strategies [843]. Mechanically, the available antitumor drugs induce cell cycle arrest or cell death unexceptionally. For instance, the mTOR inhibitors (e.g., temsirolimus and everolimus) can be classified as autophagy-related death inducers. In recent years, novel cell death inducers, such as the tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) agonist eftozanermin alfa (ABBV-621) [844] and the mitochondrial caseinolytic protease P (ClpP) activator ONC201 [845] have entered clinical trials for the treatment of solid tumors (NCT03082209 and NCT05476939), which may bring new hope for cancer treatment. In contrast, significant success has been achieved in the field of epigenetic drugs (epi-drugs), such as EZH2 inhibitor tazemetostat and isocitrate dehydrogenase 1 (IDH1) inhibitor ivosidenib (Additional file 1: Table S1, page 34). The first- and second-generation epi-drugs that use a ‘one size fits all’ strategy, such as DNA methyltransferase (DNMT) and histone deacetylase (HDAC) inhibitors, were proven to have disappointing efficacy in patients with solid tumors [846]. The third-generation epi-drugs use more precise targets, such as IDH1, EZH2, and certain bromodomain and extra-terminal domain (BET)-containing proteins (BRDs), which are showing promising efficacy [847].

Artificial intelligence (AI) improves the ability to deal with the massive amount of tumor genome information and promotes the ability to decipher protein structures. The AI technology represented by AlphaFold may significantly shorten the process of drug development [848]. In addition, gene-editing technologies, such as clustered regularly interspaced short palindromic repeats (CRISPR)-associated (Cas) systems [849], provide a potent tool to modify primary patient-derived cells in vitro. Quite a few clinical trials of CRISPR-based immune cells for cancer treatment are ongoing, especially CAR-T cells [850]. However, multiple hurdles need to be overcome before CRISPR directly targets tumor cells in vivo, including appropriate delivery carriers, off-target cutting, and chromosomal rearrangements [850, 851]. Modifying specific mutations by gene-editing technologies is undoubtedly one of the peaks of precision medicine. Engineered cell therapies should not be limited to the currently used T cell or DC cell models; many other cell types can also be incorporated into this system, such as stem cells [852, 853], natural killer (NK) cells [854], fibroblasts [855], and even engineered cancer cells [856]. Expanding the variety of cell types available for therapy can make full use of the characteristics of different cells to meet complex clinical needs [857]. Human microbiome communities have been implicated in cancer initiation, progression, metastasis, and therapy resistance [858]. With the advent of next-generation sequencing and a deeper understanding of host–microbiome interactions, microbiome analyses are being developed as a promising approach for cancer diagnosis [859], while microbiome modulation may be a practicable adjunct to existing antitumor strategies [860].

High tumor heterogeneity and tumor mutation burden are frequently associated with treatment resistance and cancer recurrence [861], failing to predict the response to available treatment [862]. For these clinical settings, systematic manipulation and domestication of cancer cells by extracellular matrix (ECM) and epigenetic remodeling or by more complicated metabolic and immune remodeling to control the progression and metastasis of tumors instead of killing tumors may be realistic strategies in the post-precision medicine era. With the development of AI and nanotechnology, the existing approaches to diagnosis and treatment will be replaced by dynamic nanosensors and intelligent nanobots, thereby promoting the transition from precision medicine to intelligent medicine (Fig. 12b).

Conclusion

The research and development pace of antitumor drugs is accelerating with the in-depth study of the tumorigenesis mechanism. Nevertheless, these 120 therapeutic drugs are still the mainstay for advanced, unresectable, or metastatic solid tumors. Although several drugs have been discontinued or withdrawn from the market due to severe adverse effects, commercial reasons, or the emergence of substituted new-generation drugs, the findings and lessons in the exploration of cancer treatment strategies will always be the milestones in antitumor history.

Acknowledgements

Not applicable.

Abbreviations

177LuLutetium-177
223RaRadium-223
5′DFUR5′-Deoxy-5-fluorouridine
5-FU5-Fluorouracil
ADCCAntibody-dependent cellular cytotoxicity
ADCsAntibody-drug conjugates
ADTAndrogen deprivation therapy
AIArtificial intelligence
AIDSAcquired immunodeficiency syndrome
ALKAnaplastic lymphoma kinase
APCAntigen-presenting cell or adenomatous polyposis coli
ARAndrogen receptor
AREsAndrogen response elements
BATBipolar androgen therapy
BCGBacillus Calmette-Guérin
BERBase excision repair
BETBromodomain and extra-terminal domain
BiTEsBispecific T cell engagers
BLAsBiologics license applications
BLBCBasal-like breast cancer
BRDsBET-containing proteins
CARChimeric antigen receptor
CDCComplement-dependent cytotoxicity
CDK4/6Cyclin-dependent kinases 4/6
CISCarcinoma in situ
ClpPCaseinolytic protease P
CNSCentral nervous system
COX-2Cyclooxygenase-2
CRISPRClustered regularly interspaced short palindromic repeats
CRSCytokine release syndrome
CSCCCutaneous squamous cell carcinoma
CSF1Colony-stimulating factor-1
CSF1RCSF1 receptor
CTLA4Cytotoxic T lymphocyte antigen 4
CYP17A1Cytochrome P450 17A1
DCsDendritic cells
DFSDisease-free survival
DHTDihydrotestosterone
dMMRMismatch repair deficiency
DNMTDNA methyltransferase
DSBsDouble-stranded breaks
dTAGDegradation tag
DTCDifferentiated thyroid cancer
ECMExtracellular matrix
EGFREpidermal growth factor receptor
EML4EMAP-like protein 4
epi-drugsEpigenetic drugs
EREstrogen receptor
ERKExtracellular signal-regulated kinase
EZH2Zeste homolog 2
FAPFamilial adenomatous polyposis
FDAFood and Drug Administration
FdUMPFluorodeoxyuridine monophosphate
FdUTPFluorodeoxyuridine triphosphate
FGFR2Fibroblast growth factor receptor 2
FKBP12FK506-binding protein 12
FRFolate receptor
FSHFollicle-stimulating hormone
FUTPFluorouridine triphosphate
GATA3GATA binding protein 3
GBMGlioblastoma multiforme
GEP-NETsGastroenteropancreatic neuroendocrine tumors
GISTsGastrointestinal stromal tumors
GLIGlioma-associated oncogene
GM-CSFGranulocyte-macrophage colony-stimulating factor
GnRHGonadotropin-releasing hormone
gp100Glycoprotein 100
GPCRG protein-coupled receptor
H3K27acAcetylated histone H3 lysine27
H3K27me3Trimethylated histone H3 lysine27
HBVHepatitis B virus
HCCHepatocellular carcinoma
HDACHistone deacetylase
HER2Epidermal growth factor receptor 2
HGFHepatocyte growth factor
HIF1αHypoxia-inducible factor-1α
HPVHuman papillomavirus
HRHormone receptor
HRDHomologous recombination deficiency
HSP90Heat shock protein 90
ICANSImmune effector cell-associated neurotoxicity syndrome
ICBsImmune checkpoint blockades
ICCIntrahepatic cholangiocarcinoma
IDH1Isocitrate dehydrogenase 1
IFNαInterferon-α
IFNγInterferon-γ
IgImmunoglobulin
IGF1RInsulin-like growth factor-1 receptor
IL-2Interleukin-2
IL-2RIL-2 receptor
irAEsImmune-related adverse events
KSHVKaposi’s sarcoma-associated herpesvirus
LAG-3Lymphocyte activation gene-3
LAKLymphokine-activated killer
LBDLigand-binding domain
LCCLarge-cell carcinoma
LHLuteinizing hormone
LH-RHLuteinizing hormone-releasing hormone
LYTACsLysosome-targeting chimaeras
mAbsMonoclonal antibodies
MAPKMitogen-activated protein kinase
MASCMammary analog secretory carcinoma
mCRPCMetastatic castration-resistant prostate cancer
MC-vc-PABMaleimidocaproyl valine-citrulline p-aminobenzyl alcohol carbamate
MDRABCG2-mediated multidrug resistance
MEKMAPK/ERK kinase
METMesenchymal–epithelial transition
MGMTO6-meG methyltransferase
MHCMajor histocompatibility complex
mLST8Mammalian lethal with SEC13 protein 8
MMAEMonomethyl auristatin E
MPMMalignant pleural mesothelioma
MRP2Multidrug resistance-associated protein 2
msAbsMultispecific antibodies
MSI-HMicrosatellite instability—high
MTCMedullary thyroid cancer
mTORMammalian target of rapamycin
mTORCMTOR complex
NAD+Nicotinamide adenine dinucleotide
NERNucleotide excision repair
NISSodium–iodide symporter
NKNatural killer
NMEsNew molecular entities
NMPANational Medical Products Administration
NMSCsNon-melanoma skin cancers
NSCLCNon-small-cell lung cancer
NTRKNeurotrophic tyrosine receptor kinase
OCTOrganic cation transporter
ORRObjective response rate
OSOverall survival
PARPsPoly (ADP-ribose) polymerases
PARPoly (ADP-ribose)
PAR-2Protease-activated receptor 2
PD1Programmed death receptor-1
PDGFRα/βPlatelet-derived growth factor receptor α/β
PDL1Programmed death-ligand 1
PFSProgression-free survival
PGE2Prostaglandin E2
PI3KPhosphatidylinositol 3-kinase
PIK3CAPhosphatidylinositol 3-kinase catalytic subunit A
PlGFPlacenta growth factor
PRProgesterone receptor
PRC2Polycomb repressor complex 2
PROTACsProteolysis-targeting chimeras
PRRTPeptide receptor radionuclide therapy
PSAProstate-specific antigen
PTCH1Patched homolog 1
RAPTORRegulatory-associated protein of mTOR
RARγRetinoic acid receptor γ
RBRetinoblastoma
RB3-SLDRB3 protein stathmin-like domain
RCCRenal cell carcinoma
RETRearranged during transfection
rhTSHRecombinant human TSH
RICTORRapamycin-insensitive companion of mTOR
ROSReactive oxygen species
ROS1ROS proto-oncogene 1
RTKsReceptor tyrosine kinases
RXRαRetinoid X receptor α
SAMS-adenosyl methionine
SERDSelective ER degrader/down-regulator
SERMsSelective ER modulators
SCCSquamous cell carcinoma
SCLCSmall-cell lung cancer
siRNASmall interfering RNA
SMCCN-succinimidyl-4-(N-maleimidomethyl) cyclohexane-1-carboxylate
SMOSmoothened
SSBsDNA single-stranded breaks
SSTRsSomatostatin receptors
SWI/SNFSwitch/sucrose non-fermentable
T-VecTalimogene laherparepvec
TFTissue factor
TIIsTumor-infiltrating immune cells
TIMCsTumor-infiltrating mononuclear cells
TKIsTyrosine kinase inhibitors
TMETumor microenvironment
TNBCTriple-negative breast cancer
TOP1Topoisomerase I
TOP1CCsTOP1 cleavage complexes
TOP2ATopoisomerase Iiα
TPThymidine phosphorylase
TRAILTumor necrosis factor-related apoptosis inducing ligand
TRKsTropomyosin receptor kinases
Trop-2Trophoblastic cell surface antigen-2
TSHThyroid-stimulating hormone
UGT1A1Uridine diphosphate glucuronosyltransferase 1A1
UPUridine phosphorylase
VEGFR2Vascular endothelial growth factor receptor 2
VEGFsVascular endothelial growth factors
VHLVon Hippel–Lindau; WT: Wild type

Author contributions

SJ drafted the manuscript with QW, WQ, and XS; SJ conceived the idea, constructed the figures, and revised the manuscript. All authors read and approved the final manuscript.

Funding

This study was supported by the Zhejiang Provincial Natural Science Foundation of China (LY21C070002 to SJ) and the Basic Research Project of Hangzhou Medical College (KYQN202007 to QW).

Availability of data and materials

Not applicable.

Declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors have declared that no competing interests exist.

Footnotes

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Qing Wu, Wei Qian, and Xiaoli Sun contributed equally to this work

Contributor Information

Qing Wu, nc.ude.cmh@6280qw.

Wei Qian, nc.ude.ujz@0104152.

Xiaoli Sun, nc.ude.ujz@hzylxs.

Shaojie Jiang, nc.ude.cmh@3891jsj.

References

1. Bray F, Laversanne M, Weiderpass E, Soerjomataram I. The ever-increasing importance of cancer as a leading cause of premature death worldwide. Cancer. 2021;127(16):3029–3030. 10.1002/cncr.33587. [Abstract] [CrossRef] [Google Scholar]
2. Sung H, Ferlay J, Siegel RL, Laversanne M, Soerjomataram I, Jemal A, Bray F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2021;71(3):209–249. 10.3322/caac.21660. [Abstract] [CrossRef] [Google Scholar]
3. Vasan N, Baselga J, Hyman DM. A view on drug resistance in cancer. Nature. 2019;575(7782):299–309. 10.1038/s41586-019-1730-1. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
4. Nurgali K, Jagoe RT, Abalo R. Editorial: adverse effects of cancer chemotherapy: anything new to improve tolerance and reduce sequelae? Front Pharmacol. 2018;9:245. 10.3389/fphar.2018.00245. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
5. Champiat S, Dercle L, Ammari S, Massard C, Hollebecque A, Postel-Vinay S, Chaput N, Eggermont A, Marabelle A, Soria JC, et al. Hyperprogressive disease is a new pattern of progression in cancer patients treated by anti-PD-1/PD-L1. Clin Cancer Res. 2017;23(8):1920–1928. 10.1158/1078-0432.CCR-16-1741. [Abstract] [CrossRef] [Google Scholar]
6. Gridelli C, Rossi A, Carbone DP, Guarize J, Karachaliou N, Mok T, Petrella F, Spaggiari L, Rosell R. Non-small-cell lung cancer. Nat Rev Dis Primers. 2015;1:15009. 10.1038/nrdp.2015.9. [Abstract] [CrossRef] [Google Scholar]
7. Helbekkmo N, Sundstrom SH, Aasebo U, Brunsvig PF, von Plessen C, Hjelde HH, Garpestad OK, Bailey A, Bremnes RM, Norwegian Lung Cancer Study G Vinorelbine/carboplatin vs gemcitabine/carboplatin in advanced NSCLC shows similar efficacy, but different impact of toxicity. Br J Cancer. 2007;97(3):283–289. 10.1038/sj.bjc.6603869. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
8. Lindeman NI, Cagle PT, Beasley MB, Chitale DA, Dacic S, Giaccone G, Jenkins RB, Kwiatkowski DJ, Saldivar JS, Squire J, et al. Molecular testing guideline for selection of lung cancer patients for EGFR and ALK tyrosine kinase inhibitors: guideline from the college of American pathologists, international association for the study of lung cancer, and association for molecular pathology. Arch Pathol Lab Med. 2013;137(6):828–860. 10.5858/arpa.2012-0720-OA. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
9. Gainor JF, Varghese AM, Ou SH, Kabraji S, Awad MM, Katayama R, Pawlak A, Mino-Kenudson M, Yeap BY, Riely GJ, et al. ALK rearrangements are mutually exclusive with mutations in EGFR or KRAS: an analysis of 1683 patients with non-small cell lung cancer. Clin Cancer Res. 2013;19(15):4273–4281. 10.1158/1078-0432.CCR-13-0318. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
10. Gigant B, Wang C, Ravelli RB, Roussi F, Steinmetz MO, Curmi PA, Sobel A, Knossow M. Structural basis for the regulation of tubulin by vinblastine. Nature. 2005;435(7041):519–522. 10.1038/nature03566. [Abstract] [CrossRef] [Google Scholar]
11. Wieczorek M, Tcherkezian J, Bernier C, Prota AE, Chaaban S, Rolland Y, Godbout C, Hancock MA, Arezzo JC, Ocal O, et al. The synthetic diazonamide DZ-2384 has distinct effects on microtubule curvature and dynamics without neurotoxicity. Sci Transl Med. 2016;8(365):365ra159. 10.1126/scitranslmed.aag1093. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
12. Capasso A. Vinorelbine in cancer therapy. Curr Drug Targets. 2012;13(8):1065–1071. 10.2174/138945012802009017. [Abstract] [CrossRef] [Google Scholar]
13. Shi Y, Au JS, Thongprasert S, Srinivasan S, Tsai CM, Khoa MT, Heeroma K, Itoh Y, Cornelio G, Yang PC. A prospective, molecular epidemiology study of EGFR mutations in Asian patients with advanced non-small-cell lung cancer of adenocarcinoma histology (PIONEER) J Thorac Oncol. 2014;9(2):154–162. 10.1097/JTO.0000000000000033. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
14. Barlesi F, Mazieres J, Merlio JP, Debieuvre D, Mosser J, Lena H, Ouafik L, Besse B, Rouquette I, Westeel V, et al. Routine molecular profiling of patients with advanced non-small-cell lung cancer: results of a 1-year nationwide programme of the French Cooperative Thoracic Intergroup (IFCT) Lancet. 2016;387(10026):1415–1426. 10.1016/S0140-6736(16)00004-0. [Abstract] [CrossRef] [Google Scholar]
15. Rosell R, Moran T, Queralt C, Porta R, Cardenal F, Camps C, Majem M, Lopez-Vivanco G, Isla D, Provencio M, et al. Screening for epidermal growth factor receptor mutations in lung cancer. N Engl J Med. 2009;361(10):958–967. 10.1056/NEJMoa0904554. [Abstract] [CrossRef] [Google Scholar]
16. Murray S, Dahabreh IJ, Linardou H, Manoloukos M, Bafaloukos D, Kosmidis P. Somatic mutations of the tyrosine kinase domain of epidermal growth factor receptor and tyrosine kinase inhibitor response to TKIs in non-small cell lung cancer: an analytical database. J Thorac Oncol. 2008;3(8):832–839. 10.1097/JTO.0b013e31818071f3. [Abstract] [CrossRef] [Google Scholar]
17. Recondo G, Facchinetti F, Olaussen KA, Besse B, Friboulet L. Making the first move in EGFR-driven or ALK-driven NSCLC: first-generation or next-generation TKI? Nat Rev Clin Oncol. 2018;15(11):694–708. 10.1038/s41571-018-0081-4. [Abstract] [CrossRef] [Google Scholar]
18. Ji H, Li D, Chen L, Shimamura T, Kobayashi S, McNamara K, Mahmood U, Mitchell A, Sun Y, Al-Hashem R, et al. The impact of human EGFR kinase domain mutations on lung tumorigenesis and in vivo sensitivity to EGFR-targeted therapies. Cancer Cell. 2006;9(6):485–495. 10.1016/j.ccr.2006.04.022. [Abstract] [CrossRef] [Google Scholar]
19. Novello S, Barlesi F, Califano R, Cufer T, Ekman S, Levra MG, Kerr K, Popat S, Reck M, Senan S, et al. Metastatic non-small-cell lung cancer: ESMO clinical practice guidelines for diagnosis, treatment and follow-up. Ann Oncol. 2016;27(suppl 5):v1–v27. 10.1093/annonc/mdw326. [Abstract] [CrossRef] [Google Scholar]
20. Lynch TJ, Bell DW, Sordella R, Gurubhagavatula S, Okimoto RA, Brannigan BW, Harris PL, Haserlat SM, Supko JG, Haluska FG, et al. Activating mutations in the epidermal growth factor receptor underlying responsiveness of non-small-cell lung cancer to gefitinib. N Engl J Med. 2004;350(21):2129–2139. 10.1056/NEJMoa040938. [Abstract] [CrossRef] [Google Scholar]
21. Shepherd FA, Rodrigues Pereira J, Ciuleanu T, Tan EH, Hirsh V, Thongprasert S, Campos D, Maoleekoonpiroj S, Smylie M, Martins R, et al. Erlotinib in previously treated non-small-cell lung cancer. N Engl J Med. 2005;353(2):123–132. 10.1056/NEJMoa050753. [Abstract] [CrossRef] [Google Scholar]
22. Wu YL, Zhou C, Hu CP, Feng J, Lu S, Huang Y, Li W, Hou M, Shi JH, Lee KY, 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(2):213–222. 10.1016/S1470-2045(13)70604-1. [Abstract] [CrossRef] [Google Scholar]
23. Wu YL, Cheng Y, Zhou X, Lee KH, Nakagawa K, Niho S, Tsuji F, Linke R, Rosell R, Corral J, et al. Dacomitinib versus gefitinib as first-line treatment for patients with EGFR-mutation-positive non-small-cell lung cancer (ARCHER 1050): a randomised, open-label, phase 3 trial. Lancet Oncol. 2017;18(11):1454–1466. 10.1016/S1470-2045(17)30608-3. [Abstract] [CrossRef] [Google Scholar]
24. Soria JC, Ohe Y, Vansteenkiste J, Reungwetwattana T, Chewaskulyong B, Lee KH, Dechaphunkul A, Imamura F, Nogami N, Kurata T, et al. Osimertinib in untreated EGFR-Mutated advanced non-small-cell lung cancer. N Engl J Med. 2018;378(2):113–125. 10.1056/NEJMoa1713137. [Abstract] [CrossRef] [Google Scholar]
25. Riely GJ, Pao W, Pham D, Li AR, Rizvi N, Venkatraman ES, Zakowski MF, Kris MG, Ladanyi M, Miller VA. Clinical course of patients with non-small cell lung cancer and epidermal growth factor receptor exon 19 and exon 21 mutations treated with gefitinib or erlotinib. Clin Cancer Res. 2006;12(3 Pt 1):839–844. 10.1158/1078-0432.CCR-05-1846. [Abstract] [CrossRef] [Google Scholar]
26. Li D, Ambrogio L, Shimamura T, Kubo S, Takahashi M, Chirieac LR, Padera RF, Shapiro GI, Baum A, Himmelsbach F, et al. BIBW2992, an irreversible EGFR/HER2 inhibitor highly effective in preclinical lung cancer models. Oncogene. 2008;27(34):4702–4711. 10.1038/onc.2008.109. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
27. Pao W, Miller VA, Politi KA, Riely GJ, Somwar R, Zakowski MF, Kris MG, Varmus H. Acquired resistance of lung adenocarcinomas to gefitinib or erlotinib is associated with a second mutation in the EGFR kinase domain. PLoS Med. 2005;2(3):e73. 10.1371/journal.pmed.0020073. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
28. Yun CH, Mengwasser KE, Toms AV, Woo MS, Greulich H, Wong KK, Meyerson M, Eck MJ. The T790M mutation in EGFR kinase causes drug resistance by increasing the affinity for ATP. Proc Natl Acad Sci U S A. 2008;105(6):2070–2075. 10.1073/pnas.0709662105. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
29. Engelman JA, Zejnullahu K, Gale CM, Lifshits E, Gonzales AJ, Shimamura T, Zhao F, Vincent PW, Naumov GN, Bradner JE, et al. PF00299804, an irreversible pan-ERBB inhibitor, is effective in lung cancer models with EGFR and ERBB2 mutations that are resistant to gefitinib. Cancer Res. 2007;67(24):11924–11932. 10.1158/0008-5472.CAN-07-1885. [Abstract] [CrossRef] [Google Scholar]
30. Ward RA, Anderton MJ, Ashton S, Bethel PA, Box M, Butterworth S, Colclough N, Chorley CG, Chuaqui C, Cross DA, et al. Structure- and reactivity-based development of covalent inhibitors of the activating and gatekeeper mutant forms of the epidermal growth factor receptor (EGFR) J Med Chem. 2013;56(17):7025–7048. 10.1021/jm400822z. [Abstract] [CrossRef] [Google Scholar]
31. Zhou W, Ercan D, Chen L, Yun CH, Li D, Capelletti M, Cortot AB, Chirieac L, Iacob RE, Padera R, et al. Novel mutant-selective EGFR kinase inhibitors against EGFR T790M. Nature. 2009;462(7276):1070–1074. 10.1038/nature08622. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
32. Cross DA, Ashton SE, Ghiorghiu S, Eberlein C, Nebhan CA, Spitzler PJ, Orme JP, Finlay MR, Ward RA, Mellor MJ, et al. AZD9291, an irreversible EGFR TKI, overcomes T790M-mediated resistance to EGFR inhibitors in lung cancer. Cancer Discov. 2014;4(9):1046–1061. 10.1158/2159-8290.CD-14-0337. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
33. Ou SI, Agarwal N, Ali SM. High MET amplification level as a resistance mechanism to osimertinib (AZD9291) in a patient that symptomatically responded to crizotinib treatment post-osimertinib progression. Lung Cancer. 2016;98:59–61. 10.1016/j.lungcan.2016.05.015. [Abstract] [CrossRef] [Google Scholar]
34. Thress KS, Paweletz CP, Felip E, Cho BC, Stetson D, Dougherty B, Lai Z, Markovets A, Vivancos A, Kuang Y, et al. Acquired EGFR C797S mutation mediates resistance to AZD9291 in non-small cell lung cancer harboring EGFR T790M. Nat Med. 2015;21(6):560–562. 10.1038/nm.3854. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
35. Kuenen B, Witteveen PO, Ruijter R, Giaccone G, Dontabhaktuni A, Fox F, Katz T, Youssoufian H, Zhu J, Rowinsky EK, et al. A phase I pharmacologic study of necitumumab (IMC-11F8), a fully human IgG1 monoclonal antibody directed against EGFR in patients with advanced solid malignancies. Clin Cancer Res. 2010;16(6):1915–1923. 10.1158/1078-0432.CCR-09-2425. [Abstract] [CrossRef] [Google Scholar]
36. Garnock-Jones KP. Necitumumab: first global approval. Drugs. 2016;76(2):283–289. 10.1007/s40265-015-0537-0. [Abstract] [CrossRef] [Google Scholar]
37. Bagchi A, Haidar JN, Eastman SW, Vieth M, Topper M, Iacolina MD, Walker JM, Forest A, Shen Y, Novosiadly RD, et al. Molecular basis for necitumumab inhibition of EGFR variants associated with acquired cetuximab resistance. Mol Cancer Ther. 2018;17(2):521–531. 10.1158/1535-7163.MCT-17-0575. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
38. Vyse S, Huang PH. Targeting EGFR exon 20 insertion mutations in non-small cell lung cancer. Signal Transduct Target Ther. 2019;4:5. 10.1038/s41392-019-0038-9. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
39. Gonzalvez F, Vincent S, Baker TE, Gould AE, Li S, Wardwell SD, Nadworny S, Ning Y, Zhang S, Huang WS, et al. Mobocertinib (TAK-788): a targeted inhibitor of EGFR Exon 20 insertion mutants in non-small cell lung cancer. Cancer Discov. 2021;11(7):1672–1687. 10.1158/2159-8290.CD-20-1683. [Abstract] [CrossRef] [Google Scholar]
40. Arcila ME, Nafa K, Chaft JE, Rekhtman N, Lau C, Reva BA, Zakowski MF, Kris MG, Ladanyi M. EGFR exon 20 insertion mutations in lung adenocarcinomas: prevalence, molecular heterogeneity, and clinicopathologic characteristics. Mol Cancer Ther. 2013;12(2):220–229. 10.1158/1535-7163.MCT-12-0620. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
41. Oxnard GR, Lo PC, Nishino M, Dahlberg SE, Lindeman NI, Butaney M, Jackman DM, Johnson BE, Janne PA. Natural history and molecular characteristics of lung cancers harboring EGFR exon 20 insertions. J Thorac Oncol. 2013;8(2):179–184. 10.1097/JTO.0b013e3182779d18. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
42. Kobayashi Y, Mitsudomi T. Not all epidermal growth factor receptor mutations in lung cancer are created equal: perspectives for individualized treatment strategy. Cancer Sci. 2016;107(9):1179–1186. 10.1111/cas.12996. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
43. Riess JW, Gandara DR, Frampton GM, Madison R, Peled N, Bufill JA, Dy GK, Ou SI, Stephens PJ, McPherson JD, et al. Diverse EGFR exon 20 insertions and co-occurring molecular alterations identified by comprehensive genomic profiling of NSCLC. J Thorac Oncol. 2018;13(10):1560–1568. 10.1016/j.jtho.2018.06.019. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
44. Yun J, Lee SH, Kim SY, Jeong SY, Kim JH, Pyo KH, Park CW, Heo SG, Yun MR, Lim S, et al. Antitumor activity of amivantamab (JNJ-61186372), an EGFR-MET bispecific antibody, in diverse models of EGFR exon 20 insertion-driven NSCLC. Cancer Discov. 2020;10(8):1194–1209. 10.1158/2159-8290.CD-20-0116. [Abstract] [CrossRef] [Google Scholar]
45. Hirose T, Ikegami M, Endo M, Matsumoto Y, Nakashima Y, Mano H, Kohsaka S. Extensive functional evaluation of exon 20 insertion mutations of EGFR. Lung Cancer. 2021;152:135–142. 10.1016/j.lungcan.2020.12.023. [Abstract] [CrossRef] [Google Scholar]
46. Yasuda H, Park E, Yun CH, Sng NJ, Lucena-Araujo AR, Yeo WL, Huberman MS, Cohen DW, Nakayama S, Ishioka K, et al. Structural, biochemical, and clinical characterization of epidermal growth factor receptor (EGFR) exon 20 insertion mutations in lung cancer. Sci Transl Med. 2013;5(216):216ra177. 10.1126/scitranslmed.3007205. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
47. Lee Y, Kim TM, Kim DW, Kim S, Kim M, Keam B, Ku JL, Heo DS. Preclinical modeling of osimertinib for NSCLC with EGFR exon 20 insertion mutations. J Thorac Oncol. 2019;14(9):1556–1566. 10.1016/j.jtho.2019.05.006. [Abstract] [CrossRef] [Google Scholar]
48. Ji J, Aredo JV, Piper-Vallillo A, Huppert L, Rotow JK, Husain H, Stewart SL, Cobb R, Wakelee HA, Blakely CM, et al. Osimertinib in non-small cell lung cancer (NSCLC) with atypical EGFR activating mutations: a retrospective multicenter study. J Clin Oncol. 2020;38(15_suppl):9570–9570. 10.1200/JCO.2020.38.15_suppl.9570. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
49. Remon J, Hendriks LEL, Cardona AF, Besse B. EGFR exon 20 insertions in advanced non-small cell lung cancer: a new history begins. Cancer Treat Rev. 2020;90:102105. 10.1016/j.ctrv.2020.102105. [Abstract] [CrossRef] [Google Scholar]
50. Moores SL, Chiu ML, Bushey BS, Chevalier K, Luistro L, Dorn K, Brezski RJ, Haytko P, Kelly T, Wu SJ, et al. A novel bispecific antibody targeting EGFR and cMet is effective against EGFR inhibitor-resistant lung tumors. Cancer Res. 2016;76(13):3942–3953. 10.1158/0008-5472.CAN-15-2833. [Abstract] [CrossRef] [Google Scholar]
51. Labrijn AF, Meesters JI, de Goeij BE, van den Bremer ET, Neijssen J, van Kampen MD, Strumane K, Verploegen S, Kundu A, Gramer MJ, et al. Efficient generation of stable bispecific IgG1 by controlled Fab-arm exchange. Proc Natl Acad Sci USA. 2013;110(13):5145–5150. 10.1073/pnas.1220145110. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
52. Neijssen J, Cardoso RMF, Chevalier KM, Wiegman L, Valerius T, Anderson GM, Moores SL, Schuurman J, Parren P, Strohl WR, et al. Discovery of amivantamab (JNJ-61186372), a bispecific antibody targeting EGFR and MET. J Biol Chem. 2021;296:100641. 10.1016/j.jbc.2021.100641. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
53. Syed YY. Amivantamab: first approval. Drugs. 2021;81(11):1349–1353. 10.1007/s40265-021-01561-7. [Abstract] [CrossRef] [Google Scholar]
54. Grugan KD, Dorn K, Jarantow SW, Bushey BS, Pardinas JR, Laquerre S, Moores SL, Chiu ML. Fc-mediated activity of EGFR x c-Met bispecific antibody JNJ-61186372 enhanced killing of lung cancer cells. MAbs. 2017;9(1):114–126. 10.1080/19420862.2016.1249079. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
55. Park K, Haura EB, Leighl NB, Mitchell P, Shu CA, Girard N, Viteri S, Han JY, Kim SW, Lee CK, et al. Amivantamab in EGFR Exon 20 insertion-mutated non-small-cell lung cancer progressing on platinum chemotherapy: initial results from the CHRYSALIS phase I study. J Clin Oncol. 2021;39(30):3391–3402. 10.1200/JCO.21.00662. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
56. Robichaux JP, Elamin YY, Tan Z, Carter BW, Zhang S, Liu S, Li S, Chen T, Poteete A, Estrada-Bernal A, et al. Mechanisms and clinical activity of an EGFR and HER2 exon 20-selective kinase inhibitor in non-small cell lung cancer. Nat Med. 2018;24(5):638–646. 10.1038/s41591-018-0007-9. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
57. Markham A. Mobocertinib: first approval. Drugs. 2021;81(17):2069–2074. 10.1007/s40265-021-01632-9. [Abstract] [CrossRef] [Google Scholar]
58. Riely GJ, Neal JW, Camidge DR, Spira AI, Piotrowska Z, Costa DB, Tsao AS, Patel JD, Gadgeel SM, Bazhenova L, et al. Activity and safety of mobocertinib (TAK-788) in previously treated non-small cell lung cancer with EGFR exon 20 insertion mutations from a phase I/II trial. Cancer Discov. 2021;11(7):1688–1699. 10.1158/2159-8290.CD-20-1598. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
59. Zhou C, Ramalingam SS, Kim TM, Kim SW, Yang JC, Riely GJ, Mekhail T, Nguyen D, Garcia Campelo MR, Felip E, et al. Treatment outcomes and safety of mobocertinib in platinum-pretreated patients with EGFR exon 20 insertion-positive metastatic non-small cell lung cancer: a phase 1/2 open-label nonrandomized clinical trial. JAMA Oncol. 2021;7(12):e214761. 10.1001/jamaoncol.2021.4761. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
60. Soda M, Choi YL, Enomoto M, Takada S, Yamashita Y, Ishikawa S, Fujiwara S, Watanabe H, Kurashina K, Hatanaka H, et al. Identification of the transforming EML4-ALK fusion gene in non-small-cell lung cancer. Nature. 2007;448(7153):561–566. 10.1038/nature05945. [Abstract] [CrossRef] [Google Scholar]
61. Salido M, Pijuan L, Martinez-Aviles L, Galvan AB, Canadas I, Rovira A, Zanui M, Martinez A, Longaron R, Sole F, et al. Increased ALK gene copy number and amplification are frequent in non-small cell lung cancer. J Thorac Oncol. 2011;6(1):21–27. 10.1097/JTO.0b013e3181fb7cd6. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
62. Horn L, Pao W. EML4-ALK: honing in on a new target in non-small-cell lung cancer. J Clin Oncol. 2009;27(26):4232–4235. 10.1200/JCO.2009.23.6661. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
63. Shaw AT, Yeap BY, Mino-Kenudson M, Digumarthy SR, Costa DB, Heist RS, Solomon B, Stubbs H, Admane S, McDermott U, et al. Clinical features and outcome of patients with non-small-cell lung cancer who harbor EML4-ALK. J Clin Oncol. 2009;27(26):4247–4253. 10.1200/JCO.2009.22.6993. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
64. Solomon BJ, Mok T, Kim DW, Wu YL, Nakagawa K, Mekhail T, Felip E, Cappuzzo F, Paolini J, Usari T, et al. First-line crizotinib versus chemotherapy in ALK-positive lung cancer. N Engl J Med. 2014;371(23):2167–2177. 10.1056/NEJMoa1408440. [Abstract] [CrossRef] [Google Scholar]
65. Shaw AT, Kim DW, Nakagawa K, Seto T, Crino L, Ahn MJ, De Pas T, Besse B, Solomon BJ, Blackhall F, et al. Crizotinib versus chemotherapy in advanced ALK-positive lung cancer. N Engl J Med. 2013;368(25):2385–2394. 10.1056/NEJMoa1214886. [Abstract] [CrossRef] [Google Scholar]
66. Shaw AT, Kim DW, Mehra R, Tan DS, Felip E, Chow LQ, Camidge DR, Vansteenkiste J, Sharma S, De Pas T, et al. Ceritinib in ALK-rearranged non-small-cell lung cancer. N Engl J Med. 2014;370(13):1189–1197. 10.1056/NEJMoa1311107. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
67. Peters S, Camidge DR, Shaw AT, Gadgeel S, Ahn JS, Kim DW, Ou SI, Perol M, Dziadziuszko R, Rosell R, et al. Alectinib versus crizotinib in untreated ALK-positive non-small-cell lung cancer. N Engl J Med. 2017;377(9):829–838. 10.1056/NEJMoa1704795. [Abstract] [CrossRef] [Google Scholar]
68. Camidge DR, Kim HR, Ahn MJ, Yang JC, Han JY, Lee JS, Hochmair MJ, Li JY, Chang GC, Lee KH, et al. Brigatinib versus crizotinib in ALK-positive non-small-cell lung cancer. N Engl J Med. 2018;379(21):2027–2039. 10.1056/NEJMoa1810171. [Abstract] [CrossRef] [Google Scholar]
69. Shaw AT, Friboulet L, Leshchiner I, Gainor JF, Bergqvist S, Brooun A, Burke BJ, Deng YL, Liu W, Dardaei L, et al. Resensitization to crizotinib by the lorlatinib ALK resistance mutation L1198F. N Engl J Med. 2016;374(1):54–61. 10.1056/NEJMoa1508887. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
70. Doebele RC, Pilling AB, Aisner DL, Kutateladze TG, Le AT, Weickhardt AJ, Kondo KL, Linderman DJ, Heasley LE, Franklin WA, et al. Mechanisms of resistance to crizotinib in patients with ALK gene rearranged non-small cell lung cancer. Clin Cancer Res. 2012;18(5):1472–1482. 10.1158/1078-0432.CCR-11-2906. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
71. Katayama R, Shaw AT, Khan TM, Mino-Kenudson M, Solomon BJ, Halmos B, Jessop NA, Wain JC, Yeo AT, Benes C, et al. Mechanisms of acquired crizotinib resistance in ALK-rearranged lung Cancers. Sci Transl Med. 2012;4(120):120ra117. 10.1126/scitranslmed.3003316. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
72. Gainor JF, Dardaei L, Yoda S, Friboulet L, Leshchiner I, Katayama R, Dagogo-Jack I, Gadgeel S, Schultz K, Singh M, et al. Molecular mechanisms of resistance to first- and second-generation ALK inhibitors in ALK-rearranged lung cancer. Cancer Discov. 2016;6(10):1118–1133. 10.1158/2159-8290.CD-16-0596. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
73. Sasaki T, Koivunen J, Ogino A, Yanagita M, Nikiforow S, Zheng W, Lathan C, Marcoux JP, Du J, Okuda K, et al. A novel ALK secondary mutation and EGFR signaling cause resistance to ALK kinase inhibitors. Cancer Res. 2011;71(18):6051–6060. 10.1158/0008-5472.CAN-11-1340. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
74. Friboulet L, Li N, Katayama R, Lee CC, Gainor JF, Crystal AS, Michellys PY, Awad MM, Yanagitani N, Kim S, et al. The ALK inhibitor ceritinib overcomes crizotinib resistance in non-small cell lung cancer. Cancer Discov. 2014;4(6):662–673. 10.1158/2159-8290.CD-13-0846. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
75. Zou HY, Friboulet L, Kodack DP, Engstrom LD, Li Q, West M, Tang RW, Wang H, Tsaparikos K, Wang J, et al. PF-06463922, an ALK/ROS1 inhibitor, overcomes resistance to first and second generation ALK inhibitors in preclinical models. Cancer Cell. 2015;28(1):70–81. 10.1016/j.ccell.2015.05.010. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
76. Lin JJ, Zhu VW, Yoda S, Yeap BY, Schrock AB, Dagogo-Jack I, Jessop NA, Jiang GY, Le LP, Gowen K, et al. Impact of EML4-ALK variant on resistance mechanisms and clinical outcomes in ALK-positive lung cancer. J Clin Oncol. 2018;36(12):1199–1206. 10.1200/JCO.2017.76.2294. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
77. Recondo G, Mezquita L, Facchinetti F, Planchard D, Gazzah A, Bigot L, Rizvi AZ, Frias RL, Thiery JP, Scoazec JY, et al. Diverse resistance mechanisms to the third-generation ALK inhibitor lorlatinib in ALK-rearranged lung cancer. Clin Cancer Res. 2020;26(1):242–255. 10.1158/1078-0432.CCR-19-1104. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
78. Awad MM, Oxnard GR, Jackman DM, Savukoski DO, Hall D, Shivdasani P, Heng JC, Dahlberg SE, Janne PA, Verma S, et al. MET exon 14 mutations in non-small-cell lung cancer are associated with advanced age and stage-dependent MET genomic amplification and c-met overexpression. J Clin Oncol. 2016;34(7):721–730. 10.1200/JCO.2015.63.4600. [Abstract] [CrossRef] [Google Scholar]
79. Schrock AB, Frampton GM, Suh J, Chalmers ZR, Rosenzweig M, Erlich RL, Halmos B, Goldman J, Forde P, Leuenberger K, et al. Characterization of 298 patients with lung cancer harboring MET exon 14 skipping alterations. J Thorac Oncol. 2016;11(9):1493–1502. 10.1016/j.jtho.2016.06.004. [Abstract] [CrossRef] [Google Scholar]
80. Tong JH, Yeung SF, Chan AW, Chung LY, Chau SL, Lung RW, Tong CY, Chow C, Tin EK, Yu YH, et al. MET amplification and exon 14 splice site mutation define unique molecular subgroups of non-small cell lung carcinoma with poor prognosis. Clin Cancer Res. 2016;22(12):3048–3056. 10.1158/1078-0432.CCR-15-2061. [Abstract] [CrossRef] [Google Scholar]
81. Onozato R, Kosaka T, Kuwano H, Sekido Y, Yatabe Y, Mitsudomi T. Activation of MET by gene amplification or by splice mutations deleting the juxtamembrane domain in primary resected lung cancers. J Thorac Oncol. 2009;4(1):5–11. 10.1097/JTO.0b013e3181913e0e. [Abstract] [CrossRef] [Google Scholar]
82. Cappuzzo F, Marchetti A, Skokan M, Rossi E, Gajapathy S, Felicioni L, Del Grammastro M, Sciarrotta MG, Buttitta F, Incarbone M, et al. Increased MET gene copy number negatively affects survival of surgically resected non-small-cell lung cancer patients. J Clin Oncol. 2009;27(10):1667–1674. 10.1200/JCO.2008.19.1635. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
83. Schildhaus HU, Schultheis AM, Ruschoff J, Binot E, Merkelbach-Bruse S, Fassunke J, Schulte W, Ko YD, Schlesinger A, Bos M, et al. MET amplification status in therapy-naive adeno- and squamous cell carcinomas of the lung. Clin Cancer Res. 2015;21(4):907–915. 10.1158/1078-0432.CCR-14-0450. [Abstract] [CrossRef] [Google Scholar]
84. Wolf J, Seto T, Han JY, Reguart N, Garon EB, Groen HJM, Tan DSW, Hida T, de Jonge M, Orlov SV, et al. Capmatinib in MET exon 14-mutated or MET-amplified non-small-cell lung cancer. N Engl J Med. 2020;383(10):944–957. 10.1056/NEJMoa2002787. [Abstract] [CrossRef] [Google Scholar]
85. Dhillon S. Capmatinib: first approval. Drugs. 2020;80(11):1125–1131. 10.1007/s40265-020-01347-3. [Abstract] [CrossRef] [Google Scholar]
86. Paez JG, Janne PA, Lee JC, Tracy S, Greulich H, Gabriel S, Herman P, Kaye FJ, Lindeman N, Boggon TJ, et al. EGFR mutations in lung cancer: correlation with clinical response to gefitinib therapy. Science. 2004;304(5676):1497–1500. 10.1126/science.1099314. [Abstract] [CrossRef] [Google Scholar]
87. Liu X, Wang Q, Yang G, Marando C, Koblish HK, Hall LM, Fridman JS, Behshad E, Wynn R, Li Y, et al. A novel kinase inhibitor, INCB28060, blocks c-MET-dependent signaling, neoplastic activities, and cross-talk with EGFR and HER-3. Clin Cancer Res. 2011;17(22):7127–7138. 10.1158/1078-0432.CCR-11-1157. [Abstract] [CrossRef] [Google Scholar]
88. Baltschukat S, Engstler BS, Huang A, Hao HX, Tam A, Wang HQ, Liang J, DiMare MT, Bhang HC, Wang Y, et al. Capmatinib (INC280) is active against models of non-small cell lung cancer and other cancer types with defined mechanisms of MET activation. Clin Cancer Res. 2019;25(10):3164–3175. 10.1158/1078-0432.CCR-18-2814. [Abstract] [CrossRef] [Google Scholar]
89. Wu YL, Smit EF, Bauer TM. Capmatinib for patients with non-small cell lung cancer with MET exon 14 skipping mutations: a review of preclinical and clinical studies. Cancer Treat Rev. 2021;95:102173. 10.1016/j.ctrv.2021.102173. [Abstract] [CrossRef] [Google Scholar]
90. Fujino T, Kobayashi Y, Suda K, Koga T, Nishino M, Ohara S, Chiba M, Shimoji M, Tomizawa K, Takemoto T, et al. Sensitivity and resistance of MET exon 14 mutations in lung cancer to eight MET tyrosine kinase inhibitors in vitro. J Thorac Oncol. 2019;14(10):1753–1765. 10.1016/j.jtho.2019.06.023. [Abstract] [CrossRef] [Google Scholar]
91. Bladt F, Faden B, Friese-Hamim M, Knuehl C, Wilm C, Fittschen C, Gradler U, Meyring M, Dorsch D, Jaehrling F, et al. EMD 1214063 and EMD 1204831 constitute a new class of potent and highly selective c-met inhibitors. Clin Cancer Res. 2013;19(11):2941–2951. 10.1158/1078-0432.CCR-12-3247. [Abstract] [CrossRef] [Google Scholar]
92. Paik PK, Felip E, Veillon R, Sakai H, Cortot AB, Garassino MC, Mazieres J, Viteri S, Senellart H, Van Meerbeeck J, et al. Tepotinib in non-small-cell lung cancer with MET exon 14 skipping mutations. N Engl J Med. 2020;383(10):931–943. 10.1056/NEJMoa2004407. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
93. Mathieu LN, Larkins E, Akinboro O, Roy P, Amatya AK, Fiero MH, Mishra-Kalyani PS, Helms WS, Myers CE, Skinner AM, et al. FDA approval summary: capmatinib and tepotinib for the treatment of metastatic NSCLC harboring MET exon 14 skipping mutations or alterations. Clin Cancer Res. 2022;28(2):249–254. 10.1158/1078-0432.CCR-21-1566. [Abstract] [CrossRef] [Google Scholar]
94. Wu YL, Cheng Y, Zhou J, Lu S, Zhang Y, Zhao J, Kim DW, Soo RA, Kim SW, Pan H, et al. Tepotinib plus gefitinib in patients with EGFR-mutant non-small-cell lung cancer with MET overexpression or MET amplification and acquired resistance to previous EGFR inhibitor (INSIGHT study): an open-label, phase 1b/2, multicentre, randomised trial. Lancet Respir Med. 2020;8(11):1132–1143. 10.1016/S2213-2600(20)30154-5. [Abstract] [CrossRef] [Google Scholar]
95. Drilon A, Oxnard GR, Tan DSW, Loong HHF, Johnson M, Gainor J, McCoach CE, Gautschi O, Besse B, Cho BC, et al. Efficacy of selpercatinib in RET fusion-positive non-small-cell lung cancer. N Engl J Med. 2020;383(9):813–824. 10.1056/NEJMoa2005653. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
96. Drilon A, Lin JJ, Filleron T, Ni A, Milia J, Bergagnini I, Hatzoglou V, Velcheti V, Offin M, Li B, et al. Frequency of brain metastases and multikinase inhibitor outcomes in patients with RET-rearranged lung cancers. J Thorac Oncol. 2018;13(10):1595–1601. 10.1016/j.jtho.2018.07.004. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
97. Markham A. Selpercatinib: first approval. Drugs. 2020;80(11):1119–1124. 10.1007/s40265-020-01343-7. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
98. Markham A. Pralsetinib: first approval. Drugs. 2020;80(17):1865–1870. 10.1007/s40265-020-01427-4. [Abstract] [CrossRef] [Google Scholar]
99. Drusbosky LM, Rodriguez E, Dawar R, Ikpeazu CV. Therapeutic strategies in RET gene rearranged non-small cell lung cancer. J Hematol Oncol. 2021;14(1):50. 10.1186/s13045-021-01063-9. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
100. Rahal R, Maynard M, Hu W, Brubaker J, Cao Q, Kim JL, Sheets MP, Wilson DP, Wilson KJ, DiPietro L et al: Abstract B151: BLU-667 is a potent and highly selective RET inhibitor being developed forRET-driven cancers. In: Therapeutic agents: small-molecule kinase inhibitors. 2018; B151–B151.
101. Subbiah V, Velcheti V, Tuch BB, Ebata K, Busaidy NL, Cabanillas ME, Wirth LJ, Stock S, Smith S, Lauriault V, et al. Selective RET kinase inhibition for patients with RET-altered cancers. Ann Oncol. 2018;29(8):1869–1876. 10.1093/annonc/mdy137. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
102. Lin JJ, Liu SV, McCoach CE, Zhu VW, Tan AC, Yoda S, Peterson J, Do A, Prutisto-Chang K, Dagogo-Jack I, et al. Mechanisms of resistance to selective RET tyrosine kinase inhibitors in RET fusion-positive non-small-cell lung cancer. Ann Oncol. 2020;31(12):1725–1733. 10.1016/j.annonc.2020.09.015. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
103. Zhu VW, Madison R, Schrock AB, Ou SI. Emergence of high level of MET amplification as off-target resistance to selpercatinib treatment in KIF5B-RET NSCLC. J Thorac Oncol. 2020;15(7):e124–e127. 10.1016/j.jtho.2020.03.020. [Abstract] [CrossRef] [Google Scholar]
104. Rosen EY, Johnson ML, Clifford SE, Somwar R, Kherani JF, Son J, Bertram AA, Davare MA, Gladstone E, Ivanova EV, et al. Overcoming MET-dependent resistance to selective RET inhibition in patients with RET fusion-positive lung cancer by combining selpercatinib with crizotinib. Clin Cancer Res. 2021;27(1):34–42. 10.1158/1078-0432.CCR-20-2278. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
105. Zhu VW, Zhang SS, Zhang J, Swensen J, Xiu J, Ou SI. Acquired tertiary MET resistance (MET D1228N and a novel LSM8-MET fusion) to selpercatinib and capmatinib in a patient with KIF5B-RET-positive NSCLC with secondary MET amplification as initial resistance to selpercatinib. J Thorac Oncol. 2021;16(7):e51–e54. 10.1016/j.jtho.2021.03.006. [Abstract] [CrossRef] [Google Scholar]
106. Subbiah V, Shen T, Terzyan SS, Liu X, Hu X, Patel KP, Hu M, Cabanillas M, Behrang A, Meric-Bernstam F, et al. Structural basis of acquired resistance to selpercatinib and pralsetinib mediated by non-gatekeeper RET mutations. Ann Oncol. 2021;32(2):261–268. 10.1016/j.annonc.2020.10.599. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
107. Gainor JF, Shaw AT. Novel targets in non-small cell lung cancer: ROS1 and RET fusions. Oncologist. 2013;18(7):865–875. 10.1634/theoncologist.2013-0095. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
108. Shaw AT, Ou SH, Bang YJ, Camidge DR, Solomon BJ, Salgia R, Riely GJ, Varella-Garcia M, Shapiro GI, Costa DB, et al. Crizotinib in ROS1-rearranged non-small-cell lung cancer. N Engl J Med. 2014;371(21):1963–1971. 10.1056/NEJMoa1406766. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
109. Arbour KC, Riely GJ. Systemic therapy for locally advanced and metastatic non-small cell lung cancer: a review. JAMA. 2019;322(8):764–774. 10.1001/jama.2019.11058. [Abstract] [CrossRef] [Google Scholar]
110. Patil T, Smith DE, Bunn PA, Aisner DL, Le AT, Hancock M, Purcell WT, Bowles DW, Camidge DR, Doebele RC. The incidence of brain metastases in stage IV ROS1-rearranged non-small cell lung cancer and rate of central nervous system progression on crizotinib. J Thorac Oncol. 2018;13(11):1717–1726. 10.1016/j.jtho.2018.07.001. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
111. Tang SC, Nguyen LN, Sparidans RW, Wagenaar E, Beijnen JH, Schinkel AH. Increased oral availability and brain accumulation of the ALK inhibitor crizotinib by coadministration of the P-glycoprotein (ABCB1) and breast cancer resistance protein (ABCG2) inhibitor elacridar. Int J Cancer. 2014;134(6):1484–1494. 10.1002/ijc.28475. [Abstract] [CrossRef] [Google Scholar]
112. Dziadziuszko R, Krebs MG, De Braud F, Siena S, Drilon A, Doebele RC, Patel MR, Cho BC, Liu SV, Ahn MJ, et al. Updated integrated analysis of the efficacy and safety of entrectinib in locally advanced or metastatic ROS1 fusion-positive non-small-cell lung cancer. J Clin Oncol. 2021;39(11):1253–1263. 10.1200/JCO.20.03025. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
113. Awad MM, Katayama R, McTigue M, Liu W, Deng YL, Brooun A, Friboulet L, Huang D, Falk MD, Timofeevski S, et al. Acquired resistance to crizotinib from a mutation in CD74-ROS1. N Engl J Med. 2013;368(25):2395–2401. 10.1056/NEJMoa1215530. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
114. Al-Salama ZT, Keam SJ. Entrectinib: first global approval. Drugs. 2019;79(13):1477–1483. 10.1007/s40265-019-01177-y. [Abstract] [CrossRef] [Google Scholar]
115. Fischer H, Ullah M, de la Cruz CC, Hunsaker T, Senn C, Wirz T, Wagner B, Draganov D, Vazvaei F, Donzelli M, et al. Entrectinib, a TRK/ROS1 inhibitor with anti-CNS tumor activity: differentiation from other inhibitors in its class due to weak interaction with P-glycoprotein. Neuro Oncol. 2020;22(6):819–829. 10.1093/neuonc/noaa052. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
116. Doebele RC, Dziadziuszko R, Drilon A, Shaw A, Wolf J, Farago AF, Dennis L, Riehl T, Simmons B, Wu C, et al. Genomic landscape of entrectinib resistance from ctDNA analysis in STARTRK-2. Annal Oncol. 2019 10.1093/annonc/mdz394.017. [CrossRef] [Google Scholar]
117. Russo M, Misale S, Wei G, Siravegna G, Crisafulli G, Lazzari L, Corti G, Rospo G, Novara L, Mussolin B, et al. Acquired resistance to the TRK inhibitor entrectinib in colorectal cancer. Cancer Discov. 2016;6(1):36–44. 10.1158/2159-8290.CD-15-0940. [Abstract] [CrossRef] [Google Scholar]
118. Drilon A, Li G, Dogan S, Gounder M, Shen R, Arcila M, Wang L, Hyman DM, Hechtman J, Wei G, et al. What hides behind the MASC: clinical response and acquired resistance to entrectinib after ETV6-NTRK3 identification in a mammary analogue secretory carcinoma (MASC) Ann Oncol. 2016;27(5):920–926. 10.1093/annonc/mdw042. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
119. MacFarland SP, Naraparaju K, Iyer R, Guan P, Kolla V, Hu Y, Tan K, Brodeur GM. Mechanisms of entrectinib resistance in a neuroblastoma xenograft model. Mol Cancer Ther. 2020;19(3):920–926. 10.1158/1535-7163.MCT-18-1044. [Abstract] [CrossRef] [Google Scholar]
120. Baker NM, Der CJ. Cancer: Drug for an 'undruggable' protein. Nature. 2013;497(7451):577–578. 10.1038/nature12248. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
121. Skoulidis F, Li BT, Dy GK, Price TJ, Falchook GS, Wolf J, Italiano A, Schuler M, Borghaei H, Barlesi F, et al. Sotorasib for Lung Cancers with KRAS p.G12C Mutation. N Engl J Med. 2021;384(25):2371–2381. 10.1056/NEJMoa2103695. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
122. Canon J, Rex K, Saiki AY, Mohr C, Cooke K, Bagal D, Gaida K, Holt T, Knutson CG, Koppada N, et al. The clinical KRAS(G12C) inhibitor AMG 510 drives anti-tumour immunity. Nature. 2019;575(7781):217–223. 10.1038/s41586-019-1694-1. [Abstract] [CrossRef] [Google Scholar]
123. Lanman BA, Allen JR, Allen JG, Amegadzie AK, Ashton KS, Booker SK, Chen JJ, Chen N, Frohn MJ, Goodman G, et al. Discovery of a covalent inhibitor of KRAS(G12C) (AMG 510) for the treatment of solid tumors. J Med Chem. 2020;63(1):52–65. 10.1021/acs.jmedchem.9b01180. [Abstract] [CrossRef] [Google Scholar]
124. Awad MM, Liu S, Rybkin II, Arbour KC, Dilly J, Zhu VW, Johnson ML, Heist RS, Patil T, Riely GJ, et al. Acquired resistance to KRAS(G12C) inhibition in cancer. N Engl J Med. 2021;384(25):2382–2393. 10.1056/NEJMoa2105281. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
125. Suzuki S, Yonesaka K, Teramura T, Takehara T, Kato R, Sakai H, Haratani K, Tanizaki J, Kawakami H, Hayashi H, et al. KRAS inhibitor resistance in MET-amplified KRAS (G12C) non-small cell lung cancer induced By RAS- and non-RAS-mediated cell signaling mechanisms. Clin Cancer Res. 2021;27(20):5697–5707. 10.1158/1078-0432.CCR-21-0856. [Abstract] [CrossRef] [Google Scholar]
126. Rudin CM, Brambilla E, Faivre-Finn C, Sage J. Small-cell lung cancer. Nat Rev Dis Primers. 2021;7(1):3. 10.1038/s41572-020-00235-0. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
127. Pommier Y. Topoisomerase I inhibitors: camptothecins and beyond. Nat Rev Cancer. 2006;6(10):789–802. 10.1038/nrc1977. [Abstract] [CrossRef] [Google Scholar]
128. Thomas A, Pommier Y. Targeting topoisomerase I in the era of precision medicine. Clin Cancer Res. 2019;25(22):6581–6589. 10.1158/1078-0432.CCR-19-1089. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
129. Pommier Y, Sun Y, Huang SN, Nitiss JL. Roles of eukaryotic topoisomerases in transcription, replication and genomic stability. Nat Rev Mol Cell Biol. 2016;17(11):703–721. 10.1038/nrm.2016.111. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
130. Leal JF, Martinez-Diez M, Garcia-Hernandez V, Moneo V, Domingo A, Bueren-Calabuig JA, Negri A, Gago F, Guillen-Navarro MJ, Aviles P, et al. PM01183, a new DNA minor groove covalent binder with potent in vitro and in vivo anti-tumour activity. Br J Pharmacol. 2010;161(5):1099–1110. 10.1111/j.1476-5381.2010.00945.x. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
131. Markham A. Lurbinectedin: first approval. Drugs. 2020;80(13):1345–1353. 10.1007/s40265-020-01374-0. [Abstract] [CrossRef] [Google Scholar]
132. Giordano SH. Breast cancer in men. N Engl J Med. 2018;378(24):2311–2320. 10.1056/NEJMra1707939. [Abstract] [CrossRef] [Google Scholar]
133. Leo CP, Leo C, Szucs TD. Breast cancer drug approvals by the US FDA from 1949 to 2018. Nat Rev Drug Discov. 2020;19(1):11. 10.1038/d41573-019-00201-w. [Abstract] [CrossRef] [Google Scholar]
134. Mullard A. 2019 FDA drug approvals. Nat Rev Drug Discov. 2020;19(2):79–84. 10.1038/d41573-020-00001-7. [Abstract] [CrossRef] [Google Scholar]
135. Mullard A. 2020 FDA drug approvals. Nat Rev Drug Discov. 2021;20(2):85–90. 10.1038/d41573-021-00002-0. [Abstract] [CrossRef] [Google Scholar]
136. Diaz JF, Andreu JM. Assembly of purified GDP-tubulin into microtubules induced by taxol and taxotere: reversibility, ligand stoichiometry, and competition. Biochemistry. 1993;32(11):2747–2755. 10.1021/bi00062a003. [Abstract] [CrossRef] [Google Scholar]
137. Wang YF, Shi QW, Dong M, Kiyota H, Gu YC, Cong B. Natural taxanes: developments since 1828. Chem Rev. 2011;111(12):7652–7709. 10.1021/cr100147u. [Abstract] [CrossRef] [Google Scholar]
138. McKeage K. Docetaxel: a review of its use for the first-line treatment of advanced castration-resistant prostate cancer. Drugs. 2012;72(11):1559–1577. 10.2165/11209660-000000000-00000. [Abstract] [CrossRef] [Google Scholar]
139. Morse DL, Gray H, Payne CM, Gillies RJ. Docetaxel induces cell death through mitotic catastrophe in human breast cancer cells. Mol Cancer Ther. 2005;4(10):1495–1504. 10.1158/1535-7163.MCT-05-0130. [Abstract] [CrossRef] [Google Scholar]
140. Goodin S, Kane MP, Rubin EH. Epothilones: mechanism of action and biologic activity. J Clin Oncol. 2004;22(10):2015–2025. 10.1200/JCO.2004.12.001. [Abstract] [CrossRef] [Google Scholar]
141. Doodhi H, Prota AE, Rodriguez-Garcia R, Xiao H, Custar DW, Bargsten K, Katrukha EA, Hilbert M, Hua S, Jiang K, et al. Termination of protofilament elongation by eribulin induces lattice defects that promote microtubule catastrophes. Curr Biol. 2016;26(13):1713–1721. 10.1016/j.cub.2016.04.053. [Abstract] [CrossRef] [Google Scholar]
142. Metzger R, Danenberg K, Leichman CG, Salonga D, Schwartz EL, Wadler S, Lenz HJ, Groshen S, Leichman L, Danenberg PV. High basal level gene expression of thymidine phosphorylase (platelet-derived endothelial cell growth factor) in colorectal tumors is associated with nonresponse to 5-fluorouracil. Clin Cancer Res. 1998;4(10):2371–2376. [Abstract] [Google Scholar]
143. Miwa M, Ura M, Nishida M, Sawada N, Ishikawa T, Mori K, Shimma N, Umeda I, Ishitsuka H. Design of a novel oral fluoropyrimidine carbamate, capecitabine, which generates 5-fluorouracil selectively in tumours by enzymes concentrated in human liver and cancer tissue. Eur J Cancer. 1998;34(8):1274–1281. 10.1016/S0959-8049(98)00058-6. [Abstract] [CrossRef] [Google Scholar]
144. Cao D, Russell RL, Zhang D, Leffert JJ, Pizzorno G. Uridine phosphorylase (-/-) murine embryonic stem cells clarify the key role of this enzyme in the regulation of the pyrimidine salvage pathway and in the activation of fluoropyrimidines. Cancer Res. 2002;62(8):2313–2317. [Abstract] [Google Scholar]
145. Wagstaff AJ, Ibbotson T, Goa KL. Capecitabine: a review of its pharmacology and therapeutic efficacy in the management of advanced breast cancer. Drugs. 2003;63(2):217–236. 10.2165/00003495-200363020-00009. [Abstract] [CrossRef] [Google Scholar]
146. Longley DB, Harkin DP, Johnston PG. 5-fluorouracil: mechanisms of action and clinical strategies. Nat Rev Cancer. 2003;3(5):330–338. 10.1038/nrc1074. [Abstract] [CrossRef] [Google Scholar]
147. Plosker GL, Faulds D. Epirubicin. A review of its pharmacodynamic and pharmacokinetic properties, and therapeutic use in cancer chemotherapy. Drugs. 1993;45(5):788–856. 10.2165/00003495-199345050-00011. [Abstract] [CrossRef] [Google Scholar]
148. Coukell AJ, Faulds D. Epirubicin. An updated review of its pharmacodynamic and pharmacokinetic properties and therapeutic efficacy in the management of breast cancer. Drugs. 1997;53(3):453–482. 10.2165/00003495-199753030-00008. [Abstract] [CrossRef] [Google Scholar]
149. Ormrod D, Holm K, Goa K, Spencer C. Epirubicin: a review of its efficacy as adjuvant therapy and in the treatment of metastatic disease in breast cancer. Drugs Aging. 1999;15(5):389–416. 10.2165/00002512-199915050-00006. [Abstract] [CrossRef] [Google Scholar]
150. DeSantis CE, Ma J, Gaudet MM, Newman LA, Miller KD, Goding Sauer A, Jemal A, Siegel RL. Breast cancer statistics, 2019. CA Cancer J Clin. 2019;69(6):438–451. 10.3322/caac.21583. [Abstract] [CrossRef] [Google Scholar]
151. Jeselsohn R, Buchwalter G, De Angelis C, Brown M, Schiff R. ESR1 mutations-a mechanism for acquired endocrine resistance in breast cancer. Nat Rev Clin Oncol. 2015;12(10):573–583. 10.1038/nrclinonc.2015.117. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
152. Guan J, Zhou W, Hafner M, Blake RA, Chalouni C, Chen IP, De Bruyn T, Giltnane JM, Hartman SJ, Heidersbach A, et al. Therapeutic ligands antagonize estrogen receptor function by impairing its mobility. Cell. 2019;178(4):949–963 e918. 10.1016/j.cell.2019.06.026. [Abstract] [CrossRef] [Google Scholar]
153. Hanker AB, Sudhan DR, Arteaga CL. Overcoming endocrine resistance in breast cancer. Cancer Cell. 2020;37(4):496–513. 10.1016/j.ccell.2020.03.009. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
154. Ma CX, Reinert T, Chmielewska I, Ellis MJ. Mechanisms of aromatase inhibitor resistance. Nat Rev Cancer. 2015;15(5):261–275. 10.1038/nrc3920. [Abstract] [CrossRef] [Google Scholar]
155. Bulun SE, Sebastian S, Takayama K, Suzuki T, Sasano H, Shozu M. The human CYP19 (aromatase P450) gene: update on physiologic roles and genomic organization of promoters. J Steroid Biochem Mol Biol. 2003;86(3–5):219–224. 10.1016/S0960-0760(03)00359-5. [Abstract] [CrossRef] [Google Scholar]
156. Simpson ER, Clyne C, Rubin G, Boon WC, Robertson K, Britt K, Speed C, Jones M. Aromatase–a brief overview. Annu Rev Physiol. 2002;64:93–127. 10.1146/annurev.physiol.64.081601.142703. [Abstract] [CrossRef] [Google Scholar]
157. Maurelli S, Chiesa M, Giamello E, Di Nardo G, Ferrero VE, Gilardi G, Van Doorslaer S. Direct spectroscopic evidence for binding of anastrozole to the iron heme of human aromatase. Peering into the mechanism of aromatase inhibition. Chem Commun. 2011;47(38):10737–10739. 10.1039/c1cc13872c. [Abstract] [CrossRef] [Google Scholar]
158. Hong Y, Rashid R, Chen S. Binding features of steroidal and nonsteroidal inhibitors. Steroids. 2011;76(8):802–806. 10.1016/j.steroids.2011.02.037. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
159. Geisler J, King N, Anker G, Ornati G, Di Salle E, Lonning PE, Dowsett M. In vivo inhibition of aromatization by exemestane, a novel irreversible aromatase inhibitor, in postmenopausal breast cancer patients. Clin Cancer Res. 1998;4(9):2089–2093. [Abstract] [Google Scholar]
160. Hong Y, Yu B, Sherman M, Yuan YC, Zhou D, Chen S. Molecular basis for the aromatization reaction and exemestane-mediated irreversible inhibition of human aromatase. Mol Endocrinol. 2007;21(2):401–414. 10.1210/me.2006-0281. [Abstract] [CrossRef] [Google Scholar]
161. Geisler J. Differences between the non-steroidal aromatase inhibitors anastrozole and letrozole–of clinical importance? Br J Cancer. 2011;104(7):1059–1066. 10.1038/bjc.2011.58. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
162. Dowsett M, Jones A, Johnston SR, Jacobs S, Trunet P, Smith IE. In vivo measurement of aromatase inhibition by letrozole (CGS 20267) in postmenopausal patients with breast cancer. Clin Cancer Res. 1995;1(12):1511–1515. [Abstract] [Google Scholar]
163. Geisler J, King N, Dowsett M, Ottestad L, Lundgren S, Walton P, Kormeset PO, Lonning PE. Influence of anastrozole (Arimidex), a selective, non-steroidal aromatase inhibitor, on in vivo aromatisation and plasma oestrogen levels in postmenopausal women with breast cancer. Br J Cancer. 1996;74(8):1286–1291. 10.1038/bjc.1996.531. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
164. Bhatnagar AS, Brodie AM, Long BJ, Evans DB, Miller WR. Intracellular aromatase and its relevance to the pharmacological efficacy of aromatase inhibitors. J Steroid Biochem Mol Biol. 2001;76(1–5):199–202. 10.1016/S0960-0760(01)00050-4. [Abstract] [CrossRef] [Google Scholar]
165. De Placido S, Gallo C, De Laurentiis M, Bisagni G, Arpino G, Sarobba MG, Riccardi F, Russo A, Del Mastro L, Cogoni AA, et al. Adjuvant anastrozole versus exemestane versus letrozole, upfront or after 2 years of tamoxifen, in endocrine-sensitive breast cancer (FATA-GIM3): a randomised, phase 3 trial. Lancet Oncol. 2018;19(4):474–485. 10.1016/S1470-2045(18)30116-5. [Abstract] [CrossRef] [Google Scholar]
166. Geisler J, Helle H, Ekse D, Duong NK, Evans DB, Nordbo Y, Aas T, Lonning PE. Letrozole is superior to anastrozole in suppressing breast cancer tissue and plasma estrogen levels. Clin Cancer Res. 2008;14(19):6330–6335. 10.1158/1078-0432.CCR-07-5221. [Abstract] [CrossRef] [Google Scholar]
167. Goss PE, Ingle JN, Pritchard KI, Ellis MJ, Sledge GW, Budd GT, Rabaglio M, Ansari RH, Johnson DB, Tozer R, et al. Exemestane versus anastrozole in postmenopausal women with early breast cancer: NCIC CTG MA.27–a randomized controlled phase III trial. J Clin Oncol. 2013;31(11):1398–1404. 10.1200/JCO.2012.44.7805. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
168. Ellis MJ, Ding L, Shen D, Luo J, Suman VJ, Wallis JW, Van Tine BA, Hoog J, Goiffon RJ, Goldstein TC, et al. Whole-genome analysis informs breast cancer response to aromatase inhibition. Nature. 2012;486(7403):353–360. 10.1038/nature11143. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
169. Buzdar AU, Hortobagyi GN. Tamoxifen and toremifene in breast cancer: comparison of safety and efficacy. J Clin Oncol. 1998;16(1):348–353. 10.1200/JCO.1998.16.1.348. [Abstract] [CrossRef] [Google Scholar]
170. Taneja SS, Smith MR, Dalton JT, Raghow S, Barnette G, Steiner M, Veverka KA. Toremifene–a promising therapy for the prevention of prostate cancer and complications of androgen deprivation therapy. Expert Opin Investig Drugs. 2006;15(3):293–305. 10.1517/13543784.15.3.293. [Abstract] [CrossRef] [Google Scholar]
171. Pink JJ, Jordan VC. Models of estrogen receptor regulation by estrogens and antiestrogens in breast cancer cell lines. Cancer Res. 1996;56(10):2321–2330. [Abstract] [Google Scholar]
172. DeFriend DJ, Howell A, Nicholson RI, Anderson E, Dowsett M, Mansel RE, Blamey RW, Bundred NJ, Robertson JF, Saunders C, et al. Investigation of a new pure antiestrogen (ICI 182780) in women with primary breast cancer. Cancer Res. 1994;54(2):408–414. [Abstract] [Google Scholar]
173. Andreano KJ, Wardell SE, Baker JG, Desautels TK, Baldi R, Chao CA, Heetderks KA, Bae Y, Xiong R, Tonetti DA, et al. G1T48, an oral selective estrogen receptor degrader, and the CDK4/6 inhibitor lerociclib inhibit tumor growth in animal models of endocrine-resistant breast cancer. Breast Cancer Res Treat. 2020;180(3):635–646. 10.1007/s10549-020-05575-9. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
174. Jiang S, Zhang M, Sun J, Yang X. Casein kinase 1alpha: biological mechanisms and theranostic potential. Cell Commun Signal. 2018;16(1):23. 10.1186/s12964-018-0236-z. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
175. Kent LN, Leone G. The broken cycle: E2F dysfunction in cancer. Nat Rev Cancer. 2019;19(6):326–338. 10.1038/s41568-019-0143-7. [Abstract] [CrossRef] [Google Scholar]
176. Turner NC, Ro J, Andre F, Loi S, Verma S, Iwata H, Harbeck N, Loibl S, Huang Bartlett C, Zhang K, et al. Palbociclib in hormone-receptor-positive advanced breast cancer. N Engl J Med. 2015;373(3):209–219. 10.1056/NEJMoa1505270. [Abstract] [CrossRef] [Google Scholar]
177. Finn RS, Martin M, Rugo HS, Jones S, Im SA, Gelmon K, Harbeck N, Lipatov ON, Walshe JM, Moulder S, et al. Palbociclib and letrozole in advanced breast cancer. N Engl J Med. 2016;375(20):1925–1936. 10.1056/NEJMoa1607303. [Abstract] [CrossRef] [Google Scholar]
178. Hortobagyi GN, Stemmer SM, Burris HA, Yap YS, Sonke GS, Paluch-Shimon S, Campone M, Blackwell KL, Andre F, Winer EP, et al. Ribociclib as first-line therapy for hr-positive, advanced breast cancer. N Engl J Med. 2016;375(18):1738–1748. 10.1056/NEJMoa1609709. [Abstract] [CrossRef] [Google Scholar]
179. Goetz MP, Toi M, Campone M, Sohn J, Paluch-Shimon S, Huober J, Park IH, Tredan O, Chen SC, Manso L, et al. MONARCH 3: abemaciclib as initial therapy for advanced breast cancer. J Clin Oncol. 2017;35(32):3638–3646. 10.1200/JCO.2017.75.6155. [Abstract] [CrossRef] [Google Scholar]
180. Sledge GW, Jr, Toi M, Neven P, Sohn J, Inoue K, Pivot X, Burdaeva O, Okera M, Masuda N, Kaufman PA, et al. MONARCH 2: abemaciclib in combination with fulvestrant in women with HR+/HER2- advanced breast cancer who had progressed while receiving endocrine therapy. J Clin Oncol. 2017;35(25):2875–2884. 10.1200/JCO.2017.73.7585. [Abstract] [CrossRef] [Google Scholar]
181. Slamon DJ, Neven P, Chia S, Fasching PA, De Laurentiis M, Im SA, Petrakova K, Bianchi GV, Esteva FJ, Martin M, et al. Phase III randomized study of ribociclib and fulvestrant in hormone receptor-positive, human epidermal growth factor receptor 2-negative advanced breast cancer: MONALEESA-3. J Clin Oncol. 2018;36(24):2465–2472. 10.1200/JCO.2018.78.9909. [Abstract] [CrossRef] [Google Scholar]
182. Braal CL, Jongbloed EM, Wilting SM, Mathijssen RHJ, Koolen SLW, Jager A. Inhibiting CDK4/6 in breast cancer with palbociclib, ribociclib, and abemaciclib: similarities and differences. Drugs. 2021;81(3):317–331. 10.1007/s40265-020-01461-2. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
183. Guiley KZ, Stevenson JW, Lou K, Barkovich KJ, Kumarasamy V, Wijeratne TU, Bunch KL, Tripathi S, Knudsen ES, Witkiewicz AK, et al. p27 allosterically activates cyclin-dependent kinase 4 and antagonizes palbociclib inhibition. Science. 2019;366(6471):eaaw2106. 10.1126/science.aaw2106. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
184. Chen P, Lee NV, Hu W, Xu M, Ferre RA, Lam H, Bergqvist S, Solowiej J, Diehl W, He YA, et al. Spectrum and degree of CDK drug interactions predicts clinical performance. Mol Cancer Ther. 2016;15(10):2273–2281. 10.1158/1535-7163.MCT-16-0300. [Abstract] [CrossRef] [Google Scholar]
185. Tripathy D, Bardia A, Sellers WR. Ribociclib (LEE011): mechanism of action and clinical impact of this selective cyclin-dependent kinase 4/6 inhibitor in various solid tumors. Clin Cancer Res. 2017;23(13):3251–3262. 10.1158/1078-0432.CCR-16-3157. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
186. Harbeck N, Penault-Llorca F, Cortes J, Gnant M, Houssami N, Poortmans P, Ruddy K, Tsang J, Cardoso F. Breast cancer Nat Rev Dis Primers. 2019;5(1):66. 10.1038/s41572-019-0111-2. [Abstract] [CrossRef] [Google Scholar]
187. Piccart-Gebhart MJ, Procter M, Leyland-Jones B, Goldhirsch A, Untch M, Smith I, Gianni L, Baselga J, Bell R, Jackisch C, et al. Trastuzumab after adjuvant chemotherapy in HER2-positive breast cancer. N Engl J Med. 2005;353(16):1659–1672. 10.1056/NEJMoa052306. [Abstract] [CrossRef] [Google Scholar]
188. Cho HS, Mason K, Ramyar KX, Stanley AM, Gabelli SB, Denney DW, Jr, Leahy DJ. Structure of the extracellular region of HER2 alone and in complex with the Herceptin Fab. Nature. 2003;421(6924):756–760. 10.1038/nature01392. [Abstract] [CrossRef] [Google Scholar]
189. Clynes RA, Towers TL, Presta LG, Ravetch JV. Inhibitory Fc receptors modulate in vivo cytotoxicity against tumor targets. Nat Med. 2000;6(4):443–446. 10.1038/74704. [Abstract] [CrossRef] [Google Scholar]
190. Molina MA, Codony-Servat J, Albanell J, Rojo F, Arribas J, Baselga J. Trastuzumab (herceptin), a humanized anti-Her2 receptor monoclonal antibody, inhibits basal and activated Her2 ectodomain cleavage in breast cancer cells. Cancer Res. 2001;61(12):4744–4749. [Abstract] [Google Scholar]
191. Junttila TT, Akita RW, Parsons K, Fields C, Lewis Phillips GD, Friedman LS, Sampath D, Sliwkowski MX. Ligand-independent HER2/HER3/PI3K complex is disrupted by trastuzumab and is effectively inhibited by the PI3K inhibitor GDC-0941. Cancer Cell. 2009;15(5):429–440. 10.1016/j.ccr.2009.03.020. [Abstract] [CrossRef] [Google Scholar]
192. Agus DB, Gordon MS, Taylor C, Natale RB, Karlan B, Mendelson DS, Press MF, Allison DE, Sliwkowski MX, Lieberman G, et al. Phase I clinical study of pertuzumab, a novel HER dimerization inhibitor, in patients with advanced cancer. J Clin Oncol. 2005;23(11):2534–2543. 10.1200/JCO.2005.03.184. [Abstract] [CrossRef] [Google Scholar]
193. Nami B, Maadi H, Wang Z. Mechanisms underlying the action and synergism of trastuzumab and pertuzumab in targeting HER2-positive breast cancer. Cancers. 2018;10(10):342. 10.3390/cancers10100342. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
194. Baselga J, Cortes J, Kim SB, Im SA, Hegg R, Im YH, Roman L, Pedrini JL, Pienkowski T, Knott A, et al. Pertuzumab plus trastuzumab plus docetaxel for metastatic breast cancer. N Engl J Med. 2012;366(2):109–119. 10.1056/NEJMoa1113216. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
195. Swain SM, Baselga J, Kim SB, Ro J, Semiglazov V, Campone M, Ciruelos E, Ferrero JM, Schneeweiss A, Heeson S, et al. Pertuzumab, trastuzumab, and docetaxel in HER2-positive metastatic breast cancer. N Engl J Med. 2015;372(8):724–734. 10.1056/NEJMoa1413513. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
196. Bang YJ, Giaccone G, Im SA, Oh DY, Bauer TM, Nordstrom JL, Li H, Chichili GR, Moore PA, Hong S, et al. First-in-human phase 1 study of margetuximab (MGAH22), an Fc-modified chimeric monoclonal antibody, in patients with HER2-positive advanced solid tumors. Ann Oncol. 2017;28(4):855–861. 10.1093/annonc/mdx002. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
197. Rugo HS, Im SA, Cardoso F, Cortes J, Curigliano G, Musolino A, Pegram MD, Wright GS, Saura C, Escriva-de-Romani S, et al. Efficacy of margetuximab vs trastuzumab in patients with pretreated ERBB2-positive advanced breast cancer: a phase 3 randomized clinical trial. JAMA Oncol. 2021;7(4):573–584. 10.1001/jamaoncol.2020.7932. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
198. Slamon DJ, Leyland-Jones B, Shak S, Fuchs H, Paton V, Bajamonde A, Fleming T, Eiermann W, Wolter J, Pegram M, et al. Use of chemotherapy plus a monoclonal antibody against HER2 for metastatic breast cancer that overexpresses HER2. N Engl J Med. 2001;344(11):783–792. 10.1056/NEJM200103153441101. [Abstract] [CrossRef] [Google Scholar]
199. Romond EH, Perez EA, Bryant J, Suman VJ, Geyer CE, Jr, Davidson NE, Tan-Chiu E, Martino S, Paik S, Kaufman PA, et al. Trastuzumab plus adjuvant chemotherapy for operable HER2-positive breast cancer. N Engl J Med. 2005;353(16):1673–1684. 10.1056/NEJMoa052122. [Abstract] [CrossRef] [Google Scholar]
200. Lavaud P, Andre F. Strategies to overcome trastuzumab resistance in HER2-overexpressing breast cancers: focus on new data from clinical trials. BMC Med. 2014;12:132. 10.1186/s12916-014-0132-3. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
201. Pohlmann PR, Mayer IA, Mernaugh R. Resistance to trastuzumab in breast cancer. Clin Cancer Res. 2009;15(24):7479–7491. 10.1158/1078-0432.CCR-09-0636. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
202. Lewis Phillips GD, Li G, Dugger DL, Crocker LM, Parsons KL, Mai E, Blattler WA, Lambert JM, Chari RV, Lutz RJ, et al. Targeting HER2-positive breast cancer with trastuzumab-DM1, an antibody-cytotoxic drug conjugate. Cancer Res. 2008;68(22):9280–9290. 10.1158/0008-5472.CAN-08-1776. [Abstract] [CrossRef] [Google Scholar]
203. Junttila TT, Li G, Parsons K, Phillips GL, Sliwkowski MX. Trastuzumab-DM1 (T-DM1) retains all the mechanisms of action of trastuzumab and efficiently inhibits growth of lapatinib insensitive breast cancer. Breast Cancer Res Treat. 2011;128(2):347–356. 10.1007/s10549-010-1090-x. [Abstract] [CrossRef] [Google Scholar]
204. Kupchan SM, Komoda Y, Court WA, Thomas GJ, Smith RM, Karim A, Gilmore CJ, Haltiwanger RC, Bryan RF. Maytansine, a novel antileukemic ansa macrolide from Maytenus ovatus. J Am Chem Soc. 1972;94(4):1354–1356. 10.1021/ja00759a054. [Abstract] [CrossRef] [Google Scholar]
205. Remillard S, Rebhun LI, Howie GA, Kupchan SM. Antimitotic activity of the potent tumor inhibitor maytansine. Science. 1975;189(4207):1002–1005. 10.1126/science.1241159. [Abstract] [CrossRef] [Google Scholar]
206. Verma S, Miles D, Gianni L, Krop IE, Welslau M, Baselga J, Pegram M, Oh DY, Dieras V, Guardino E, et al. Trastuzumab emtansine for HER2-positive advanced breast cancer. N Engl J Med. 2012;367(19):1783–1791. 10.1056/NEJMoa1209124. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
207. Hurvitz SA, Dirix L, Kocsis J, Bianchi GV, Lu J, Vinholes J, Guardino E, Song C, Tong B, Ng V, et al. Phase II randomized study of trastuzumab emtansine versus trastuzumab plus docetaxel in patients with human epidermal growth factor receptor 2-positive metastatic breast cancer. J Clin Oncol. 2013;31(9):1157–1163. 10.1200/JCO.2012.44.9694. [Abstract] [CrossRef] [Google Scholar]
208. Dieras V, Miles D, Verma S, Pegram M, Welslau M, Baselga J, Krop IE, Blackwell K, Hoersch S, Xu J, et al. Trastuzumab emtansine versus capecitabine plus lapatinib in patients with previously treated HER2-positive advanced breast cancer (EMILIA): a descriptive analysis of final overall survival results from a randomised, open-label, phase 3 trial. Lancet Oncol. 2017;18(6):732–742. 10.1016/S1470-2045(17)30312-1. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
209. Krop IE, Kim SB, Martin AG, LoRusso PM, Ferrero JM, Badovinac-Crnjevic T, Hoersch S, Smitt M, Wildiers H. Trastuzumab emtansine versus treatment of physician's choice in patients with previously treated HER2-positive metastatic breast cancer (TH3RESA): final overall survival results from a randomised open-label phase 3 trial. Lancet Oncol. 2017;18(6):743–754. 10.1016/S1470-2045(17)30313-3. [Abstract] [CrossRef] [Google Scholar]
210. Kovtun YV, Audette CA, Mayo MF, Jones GE, Doherty H, Maloney EK, Erickson HK, Sun X, Wilhelm S, Ab O, et al. Antibody-maytansinoid conjugates designed to bypass multidrug resistance. Cancer Res. 2010;70(6):2528–2537. 10.1158/0008-5472.CAN-09-3546. [Abstract] [CrossRef] [Google Scholar]
211. Garcia-Alonso S, Ocana A, Pandiella A. Trastuzumab emtansine: mechanisms of action and resistance, clinical progress, and beyond. Trends Cancer. 2020;6(2):130–146. 10.1016/j.trecan.2019.12.010. [Abstract] [CrossRef] [Google Scholar]
212. Ogitani Y, Aida T, Hagihara K, Yamaguchi J, Ishii C, Harada N, Soma M, Okamoto H, Oitate M, Arakawa S, et al. DS-8201a, a novel HER2-targeting ADC with a novel DNA topoisomerase I inhibitor, demonstrates a promising antitumor efficacy with differentiation from T-DM1. Clin Cancer Res. 2016;22(20):5097–5108. 10.1158/1078-0432.CCR-15-2822. [Abstract] [CrossRef] [Google Scholar]
213. Modi S, Saura C, Yamashita T, Park YH, Kim SB, Tamura K, Andre F, Iwata H, Ito Y, Tsurutani J, et al. Trastuzumab deruxtecan in previously treated HER2-positive breast cancer. N Engl J Med. 2020;382(7):610–621. 10.1056/NEJMoa1914510. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
214. Cortes J, Kim SB, Chung WP, Im SA, Park YH, Hegg R, Kim MH, Tseng LM, Petry V, Chung CF, et al. Trastuzumab deruxtecan versus trastuzumab emtansine for breast cancer. N Engl J Med. 2022;386(12):1143–1154. 10.1056/NEJMoa2115022. [Abstract] [CrossRef] [Google Scholar]
215. Duchnowska R, Loibl S, Jassem J. Tyrosine kinase inhibitors for brain metastases in HER2-positive breast cancer. Cancer Treat Rev. 2018;67:71–77. 10.1016/j.ctrv.2018.05.004. [Abstract] [CrossRef] [Google Scholar]
216. Rusnak DW, Affleck K, Cockerill SG, Stubberfield C, Harris R, Page M, Smith KJ, Guntrip SB, Carter MC, Shaw RJ, et al. The characterization of novel, dual ErbB-2/EGFR, tyrosine kinase inhibitors: potential therapy for cancer. Cancer Res. 2001;61(19):7196–7203. [Abstract] [Google Scholar]
217. Rusnak DW, Lackey K, Affleck K, Wood ER, Alligood KJ, Rhodes N, Keith BR, Murray DM, Knight WB, Mullin RJ, et al. The effects of the novel, reversible epidermal growth factor receptor/ErbB-2 tyrosine kinase inhibitor, GW2016, on the growth of human normal and tumor-derived cell lines in vitro and in vivo. Mol Cancer Ther. 2001;1(2):85–94. [Abstract] [Google Scholar]
218. Tsang RY, Sadeghi S, Finn RS. Lapatinib, a dual-targeted small molecule inhibitor of EGFR and HER2, in HER2-amplified breast cancer: from bench to bedside. Clin Med Insights Ther. 2011 10.4137/CMT.S3783. [CrossRef] [Google Scholar]
219. Dai CL, Tiwari AK, Wu CP, Su XD, Wang SR, Liu DG, Ashby CR, Jr, Huang Y, Robey RW, Liang YJ, et al. Lapatinib (Tykerb, GW572016) reverses multidrug resistance in cancer cells by inhibiting the activity of ATP-binding cassette subfamily B member 1 and G member 2. Cancer Res. 2008;68(19):7905–7914. 10.1158/0008-5472.CAN-08-0499. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
220. Trowe T, Boukouvala S, Calkins K, Cutler RE, Jr, Fong R, Funke R, Gendreau SB, Kim YD, Miller N, Woolfrey JR, et al. EXEL-7647 inhibits mutant forms of ErbB2 associated with lapatinib resistance and neoplastic transformation. Clin Cancer Res. 2008;14(8):2465–2475. 10.1158/1078-0432.CCR-07-4367. [Abstract] [CrossRef] [Google Scholar]
221. Rexer BN, Ghosh R, Narasanna A, Estrada MV, Chakrabarty A, Song Y, Engelman JA, Arteaga CL. Human breast cancer cells harboring a gatekeeper T798M mutation in HER2 overexpress EGFR ligands and are sensitive to dual inhibition of EGFR and HER2. Clin Cancer Res. 2013;19(19):5390–5401. 10.1158/1078-0432.CCR-13-1038. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
222. D'Amato V, Raimondo L, Formisano L, Giuliano M, De Placido S, Rosa R, Bianco R. Mechanisms of lapatinib resistance in HER2-driven breast cancer. Cancer Treat Rev. 2015;41(10):877–883. 10.1016/j.ctrv.2015.08.001. [Abstract] [CrossRef] [Google Scholar]
223. Rabindran SK, Discafani CM, Rosfjord EC, Baxter M, Floyd MB, Golas J, Hallett WA, Johnson BD, Nilakantan R, Overbeek E, et al. Antitumor activity of HKI-272, an orally active, irreversible inhibitor of the HER-2 tyrosine kinase. Cancer Res. 2004;64(11):3958–3965. 10.1158/0008-5472.CAN-03-2868. [Abstract] [CrossRef] [Google Scholar]
224. Paranjpe R, Basatneh D, Tao G, De Angelis C, Noormohammed S, Ekinci E, Abughosh S, Ghose R, Trivedi MV. Neratinib in HER2-positive breast cancer patients. Ann Pharmacother. 2019;53(6):612–620. 10.1177/1060028018824088. [Abstract] [CrossRef] [Google Scholar]
225. Burstein HJ, Sun Y, Dirix LY, Jiang Z, Paridaens R, Tan AR, Awada A, Ranade A, Jiao S, Schwartz G, et al. Neratinib, an irreversible ErbB receptor tyrosine kinase inhibitor, in patients with advanced ErbB2-positive breast cancer. J Clin Oncol. 2010;28(8):1301–1307. 10.1200/JCO.2009.25.8707. [Abstract] [CrossRef] [Google Scholar]
226. Park JW, Liu MC, Yee D, Yau C, van’t-Veer LJ, Symmans WF, Paoloni M, Perlmutter J, Hylton NM, Hogarth M, et al. Adaptive randomization of neratinib in early breast cancer. N Engl J Med. 2016;375(1):11–22. 10.1056/NEJMoa1513750. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
227. Hanker AB, Brewer MR, Sheehan JH, Koch JP, Sliwoski GR, Nagy R, Lanman R, Berger MF, Hyman DM, Solit DB, et al. An acquired HER2(T798I) gatekeeper mutation induces resistance to neratinib in a patient with HER2 mutant-driven breast cancer. Cancer Discov. 2017;7(6):575–585. 10.1158/2159-8290.CD-16-1431. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
228. Conlon NT, Kooijman JJ, van Gerwen SJC, Mulder WR, Zaman GJR, Diala I, Eli LD, Lalani AS, Crown J, Collins DM. Comparative analysis of drug response and gene profiling of HER2-targeted tyrosine kinase inhibitors. Br J Cancer. 2021;124(7):1249–1259. 10.1038/s41416-020-01257-x. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
229. Schlam I, Swain SM. HER2-positive breast cancer and tyrosine kinase inhibitors: the time is now. NPJ Breast Cancer. 2021;7(1):56. 10.1038/s41523-021-00265-1. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
230. Murthy R, Borges VF, Conlin A, Chaves J, Chamberlain M, Gray T, Vo A, Hamilton E. Tucatinib with capecitabine and trastuzumab in advanced HER2-positive metastatic breast cancer with and without brain metastases: a non-randomised, open-label, phase 1b study. Lancet Oncol. 2018;19(7):880–888. 10.1016/S1470-2045(18)30256-0. [Abstract] [CrossRef] [Google Scholar]
231. Murthy RK, Loi S, Okines A, Paplomata E, Hamilton E, Hurvitz SA, Lin NU, Borges V, Abramson V, Anders C, et al. Tucatinib, trastuzumab, and capecitabine for HER2-positive metastatic breast cancer. N Engl J Med. 2020;382(7):597–609. 10.1056/NEJMoa1914609. [Abstract] [CrossRef] [Google Scholar]
232. Lin NU, Borges V, Anders C, Murthy RK, Paplomata E, Hamilton E, Hurvitz S, Loi S, Okines A, Abramson V, et al. Intracranial efficacy and survival with tucatinib plus trastuzumab and capecitabine for previously treated HER2-positive breast cancer with brain metastases in the HER2CLIMB trial. J Clin Oncol. 2020;38(23):2610–2619. 10.1200/JCO.20.00775. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
233. Cancer Genome Atlas N Comprehensive molecular portraits of human breast tumours. Nature. 2012;490(7418):61–70. 10.1038/nature11412. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
234. Jonsson P, Bandlamudi C, Cheng ML, Srinivasan P, Chavan SS, Friedman ND, Rosen EY, Richards AL, Bouvier N, Selcuklu SD, et al. Tumour lineage shapes BRCA-mediated phenotypes. Nature. 2019;571(7766):576–579. 10.1038/s41586-019-1382-1. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
235. King MC, Marks JH, Mandell JB. New York breast cancer study G: breast and ovarian cancer risks due to inherited mutations in BRCA1 and BRCA2. Science. 2003;302(5645):643–646. 10.1126/science.1088759. [Abstract] [CrossRef] [Google Scholar]
236. Wooster R, Weber BL. Breast and ovarian cancer. N Engl J Med. 2003;348(23):2339–2347. 10.1056/NEJMra012284. [Abstract] [CrossRef] [Google Scholar]
237. Zhao W, Wiese C, Kwon Y, Hromas R, Sung P. The BRCA tumor suppressor network in chromosome damage repair by homologous recombination. Annu Rev Biochem. 2019;88:221–245. 10.1146/annurev-biochem-013118-111058. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
238. Hegde ML, Hazra TK, Mitra S. Early steps in the DNA base excision/single-strand interruption repair pathway in mammalian cells. Cell Res. 2008;18(1):27–47. 10.1038/cr.2008.8. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
239. Murai J, Huang SY, Renaud A, Zhang Y, Ji J, Takeda S, Morris J, Teicher B, Doroshow JH, Pommier Y. Stereospecific PARP trapping by BMN 673 and comparison with olaparib and rucaparib. Mol Cancer Ther. 2014;13(2):433–443. 10.1158/1535-7163.MCT-13-0803. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
240. Murai J, Huang SY, Das BB, Renaud A, Zhang Y, Doroshow JH, Ji J, Takeda S, Pommier Y. Trapping of PARP1 and PARP2 by Clinical PARP Inhibitors. Cancer Res. 2012;72(21):5588–5599. 10.1158/0008-5472.CAN-12-2753. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
241. Hughes DL. Patent review of manufacturing routes to recently approved parp inhibitors: olaparib, rucaparib, and niraparib. Org Process Res Dev. 2017;21(9):1227–1244. 10.1021/acs.oprd.7b00235. [CrossRef] [Google Scholar]
242. Min A, Im SA. PARP inhibitors as therapeutics: beyond modulation of PARylation. Cancers. 2020;12(2):394. 10.3390/cancers12020394. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
243. Calvert H, Azzariti A. The clinical development of inhibitors of poly(ADP-ribose) polymerase. Ann Oncol. 2011;22:i53–i59. 10.1093/annonc/mdq667. [Abstract] [CrossRef] [Google Scholar]
244. Pilie PG, Tang C, Mills GB, Yap TA. State-of-the-art strategies for targeting the DNA damage response in cancer. Nat Rev Clin Oncol. 2019;16(2):81–104. 10.1038/s41571-018-0114-z. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
245. Lord CJ, Ashworth A. PARP inhibitors: synthetic lethality in the clinic. Science. 2017;355(6330):1152–1158. 10.1126/science.aam7344. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
246. Farmer H, McCabe N, Lord CJ, Tutt AN, Johnson DA, Richardson TB, Santarosa M, Dillon KJ, Hickson I, Knights C, et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature. 2005;434(7035):917–921. 10.1038/nature03445. [Abstract] [CrossRef] [Google Scholar]
247. Mateo J, Lord CJ, Serra V, Tutt A, Balmana J, Castroviejo-Bermejo M, Cruz C, Oaknin A, Kaye SB, de Bono JS. A decade of clinical development of PARP inhibitors in perspective. Ann Oncol. 2019;30(9):1437–1447. 10.1093/annonc/mdz192. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
248. Aoyagi-Scharber M, Gardberg AS, Yip BK, Wang B, Shen Y, Fitzpatrick PA. Structural basis for the inhibition of poly(ADP-ribose) polymerases 1 and 2 by BMN 673, a potent inhibitor derived from dihydropyridophthalazinone. Acta Crystallogr F Struct Biol Commun. 2014;70(Pt 9):1143–1149. 10.1107/S2053230X14015088. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
249. Wang B, Chu D, Feng Y, Shen Y, Aoyagi-Scharber M, Post LE. Discovery and characterization of (8S,9R)-5-fluoro-8-(4-fluorophenyl)-9-(1-methyl-1H-1,2,4-triazol-5-yl)-2,7,8,9-te trahydro-3H-pyrido[4,3,2-de]phthalazin-3-one (BMN 673, talazoparib), a novel, highly potent, and orally efficacious Poly(ADP-ribose) polymerase-1/2 inhibitor, as an anticancer agent. J Med Chem. 2016;59(1):335–357. 10.1021/acs.jmedchem.5b01498. [Abstract] [CrossRef] [Google Scholar]
250. Jones P, Altamura S, Boueres J, Ferrigno F, Fonsi M, Giomini C, Lamartina S, Monteagudo E, Ontoria JM, Orsale MV, et al. Discovery of 2-{4-[(3S)-piperidin-3-yl]phenyl}-2H-indazole-7-carboxamide (MK-4827): a novel oral poly(ADP-ribose)polymerase (PARP) inhibitor efficacious in BRCA-1 and -2 mutant tumors. J Med Chem. 2009;52(22):7170–7185. 10.1021/jm901188v. [Abstract] [CrossRef] [Google Scholar]
251. Shen Y, Rehman FL, Feng Y, Boshuizen J, Bajrami I, Elliott R, Wang B, Lord CJ, Post LE, Ashworth A. BMN 673, a novel and highly potent PARP1/2 inhibitor for the treatment of human cancers with DNA repair deficiency. Clin Cancer Res. 2013;19(18):5003–5015. 10.1158/1078-0432.CCR-13-1391. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
252. de Bono J, Ramanathan RK, Mina L, Chugh R, Glaspy J, Rafii S, Kaye S, Sachdev J, Heymach J, Smith DC, et al. Phase I, dose-escalation, two-part trial of the PARP inhibitor talazoparib in patients with advanced germline BRCA1/2 mutations and selected sporadic cancers. Cancer Discov. 2017;7(6):620–629. 10.1158/2159-8290.CD-16-1250. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
253. Litton JK, Rugo HS, Ettl J, Hurvitz SA, Goncalves A, Lee KH, Fehrenbacher L, Yerushalmi R, Mina LA, Martin M, et al. Talazoparib in patients with advanced breast cancer and a germline BRCA mutation. N Engl J Med. 2018;379(8):753–763. 10.1056/NEJMoa1802905. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
254. Goncalves MD, Hopkins BD, Cantley LC. Phosphatidylinositol 3-Kinase, growth disorders, and cancer. N Engl J Med. 2018;379(21):2052–2062. 10.1056/NEJMra1704560. [Abstract] [CrossRef] [Google Scholar]
255. Andre F, Ciruelos E, Rubovszky G, Campone M, Loibl S, Rugo HS, Iwata H, Conte P, Mayer IA, Kaufman B, et al. Alpelisib for PIK3CA-mutated, hormone receptor-positive advanced breast cancer. N Engl J Med. 2019;380(20):1929–1940. 10.1056/NEJMoa1813904. [Abstract] [CrossRef] [Google Scholar]
256. Furet P, Guagnano V, Fairhurst RA, Imbach-Weese P, Bruce I, Knapp M, Fritsch C, Blasco F, Blanz J, Aichholz R, et al. Discovery of NVP-BYL719 a potent and selective phosphatidylinositol-3 kinase alpha inhibitor selected for clinical evaluation. Bioorg Med Chem Lett. 2013;23(13):3741–3748. 10.1016/j.bmcl.2013.05.007. [Abstract] [CrossRef] [Google Scholar]
257. Vanhaesebroeck B, Perry MWD, Brown JR, Andre F, Okkenhaug K. PI3K inhibitors are finally coming of age. Nat Rev Drug Discov. 2021 10.1038/s41573-021-00209-1. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
258. Juric D, Rodon J, Tabernero J, Janku F, Burris HA, Schellens JHM, Middleton MR, Berlin J, Schuler M, Gil-Martin M, et al. Phosphatidylinositol 3-kinase alpha-selective inhibition with alpelisib (BYL719) in PIK3CA-altered solid tumors: results from the first-in-human study. J Clin Oncol. 2018;36(13):1291–1299. 10.1200/JCO.2017.72.7107. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
259. Juric D, Janku F, Rodon J, Burris HA, Mayer IA, Schuler M, Seggewiss-Bernhardt R, Gil-Martin M, Middleton MR, Baselga J, et al. Alpelisib plus fulvestrant in PIK3CA-altered and PIK3CA-wild-type estrogen receptor-positive advanced breast cancer: a phase 1b clinical trial. JAMA Oncol. 2019;5(2):e184475. 10.1001/jamaoncol.2018.4475. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
260. Rugo HS, Andre F, Yamashita T, Cerda H, Toledano I, Stemmer SM, Jurado JC, Juric D, Mayer I, Ciruelos EM, et al. Time course and management of key adverse events during the randomized phase III SOLAR-1 study of PI3K inhibitor alpelisib plus fulvestrant in patients with HR-positive advanced breast cancer. Ann Oncol. 2020;31(8):1001–1010. 10.1016/j.annonc.2020.05.001. [Abstract] [CrossRef] [Google Scholar]
261. Mayer IA, Abramson VG, Formisano L, Balko JM, Estrada MV, Sanders ME, Juric D, Solit D, Berger MF, Won HH, et al. A phase Ib study of alpelisib (BYL719), a PI3Kalpha-specific inhibitor, with letrozole in ER+/HER2- metastatic breast cancer. Clin Cancer Res. 2017;23(1):26–34. 10.1158/1078-0432.CCR-16-0134. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
262. Foulkes WD, Smith IE, Reis-Filho JS. Triple-negative breast cancer. N Engl J Med. 2010;363(20):1938–1948. 10.1056/NEJMra1001389. [Abstract] [CrossRef] [Google Scholar]
263. Cardoso F, Senkus E, Costa A, Papadopoulos E, Aapro M, Andre F, Harbeck N, Aguilar Lopez B, Barrios CH, Bergh J, et al. 4th ESO-ESMO international consensus guidelines for advanced breast cancer (ABC 4)dagger. Ann Oncol. 2018;29(8):1634–1657. 10.1093/annonc/mdy192. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
264. Yardley DA, Coleman R, Conte P, Cortes J, Brufsky A, Shtivelband M, Young R, Bengala C, Ali H, Eakel J, et al. nab-Paclitaxel plus carboplatin or gemcitabine versus gemcitabine plus carboplatin as first-line treatment of patients with triple-negative metastatic breast cancer: results from the tnAcity trial. Ann Oncol. 2018;29(8):1763–1770. 10.1093/annonc/mdy201. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
265. Schmid P, Adams S, Rugo HS, Schneeweiss A, Barrios CH, Iwata H, Dieras V, Hegg R, Im SA, Shaw Wright G, et al. Atezolizumab and nab-paclitaxel in advanced triple-negative breast cancer. N Engl J Med. 2018;379(22):2108–2121. 10.1056/NEJMoa1809615. [Abstract] [CrossRef] [Google Scholar]
266. Narayan P, Wahby S, Gao JJ, Amiri-Kordestani L, Ibrahim A, Bloomquist E, Tang S, Xu Y, Liu J, Fu W, et al. FDA approval summary: atezolizumab plus paclitaxel protein-bound for the treatment of patients with advanced or metastatic TNBC whose tumors express PD-L1. Clin Cancer Res. 2020;26(10):2284–2289. 10.1158/1078-0432.CCR-19-3545. [Abstract] [CrossRef] [Google Scholar]
267. Trerotola M, Cantanelli P, Guerra E, Tripaldi R, Aloisi AL, Bonasera V, Lattanzio R, de Lange R, Weidle UH, Piantelli M, et al. Upregulation of Trop-2 quantitatively stimulates human cancer growth. Oncogene. 2013;32(2):222–233. 10.1038/onc.2012.36. [Abstract] [CrossRef] [Google Scholar]
268. Coates JT, Sun S, Leshchiner I, Thimmiah N, Martin EE, McLoughlin D, Danysh BP, Slowik K, Jacobs RA, Rhrissorrakrai K, et al. Parallel genomic alterations of antigen and payload targets mediate polyclonal acquired clinical resistance to sacituzumab govitecan in triple-negative breast cancer. Cancer Discov. 2021;11(10):2436–2445. 10.1158/2159-8290.CD-21-0702. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
269. Cardillo TM, Govindan SV, Sharkey RM, Trisal P, Goldenberg DM. Humanized anti-Trop-2 IgG-SN-38 conjugate for effective treatment of diverse epithelial cancers: preclinical studies in human cancer xenograft models and monkeys. Clin Cancer Res. 2011;17(10):3157–3169. 10.1158/1078-0432.CCR-10-2939. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
270. Cardillo TM, Govindan SV, Sharkey RM, Trisal P, Arrojo R, Liu D, Rossi EA, Chang CH, Goldenberg DM. Sacituzumab govitecan (IMMU-132), an anti-trop-2/SN-38 antibody-drug conjugate: characterization and efficacy in pancreatic, gastric, and other cancers. Bioconjug Chem. 2015;26(5):919–931. 10.1021/acs.bioconjchem.5b00223. [Abstract] [CrossRef] [Google Scholar]
271. Stein R, Basu A, Chen S, Shih LB, Goldenberg DM. Specificity and properties of MAb RS7-3G11 and the antigen defined by this pancarcinoma monoclonal antibody. Int J Cancer. 1993;55(6):938–946. 10.1002/ijc.2910550611. [Abstract] [CrossRef] [Google Scholar]
272. Varughese J, Cocco E, Bellone S, de Leon M, Bellone M, Todeschini P, Schwartz PE, Rutherford TJ, Pecorelli S, Santin AD. Uterine serous papillary carcinomas overexpress human trophoblast-cell-surface marker (Trop-2) and are highly sensitive to immunotherapy with hRS7, a humanized anti-Trop-2 monoclonal antibody. Cancer. 2011;117(14):3163–3172. 10.1002/cncr.25891. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
273. Rivory LP, Bowles MR, Robert J, Pond SM. Conversion of irinotecan (CPT-11) to its active metabolite, 7-ethyl-10-hydroxycamptothecin (SN-38), by human liver carboxylesterase. Biochem Pharmacol. 1996;52(7):1103–1111. 10.1016/0006-2952(96)00457-1. [Abstract] [CrossRef] [Google Scholar]
274. Starodub AN, Ocean AJ, Shah MA, Guarino MJ, Picozzi VJ, Jr, Vahdat LT, Thomas SS, Govindan SV, Maliakal PP, Wegener WA, et al. First-in-human trial of a novel anti-trop-2 antibody-SN-38 conjugate, sacituzumab govitecan, for the treatment of diverse metastatic solid tumors. Clin Cancer Res. 2015;21(17):3870–3878. 10.1158/1078-0432.CCR-14-3321. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
275. Bardia A, Mayer IA, Diamond JR, Moroose RL, Isakoff SJ, Starodub AN, Shah NC, O'Shaughnessy J, Kalinsky K, Guarino M, et al. Efficacy and safety of anti-trop-2 antibody drug conjugate sacituzumab govitecan (IMMU-132) in heavily pretreated patients with metastatic triple-negative breast cancer. J Clin Oncol. 2017;35(19):2141–2148. 10.1200/JCO.2016.70.8297. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
276. Bardia A, Mayer IA, Vahdat LT, Tolaney SM, Isakoff SJ, Diamond JR, O'Shaughnessy J, Moroose RL, Santin AD, Abramson VG, et al. Sacituzumab govitecan-hziy in refractory metastatic triple-negative breast cancer. N Engl J Med. 2019;380(8):741–751. 10.1056/NEJMoa1814213. [Abstract] [CrossRef] [Google Scholar]
277. Bardia A, Hurvitz SA, Tolaney SM, Loirat D, Punie K, Oliveira M, Brufsky A, Sardesai SD, Kalinsky K, Zelnak AB, et al. Sacituzumab govitecan in metastatic triple-negative breast cancer. N Engl J Med. 2021;384(16):1529–1541. 10.1056/NEJMoa2028485. [Abstract] [CrossRef] [Google Scholar]
278. Bardia A, Tolaney SM, Punie K, Loirat D, Oliveira M, Kalinsky K, Zelnak A, Aftimos P, Dalenc F, Sardesai S, et al. Biomarker analyses in the phase III ASCENT study of sacituzumab govitecan versus chemotherapy in patients with metastatic triple-negative breast cancer. Ann Oncol. 2021;32(9):1148–1156. 10.1016/j.annonc.2021.06.002. [Abstract] [CrossRef] [Google Scholar]
279. Wani MC, Taylor HL, Wall ME, Coggon P, McPhail AT. Plant antitumor agents. VI. The isolation and structure of taxol, a novel antileukemic and antitumor agent from Taxus brevifolia. J Am Chem Soc. 1971;93(9):2325–2327. 10.1021/ja00738a045. [Abstract] [CrossRef] [Google Scholar]
280. Crown J, O'Leary M. The taxanes: an update. Lancet. 2000;355(9210):1176–1178. 10.1016/S0140-6736(00)02074-2. [Abstract] [CrossRef] [Google Scholar]
281. Rao S, Krauss NE, Heerding JM, Swindell CS, Ringel I, Orr GA, Horwitz SB. 3′-(p-azidobenzamido)taxol photolabels the N-terminal 31 amino acids of beta-tubulin. J Biol Chem. 1994;269(5):3132–3134. 10.1016/S0021-9258(17)41836-9. [Abstract] [CrossRef] [Google Scholar]
282. Rao S, Orr GA, Chaudhary AG, Kingston DG, Horwitz SB. Characterization of the taxol binding site on the microtubule. 2-(m-Azidobenzoyl)taxol photolabels a peptide (amino acids 217–231) of beta-tubulin. J Biol Chem. 1995;270(35):20235–20238. 10.1074/jbc.270.35.20235. [Abstract] [CrossRef] [Google Scholar]
283. Rao S, He L, Chakravarty S, Ojima I, Orr GA, Horwitz SB. Characterization of the taxol binding site on the microtubule. Identification of Arg(282) in beta-tubulin as the site of photoincorporation of a 7-benzophenone analogue of Taxol. J Biol Chem. 1999;274(53):37990–37994. 10.1074/jbc.274.53.37990. [Abstract] [CrossRef] [Google Scholar]
284. Prota AE, Bargsten K, Zurwerra D, Field JJ, Diaz JF, Altmann KH, Steinmetz MO. Molecular mechanism of action of microtubule-stabilizing anticancer agents. Science. 2013;339(6119):587–590. 10.1126/science.1230582. [Abstract] [CrossRef] [Google Scholar]
285. Alsop K, Fereday S, Meldrum C, deFazio A, Emmanuel C, George J, Dobrovic A, Birrer MJ, Webb PM, Stewart C, et al. BRCA mutation frequency and patterns of treatment response in BRCA mutation-positive women with ovarian cancer: a report from the Australian ovarian cancer study group. J Clin Oncol. 2012;30(21):2654–2663. 10.1200/JCO.2011.39.8545. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
286. Menear KA, Adcock C, Boulter R, Cockcroft X-l, Copsey L, Cranston A, Dillon KJ, Drzewiecki J, Garman S, Gomez S, et al. 4-[3-(4-Cyclopropanecarbonylpiperazine-1-carbonyl)-4-fluorobenzyl]-2H-phthalazin-1-one: a novel bioavailable inhibitor of poly(ADP-ribose) polymerase-1. J Med Chem. 2008;51(20):6581–6591. 10.1021/jm8001263. [Abstract] [CrossRef] [Google Scholar]
287. Thomas HD, Calabrese CR, Batey MA, Canan S, Hostomsky Z, Kyle S, Maegley KA, Newell DR, Skalitzky D, Wang LZ, et al. Preclinical selection of a novel poly(ADP-ribose) polymerase inhibitor for clinical trial. Mol Cancer Ther. 2007;6(3):945–956. 10.1158/1535-7163.MCT-06-0552. [Abstract] [CrossRef] [Google Scholar]
288. Deeks ED. Olaparib: first global approval. Drugs. 2015;75(2):231–240. 10.1007/s40265-015-0345-6. [Abstract] [CrossRef] [Google Scholar]
289. Syed YY. Rucaparib: first global approval. Drugs. 2017;77(5):585–592. 10.1007/s40265-017-0716-2. [Abstract] [CrossRef] [Google Scholar]
290. Scott LJ. Niraparib: first global approval. Drugs. 2017;77(9):1029–1034. 10.1007/s40265-017-0752-y. [Abstract] [CrossRef] [Google Scholar]
291. Fong PC, Yap TA, Boss DS, Carden CP, Mergui-Roelvink M, Gourley C, De Greve J, Lubinski J, Shanley S, Messiou C, et al. Poly(ADP)-ribose polymerase inhibition: frequent durable responses in BRCA carrier ovarian cancer correlating with platinum-free interval. J Clin Oncol. 2010;28(15):2512–2519. 10.1200/JCO.2009.26.9589. [Abstract] [CrossRef] [Google Scholar]
292. Ceccaldi R, O'Connor KW, Mouw KW, Li AY, Matulonis UA, D'Andrea AD, Konstantinopoulos PA. A unique subset of epithelial ovarian cancers with platinum sensitivity and PARP inhibitor resistance. Cancer Res. 2015;75(4):628–634. 10.1158/0008-5472.CAN-14-2593. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
293. Sackeyfio A, Nussey F, Friedlander M, Pujade-Lauraine E. Comparative efficacy and tolerability of the PARP inhibitors, olaparib 300 mg tablets BID, niraparib 300 mg capsules QD and rucaparib 600 mg tablets BID as maintenance treatment in BRCA -mutated ( BRCA m) platinum-sensitive relapsed ovarian. Gynecol Oncol. 2018;149:43–44. 10.1016/j.ygyno.2018.04.094. [Abstract] [CrossRef] [Google Scholar]
294. Mohyuddin GR, Aziz M, Britt A, Wade L, Sun W, Baranda J, Al-Rajabi R, Saeed A, Kasi A. Similar response rates and survival with PARP inhibitors for patients with solid tumors harboring somatic versus germline BRCA mutations: a meta-analysis and systematic review. BMC Cancer. 2020;20(1):507. 10.1186/s12885-020-06948-5. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
295. Wolford JE, Bai J, Moore KN, Kristeleit R, Monk BJ, Tewari KS. Cost-effectiveness of niraparib, rucaparib, and olaparib for treatment of platinum-resistant, recurrent ovarian carcinoma. Gynecol Oncol. 2020;157(2):500–507. 10.1016/j.ygyno.2020.02.030. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
296. Ray-Coquard I, Pautier P, Pignata S, Perol D, Gonzalez-Martin A, Berger R, Fujiwara K, Vergote I, Colombo N, Maenpaa J, et al. Olaparib plus bevacizumab as first-line maintenance in ovarian cancer. N Engl J Med. 2019;381(25):2416–2428. 10.1056/NEJMoa1911361. [Abstract] [CrossRef] [Google Scholar]
297. Arora S, Balasubramaniam S, Zhang H, Berman T, Narayan P, Suzman D, Bloomquist E, Tang S, Gong Y, Sridhara R, et al. FDA approval summary: olaparib monotherapy or in combination with bevacizumab for the maintenance treatment of patients with advanced ovarian cancer. Oncologist. 2021;26(1):e164–e172. 10.1002/onco.13551. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
298. Mirza MR, Avall Lundqvist E, Birrer MJ, dePont CR, Nyvang GB, Malander S, Anttila M, Werner TL, Lund B, Lindahl G, et al. Niraparib plus bevacizumab versus niraparib alone for platinum-sensitive recurrent ovarian cancer (NSGO-AVANOVA2/ENGOT-ov24): a randomised, phase 2, superiority trial. Lancet Oncol. 2019;20(10):1409–1419. 10.1016/S1470-2045(19)30515-7. [Abstract] [CrossRef] [Google Scholar]
299. Lorusso D, Maltese G, Sabatucci I, Cresta S, Matteo C, Ceruti T, D'Incalci M, Zucchetti M, Raspagliesi F, Sonetto C, et al. Phase I study of rucaparib in combination with bevacizumab in ovarian cancer patients: maximum tolerated dose and pharmacokinetic profile. Target Oncol. 2021;16(1):59–68. 10.1007/s11523-020-00780-4. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
300. Makker V, MacKay H, Ray-Coquard I, Levine DA, Westin SN, Aoki D, Oaknin A. Endometrial cancer. Nat Rev Dis Primers. 2021;7(1):88. 10.1038/s41572-021-00324-8. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
301. Kawakami H, Zaanan A, Sinicrope FA. Implications of mismatch repair-deficient status on management of early stage colorectal cancer. J Gastrointest Oncol. 2015;6(6):676–684. [Europe PMC free article] [Abstract] [Google Scholar]
302. Lorenzi M, Amonkar M, Zhang J, Mehta S, Liaw K-L. Epidemiology of microsatellite instability high (MSI-H) and deficient mismatch repair (dMMR) in solid tumors: a structured literature review. J Oncol. 2020;2020:1–17. 10.1155/2020/1807929. [CrossRef] [Google Scholar]
303. Cancer Genome Atlas Research N. Kandoth C, Schultz N, Cherniack AD, Akbani R, Liu Y, Shen H, Robertson AG, Pashtan I, Shen R, et al. Integrated genomic characterization of endometrial carcinoma. Nature. 2013;497(7447):67–73. 10.1038/nature12113. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
304. Le DT, Durham JN, Smith KN, Wang H, Bartlett BR, Aulakh LK, Lu S, Kemberling H, Wilt C, Luber BS, et al. Mismatch repair deficiency predicts response of solid tumors to PD-1 blockade. Science. 2017;357(6349):409–413. 10.1126/science.aan6733. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
305. Howitt BE, Shukla SA, Sholl LM, Ritterhouse LL, Watkins JC, Rodig S, Stover E, Strickland KC, D'Andrea AD, Wu CJ, et al. Association of polymerase e-mutated and microsatellite-instable endometrial cancers with neoantigen load, number of tumor-infiltrating lymphocytes, and expression of PD-1 and PD-L1. JAMA Oncol. 2015;1(9):1319–1323. 10.1001/jamaoncol.2015.2151. [Abstract] [CrossRef] [Google Scholar]
306. 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(1):28. 10.1186/s12943-021-01489-2. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
307. Laken H, Kehry M, McNeeley P, Neben T, Zhang J, Jenkins D, Wilcoxen K. Identification and characterization of TSR-042, a novel anti-human PD-1 therapeutic antibody. European J Cancer 2016; 69.
308. Oaknin A, Tinker AV, Gilbert L, Samouelian V, Mathews C, Brown J, Barretina-Ginesta MP, Moreno V, Gravina A, Abdeddaim C, et al. Clinical activity and safety of the anti-programmed death 1 monoclonal antibody dostarlimab for patients with recurrent or advanced mismatch repair-deficient endometrial cancer: a nonrandomized phase 1 clinical trial. JAMA Oncol. 2020;6(11):1766–1772. 10.1001/jamaoncol.2020.4515. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
309. Oaknin A, Gilbert L, Tinker AV, Brown J, Mathews C, Press J, Sabatier R, O'Malley DM, Samouelian V, Boni V, et al. Safety and antitumor activity of dostarlimab in patients with advanced or recurrent DNA mismatch repair deficient/microsatellite instability-high (dMMR/MSI-H) or proficient/stable (MMRp/MSS) endometrial cancer: interim results from GARNET-a phase I, single-arm study. J Immunother Cancer. 2022 10.1136/jitc-2021-003777. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
310. Cercek A, Lumish M, Sinopoli J, Weiss J, Shia J, Lamendola-Essel M, El Dika IH, Segal N, Shcherba M, Sugarman R, et al. PD-1 blockade in mismatch repair-deficient, locally advanced rectal cancer. N Engl J Med. 2022 10.1056/NEJMoa2201445. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
311. Zur Hausen H. Papillomaviruses and cancer: from basic studies to clinical application. Nat Rev Cancer. 2002;2(5):342–350. 10.1038/nrc798. [Abstract] [CrossRef] [Google Scholar]
312. Schwarz E, Freese UK, Gissmann L, Mayer W, Roggenbuck B, Stremlau A, Zur Hausen H. Structure and transcription of human papillomavirus sequences in cervical carcinoma cells. Nature. 1985;314(6006):111–114. 10.1038/314111a0. [Abstract] [CrossRef] [Google Scholar]
313. de Goeij BE, Satijn D, Freitag CM, Wubbolts R, Bleeker WK, Khasanov A, Zhu T, Chen G, Miao D, van Berkel PH, et al. High turnover of tissue factor enables efficient intracellular delivery of antibody-drug conjugates. Mol Cancer Ther. 2015;14(5):1130–1140. 10.1158/1535-7163.MCT-14-0798. [Abstract] [CrossRef] [Google Scholar]
314. Cocco E, Varughese J, Buza N, Bellone S, Glasgow M, Bellone M, Todeschini P, Carrara L, Silasi DA, Azodi M, et al. Expression of tissue factor in adenocarcinoma and squamous cell carcinoma of the uterine cervix: implications for immunotherapy with hI-con1, a factor VII-IgGFc chimeric protein targeting tissue factor. BMC Cancer. 2011;11:263. 10.1186/1471-2407-11-263. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
315. Kasthuri RS, Taubman MB, Mackman N. Role of tissue factor in cancer. J Clin Oncol. 2009;27(29):4834–4838. 10.1200/JCO.2009.22.6324. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
316. Breij EC, de Goeij BE, Verploegen S, Schuurhuis DH, Amirkhosravi A, Francis J, Miller VB, Houtkamp M, Bleeker WK, Satijn D, et al. An antibody-drug conjugate that targets tissue factor exhibits potent therapeutic activity against a broad range of solid tumors. Cancer Res. 2014;74(4):1214–1226. 10.1158/0008-5472.CAN-13-2440. [Abstract] [CrossRef] [Google Scholar]
317. Hong DS, Concin N, Vergote I, de Bono JS, Slomovitz BM, Drew Y, Arkenau HT, Machiels JP, Spicer JF, Jones R, et al. Tisotumab vedotin in previously treated recurrent or metastatic cervical cancer. Clin Cancer Res. 2020;26(6):1220–1228. 10.1158/1078-0432.CCR-19-2962. [Abstract] [CrossRef] [Google Scholar]
318. Coleman RL, Lorusso D, Gennigens C, Gonzalez-Martin A, Randall L, Cibula D, Lund B, Woelber L, Pignata S, Forget F, et al. Efficacy and safety of tisotumab vedotin in previously treated recurrent or metastatic cervical cancer (innovaTV 204/GOG-3023/ENGOT-cx6): a multicentre, open-label, single-arm, phase 2 study. Lancet Oncol. 2021;22(5):609–619. 10.1016/S1470-2045(21)00056-5. [Abstract] [CrossRef] [Google Scholar]
319. Ormond AB, Freeman HS. Dye sensitizers for photodynamic therapy. Materials. 2013;6(3):817–840. 10.3390/ma6030817. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
320. Biel M. Advances in photodynamic therapy for the treatment of head and neck cancers. Lasers Surg Med. 2006;38(5):349–355. 10.1002/lsm.20368. [Abstract] [CrossRef] [Google Scholar]
321. Yano T, Hatogai K, Morimoto H, Yoda Y, Kaneko K. Photodynamic therapy for esophageal cancer. Ann Transl Med. 2014;2(3):29. [Europe PMC free article] [Abstract] [Google Scholar]
322. Cunningham D, Starling N, Rao S, Iveson T, Nicolson M, Coxon F, Middleton G, Daniel F, Oates J, Norman AR, et al. Capecitabine and oxaliplatin for advanced esophagogastric cancer. N Engl J Med. 2008;358(1):36–46. 10.1056/NEJMoa073149. [Abstract] [CrossRef] [Google Scholar]
323. Sun JM, Shen L, Shah MA, Enzinger P, Adenis A, Doi T, Kojima T, Metges JP, Li Z, Kim SB, et al. Pembrolizumab plus chemotherapy versus chemotherapy alone for first-line treatment of advanced oesophageal cancer (KEYNOTE-590): a randomised, placebo-controlled, phase 3 study. Lancet. 2021;398(10302):759–771. 10.1016/S0140-6736(21)01234-4. [Abstract] [CrossRef] [Google Scholar]
324. Smith NR, Baker D, James NH, Ratcliffe K, Jenkins M, Ashton SE, Sproat G, Swann R, Gray N, Ryan A, et al. Vascular endothelial growth factor receptors VEGFR-2 and VEGFR-3 are localized primarily to the vasculature in human primary solid cancers. Clin Cancer Res. 2010;16(14):3548–3561. 10.1158/1078-0432.CCR-09-2797. [Abstract] [CrossRef] [Google Scholar]
325. Lu D, Shen J, Vil MD, Zhang H, Jimenez X, Bohlen P, Witte L, Zhu Z. Tailoring in vitro selection for a picomolar affinity human antibody directed against vascular endothelial growth factor receptor 2 for enhanced neutralizing activity. J Biol Chem. 2003;278(44):43496–43507. 10.1074/jbc.M307742200. [Abstract] [CrossRef] [Google Scholar]
326. Franklin MC, Navarro EC, Wang Y, Patel S, Singh P, Zhang Y, Persaud K, Bari A, Griffith H, Shen L, et al. The structural basis for the function of two anti-VEGF receptor 2 antibodies. Structure. 2011;19(8):1097–1107. 10.1016/j.str.2011.01.019. [Abstract] [CrossRef] [Google Scholar]
327. Poole RM, Vaidya A. Ramucirumab: first global approval. Drugs. 2014;74(9):1047–1058. 10.1007/s40265-014-0244-2. [Abstract] [CrossRef] [Google Scholar]
328. Fuchs CS, Tomasek J, Yong CJ, Dumitru F, Passalacqua R, Goswami C, Safran H, Dos Santos LV, Aprile G, Ferry DR, et al. Ramucirumab monotherapy for previously treated advanced gastric or gastro-oesophageal junction adenocarcinoma (REGARD): an international, randomised, multicentre, placebo-controlled, phase 3 trial. Lancet. 2014;383(9911):31–39. 10.1016/S0140-6736(13)61719-5. [Abstract] [CrossRef] [Google Scholar]
329. Wilke H, Muro K, Van Cutsem E, Oh SC, Bodoky G, Shimada Y, Hironaka S, Sugimoto N, Lipatov O, Kim TY, et al. Ramucirumab plus paclitaxel versus placebo plus paclitaxel in patients with previously treated advanced gastric or gastro-oesophageal junction adenocarcinoma (RAINBOW): a double-blind, randomised phase 3 trial. Lancet Oncol. 2014;15(11):1224–1235. 10.1016/S1470-2045(14)70420-6. [Abstract] [CrossRef] [Google Scholar]
330. Joensuu H, Roberts PJ, Sarlomo-Rikala M, Andersson LC, Tervahartiala P, Tuveson D, Silberman S, Capdeville R, Dimitrijevic S, Druker B, et al. Effect of the tyrosine kinase inhibitor STI571 in a patient with a metastatic gastrointestinal stromal tumor. N Engl J Med. 2001;344(14):1052–1056. 10.1056/NEJM200104053441404. [Abstract] [CrossRef] [Google Scholar]
331. Miettinen M, Sarlomo-Rikala M, Sobin LH. Mesenchymal tumors of muscularis mucosae of colon and rectum are benign leiomyomas that should be separated from gastrointestinal stromal tumors–a clinicopathologic and immunohistochemical study of eighty-eight cases. Mod Pathol. 2001;14(10):950–956. 10.1038/modpathol.3880417. [Abstract] [CrossRef] [Google Scholar]
332. Miettinen M, Sarlomo-Rikala M, Lasota J. Gastrointestinal stromal tumors: recent advances in understanding of their biology. Hum Pathol. 1999;30(10):1213–1220. 10.1016/S0046-8177(99)90040-0. [Abstract] [CrossRef] [Google Scholar]
333. Corless CL, Fletcher JA, Heinrich MC. Biology of gastrointestinal stromal tumors. J Clin Oncol. 2004;22(18):3813–3825. 10.1200/JCO.2004.05.140. [Abstract] [CrossRef] [Google Scholar]
334. Corless CL, Barnett CM, Heinrich MC. Gastrointestinal stromal tumours: origin and molecular oncology. Nat Rev Cancer. 2011;11(12):865–878. 10.1038/nrc3143. [Abstract] [CrossRef] [Google Scholar]
335. Demetri GD, von Mehren M, Blanke CD, Van den Abbeele AD, Eisenberg B, Roberts PJ, Heinrich MC, Tuveson DA, Singer S, Janicek M, et al. Efficacy and safety of imatinib mesylate in advanced gastrointestinal stromal tumors. N Engl J Med. 2002;347(7):472–480. 10.1056/NEJMoa020461. [Abstract] [CrossRef] [Google Scholar]
336. Verweij J, Casali PG, Zalcberg J, LeCesne A, Reichardt P, Blay JY, Issels R, van Oosterom A, Hogendoorn PC, Van Glabbeke M, et al. Progression-free survival in gastrointestinal stromal tumours with high-dose imatinib: randomised trial. Lancet. 2004;364(9440):1127–1134. 10.1016/S0140-6736(04)17098-0. [Abstract] [CrossRef] [Google Scholar]
337. Gajiwala KS, Wu JC, Christensen J, Deshmukh GD, Diehl W, DiNitto JP, English JM, Greig MJ, He YA, Jacques SL, et al. KIT kinase mutants show unique mechanisms of drug resistance to imatinib and sunitinib in gastrointestinal stromal tumor patients. Proc Natl Acad Sci U S A. 2009;106(5):1542–1547. 10.1073/pnas.0812413106. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
338. Deeks ED, Keating GM. Sunitinib. Drugs. 2006;66(17):2255–2266. 10.2165/00003495-200666170-00007. [Abstract] [CrossRef] [Google Scholar]
339. Joensuu H. Sunitinib for imatinib-resistant GIST. Lancet. 2006;368(9544):1303–1304. 10.1016/S0140-6736(06)69489-0. [Abstract] [CrossRef] [Google Scholar]
340. Heinrich MC, Maki RG, Corless CL, Antonescu CR, Harlow A, Griffith D, Town A, McKinley A, Ou WB, Fletcher JA, et al. Primary and secondary kinase genotypes correlate with the biological and clinical activity of sunitinib in imatinib-resistant gastrointestinal stromal tumor. J Clin Oncol. 2008;26(33):5352–5359. 10.1200/JCO.2007.15.7461. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
341. Prenen H, Cools J, Mentens N, Folens C, Sciot R, Schoffski P, Van Oosterom A, Marynen P, Debiec-Rychter M. Efficacy of the kinase inhibitor SU11248 against gastrointestinal stromal tumor mutants refractory to imatinib mesylate. Clin Cancer Res. 2006;12(8):2622–2627. 10.1158/1078-0432.CCR-05-2275. [Abstract] [CrossRef] [Google Scholar]
342. Heinrich MC, Jones RL, von Mehren M, Schoffski P, Serrano C, Kang YK, Cassier PA, Mir O, Eskens F, Tap WD, et al. Avapritinib in advanced PDGFRA D842V-mutant gastrointestinal stromal tumour (NAVIGATOR): a multicentre, open-label, phase 1 trial. Lancet Oncol. 2020;21(7):935–946. 10.1016/S1470-2045(20)30269-2. [Abstract] [CrossRef] [Google Scholar]
343. Dhillon S. Avapritinib: first approval. Drugs. 2020;80(4):433–439. 10.1007/s40265-020-01275-2. [Abstract] [CrossRef] [Google Scholar]
344. Evans EK, Gardino AK, Kim JL, Hodous BL, Shutes A, Davis A, Zhu XJ, Schmidt-Kittler O, Wilson D, Wilson K, et al. A precision therapy against cancers driven by KIT/PDGFRA mutations. Sci Transl Med. 2017;9(414):eaao1690. 10.1126/scitranslmed.aao1690. [Abstract] [CrossRef] [Google Scholar]
345. Dhillon S. Ripretinib: first approval. Drugs. 2020;80(11):1133–1138. 10.1007/s40265-020-01348-2. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
346. Smith BD, Kaufman MD, Lu WP, Gupta A, Leary CB, Wise SC, Rutkoski TJ, Ahn YM, Al-Ani G, Bulfer SL, et al. Ripretinib (DCC-2618) is a switch control kinase inhibitor of a broad spectrum of oncogenic and drug-resistant KIT and PDGFRA variants. Cancer Cell. 2019;35(5):738–751 e739. 10.1016/j.ccell.2019.04.006. [Abstract] [CrossRef] [Google Scholar]
347. Grunewald S, Klug LR, Muhlenberg T, Lategahn J, Falkenhorst J, Town A, Ehrt C, Wardelmann E, Hartmann W, Schildhaus HU, et al. Resistance to avapritinib in PDGFRA-driven GIST is caused by secondary mutations in the PDGFRA kinase domain. Cancer Discov. 2021;11(1):108–125. 10.1158/2159-8290.CD-20-0487. [Abstract] [CrossRef] [Google Scholar]
348. Frilling A, Akerstrom G, Falconi M, Pavel M, Ramos J, Kidd M, Modlin IM. Neuroendocrine tumor disease: an evolving landscape. Endocr Relat Cancer. 2012;19(5):R163–185. 10.1530/ERC-12-0024. [Abstract] [CrossRef] [Google Scholar]
349. Cives M, Strosberg JR. Gastroenteropancreatic neuroendocrine tumors. CA Cancer J Clin. 2018;68(6):471–487. 10.3322/caac.21493. [Abstract] [CrossRef] [Google Scholar]
350. Papotti M, Bongiovanni M, Volante M, Allia E, Landolfi S, Helboe L, Schindler M, Cole SL, Bussolati G. Expression of somatostatin receptor types 1–5 in 81 cases of gastrointestinal and pancreatic endocrine tumors. A correlative immunohistochemical and reverse-transcriptase polymerase chain reaction analysis. Virchows Arch. 2002;440(5):461–475. 10.1007/s00428-002-0609-x. [Abstract] [CrossRef] [Google Scholar]
351. Reichlin S. Somatostatin. N Engl J Med. 1983;309(24):1495–1501. 10.1056/NEJM198312153092406. [Abstract] [CrossRef] [Google Scholar]
352. Kvols LK, Reubi JC, Horisberger U, Moertel CG, Rubin J, Charboneau JW. The presence of somatostatin receptors in malignant neuroendocrine tumor tissue predicts responsiveness to octreotide. Yale J Biol Med. 1992;65(5):505–518. [Europe PMC free article] [Abstract] [Google Scholar]
353. Kuhar MB, Kuhar BM. Evaluation of selected new drugs: 1988. AAOHN J. 1989;37(10):428–433. 10.1177/216507998903701005. [Abstract] [CrossRef] [Google Scholar]
354. van der Zwan WA, Bodei L, Mueller-Brand J, de Herder WW, Kvols LK, Kwekkeboom DJ. GEPNETs update: radionuclide therapy in neuroendocrine tumors. Eur J Endocrinol. 2015;172(1):R1–8. 10.1530/EJE-14-0488. [Abstract] [CrossRef] [Google Scholar]
355. Strosberg J, El-Haddad G, Wolin E, Hendifar A, Yao J, Chasen B, Mittra E, Kunz PL, Kulke MH, Jacene H, et al. Phase 3 trial of (177)Lu-dotatate for midgut neuroendocrine tumors. N Engl J Med. 2017;376(2):125–135. 10.1056/NEJMoa1607427. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
356. Reubi JC, Schar JC, Waser B, Wenger S, Heppeler A, Schmitt JS, Macke HR. Affinity profiles for human somatostatin receptor subtypes SST1-SST5 of somatostatin radiotracers selected for scintigraphic and radiotherapeutic use. Eur J Nucl Med. 2000;27(3):273–282. 10.1007/s002590050034. [Abstract] [CrossRef] [Google Scholar]
357. Lee WS, Lee KW, Heo JS, Kim SJ, Choi SH, Kim YI, Joh JW. Comparison of combined hepatocellular and cholangiocarcinoma with hepatocellular carcinoma and intrahepatic cholangiocarcinoma. Surg Today. 2006;36(10):892–897. 10.1007/s00595-006-3276-8. [Abstract] [CrossRef] [Google Scholar]
358. Mazzaferro V, El-Rayes BF, Droz Dit Busset M, Cotsoglou C, Harris WP, Damjanov N, Masi G, Rimassa L, Personeni N, Braiteh F, et al. Derazantinib (ARQ 087) in advanced or inoperable FGFR2 gene fusion-positive intrahepatic cholangiocarcinoma. Br J Cancer. 2019;120(2):165–171. 10.1038/s41416-018-0334-0. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
359. Silverman IM, Hollebecque A, Friboulet L, Owens S, Newton RC, Zhen H, Feliz L, Zecchetto C, Melisi D, Burn TC. Clinicogenomic analysis of FGFR2-rearranged cholangiocarcinoma identifies correlates of response and mechanisms of resistance to pemigatinib. Cancer Discov. 2021;11(2):326–339. 10.1158/2159-8290.CD-20-0766. [Abstract] [CrossRef] [Google Scholar]
360. Wu L, Zhang C, He C, Qian D, Lu L, Sun Y, Xu M, Zhuo J, Liu PCC, Klabe R, et al. Discovery of pemigatinib: a potent and selective fibroblast growth factor receptor (FGFR) inhibitor. J Med Chem. 2021;64(15):10666–10679. 10.1021/acs.jmedchem.1c00713. [Abstract] [CrossRef] [Google Scholar]
361. Hoy SM. Pemigatinib: first approval. Drugs. 2020;80(9):923–929. 10.1007/s40265-020-01330-y. [Abstract] [CrossRef] [Google Scholar]
362. Abou-Alfa GK, Sahai V, Hollebecque A, Vaccaro G, Melisi D, Al-Rajabi R, Paulson AS, Borad MJ, Gallinson D, Murphy AG, et al. Pemigatinib for previously treated, locally advanced or metastatic cholangiocarcinoma: a multicentre, open-label, phase 2 study. Lancet Oncol. 2020;21(5):671–684. 10.1016/S1470-2045(20)30109-1. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
363. Bekaii-Saab TS, Valle JW, Van Cutsem E, Rimassa L, Furuse J, Ioka T, Melisi D, Macarulla T, Bridgewater J, Wasan H, et al. FIGHT-302: first-line pemigatinib vs gemcitabine plus cisplatin for advanced cholangiocarcinoma with FGFR2 rearrangements. Future Oncol. 2020;16(30):2385–2399. 10.2217/fon-2020-0429. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
364. Guagnano V, Furet P, Spanka C, Bordas V, Le Douget M, Stamm C, Brueggen J, Jensen MR, Schnell C, Schmid H, et al. Discovery of 3-(2,6-dichloro-3,5-dimethoxy-phenyl)-1-{6-[4-(4-ethyl-piperazin-1-yl)-phenylamin o]-pyrimidin-4-yl}-1-methyl-urea (NVP-BGJ398), a potent and selective inhibitor of the fibroblast growth factor receptor family of receptor tyrosine kinase. J Med Chem. 2011;54(20):7066–7083. 10.1021/jm2006222. [Abstract] [CrossRef] [Google Scholar]
365. Guagnano V, Kauffmann A, Wohrle S, Stamm C, Ito M, Barys L, Pornon A, Yao Y, Li F, Zhang Y, et al. FGFR genetic alterations predict for sensitivity to NVP-BGJ398, a selective pan-FGFR inhibitor. Cancer Discov. 2012;2(12):1118–1133. 10.1158/2159-8290.CD-12-0210. [Abstract] [CrossRef] [Google Scholar]
366. Nogova L, Sequist LV, Perez Garcia JM, Andre F, Delord JP, Hidalgo M, Schellens JH, Cassier PA, Camidge DR, Schuler M, et al. Evaluation of BGJ398, a fibroblast growth factor receptor 1–3 kinase inhibitor, in patients with advanced solid tumors harboring genetic alterations in fibroblast growth factor receptors: results of a global phase I, dose-escalation and dose-expansion study. J Clin Oncol. 2017;35(2):157–165. 10.1200/JCO.2016.67.2048. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
367. Javle M, Lowery M, Shroff RT, Weiss KH, Springfeld C, Borad MJ, Ramanathan RK, Goyal L, Sadeghi S, Macarulla T, et al. Phase II study of BGJ398 in patients with FGFR-altered advanced cholangiocarcinoma. J Clin Oncol. 2018;36(3):276–282. 10.1200/JCO.2017.75.5009. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
368. Goyal L, Saha SK, Liu LY, Siravegna G, Leshchiner I, Ahronian LG, Lennerz JK, Vu P, Deshpande V, Kambadakone A, et al. Polyclonal secondary FGFR2 mutations drive acquired resistance to FGFR inhibition in patients with FGFR2 fusion-positive cholangiocarcinoma. Cancer Discov. 2017;7(3):252–263. 10.1158/2159-8290.CD-16-1000. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
369. Krook MA, Bonneville R, Chen HZ, Reeser JW, Wing MR, Martin DM, Smith AM, Dao T, Samorodnitsky E, Paruchuri A, et al. Tumor heterogeneity and acquired drug resistance in FGFR2-fusion-positive cholangiocarcinoma through rapid research autopsy. Cold Spring Harb Mol Case Stud. 2019 10.1101/mcs.a004002. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
370. Cancer Genome Atlas Research Network Electronic address aadhe, Cancer Genome Atlas Research N: Integrated genomic characterization of pancreatic ductal adenocarcinoma. Cancer Cell. 2017;32(2):185–203 e113. 10.1016/j.ccell.2017.07.007. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
371. Ligorio M, Sil S, Malagon-Lopez J, Nieman LT, Misale S, Di Pilato M, Ebright RY, Karabacak MN, Kulkarni AS, Liu A, et al. Stromal microenvironment shapes the intratumoral architecture of pancreatic cancer. Cell. 2019;178(1):160–175 e127. 10.1016/j.cell.2019.05.012. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
372. Plunkett W, Huang P, Xu YZ, Heinemann V, Grunewald R, Gandhi V. Gemcitabine: metabolism, mechanisms of action, and self-potentiation. Semin Oncol. 1995;22(4 Suppl 11):3–10. [Abstract] [Google Scholar]
373. Burris HA, 3rd, Moore MJ, Andersen J, Green MR, Rothenberg ML, Modiano MR, Cripps MC, Portenoy RK, Storniolo AM, Tarassoff P, et al. Improvements in survival and clinical benefit with gemcitabine as first-line therapy for patients with advanced pancreas cancer: a randomized trial. J Clin Oncol. 1997;15(6):2403–2413. 10.1200/JCO.1997.15.6.2403. [Abstract] [CrossRef] [Google Scholar]
374. Conroy T, Desseigne F, Ychou M, Bouche O, Guimbaud R, Becouarn Y, Adenis A, Raoul JL, Gourgou-Bourgade S, de la Fouchardiere C, et al. FOLFIRINOX versus gemcitabine for metastatic pancreatic cancer. N Engl J Med. 2011;364(19):1817–1825. 10.1056/NEJMoa1011923. [Abstract] [CrossRef] [Google Scholar]
375. Von Hoff DD, Ervin T, Arena FP, Chiorean EG, Infante J, Moore M, Seay T, Tjulandin SA, Ma WW, Saleh MN, et al. Increased survival in pancreatic cancer with nab-paclitaxel plus gemcitabine. N Engl J Med. 2013;369(18):1691–1703. 10.1056/NEJMoa1304369. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
376. Mizrahi JD, Surana R, Valle JW, Shroff RT. Pancreatic cancer. Lancet. 2020;395(10242):2008–2020. 10.1016/S0140-6736(20)30974-0. [Abstract] [CrossRef] [Google Scholar]
377. Cho I, Kang H, Jo J, Lee H, Chung M, Park J, Park S, Song S, Park M, An C, et al. FOLFIRINOX versus gemcitabine plus nab-paclitaxel for treatment of metastatic pancreatic cancer: a single-center cohort study. Annals Oncol. 2018 10.1093/annonc/mdy151.160. [CrossRef] [Google Scholar]
378. Kawato Y, Aonuma M, Hirota Y, Kuga H, Sato K. Intracellular roles of SN-38, a metabolite of the camptothecin derivative CPT-11, in the antitumor effect of CPT-11. Cancer Res. 1991;51(16):4187–4191. [Abstract] [Google Scholar]
379. Jonsson E, Dhar S, Jonsson B, Nygren P, Graf W, Larsson R. Differential activity of topotecan, irinotecan and SN-38 in fresh human tumour cells but not in cell lines. Eur J Cancer. 2000;36(16):2120–2127. 10.1016/S0959-8049(00)00289-6. [Abstract] [CrossRef] [Google Scholar]
380. Raymond E, Faivre S, Woynarowski JM, Chaney SG. Oxaliplatin: mechanism of action and antineoplastic activity. Semin Oncol. 1998;25(2 Suppl 5):4–12. [Abstract] [Google Scholar]
381. Rottenberg S, Disler C, Perego P. The rediscovery of platinum-based cancer therapy. Nat Rev Cancer. 2021;21(1):37–50. 10.1038/s41568-020-00308-y. [Abstract] [CrossRef] [Google Scholar]
382. Vaisman A, Varchenko M, Umar A, Kunkel TA, Risinger JI, Barrett JC, Hamilton TC, Chaney SG. The role of hMLH1, hMSH3, and hMSH6 defects in cisplatin and oxaliplatin resistance: correlation with replicative bypass of platinum-DNA adducts. Cancer Res. 1998;58(16):3579–3585. [Abstract] [Google Scholar]
383. Wen X, Buckley B, McCandlish E, Goedken MJ, Syed S, Pelis R, Manautou JE, Aleksunes LM. Transgenic expression of the human MRP2 transporter reduces cisplatin accumulation and nephrotoxicity in Mrp2-null mice. Am J Pathol. 2014;184(5):1299–1308. 10.1016/j.ajpath.2014.01.025. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
384. Myint K, Li Y, Paxton J, McKeage M. Multidrug resistance-associated protein 2 (MRP2) mediated transport of oxaliplatin-derived platinum in membrane vesicles. PLoS ONE. 2015;10(7):e0130727. 10.1371/journal.pone.0130727. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
385. Myint K, Biswas R, Li Y, Jong N, Jamieson S, Liu J, Han C, Squire C, Merien F, Lu J, et al. Identification of MRP2 as a targetable factor limiting oxaliplatin accumulation and response in gastrointestinal cancer. Sci Rep. 2019;9(1):2245. 10.1038/s41598-019-38667-8. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
386. Hall MD, Okabe M, Shen DW, Liang XJ, Gottesman MM. The role of cellular accumulation in determining sensitivity to platinum-based chemotherapy. Annu Rev Pharmacol Toxicol. 2008;48:495–535. 10.1146/annurev.pharmtox.48.080907.180426. [Abstract] [CrossRef] [Google Scholar]
387. Falcone A, Ricci S, Brunetti I, Pfanner E, Allegrini G, Barbara C, Crino L, Benedetti G, Evangelista W, Fanchini L, et al. Phase III trial of infusional fluorouracil, leucovorin, oxaliplatin, and irinotecan (FOLFOXIRI) compared with infusional fluorouracil, leucovorin, and irinotecan (FOLFIRI) as first-line treatment for metastatic colorectal cancer: the Gruppo Oncologico Nord Ovest. J Clin Oncol. 2007;25(13):1670–1676. 10.1200/JCO.2006.09.0928. [Abstract] [CrossRef] [Google Scholar]
388. Landry JC, Feng Y, Prabhu RS, Cohen SJ, Staley CA, Whittington R, Sigurdson ER, Nimeiri H, Verma U, Benson AB. Phase II trial of preoperative radiation with concurrent capecitabine, oxaliplatin, and bevacizumab followed by surgery and postoperative 5-fluorouracil, leucovorin, oxaliplatin (FOLFOX), and bevacizumab in patients with locally advanced rectal cancer: 5-year clinical outcomes ECOG-ACRIN cancer research group E3204. Oncologist. 2015;20(6):615–616. 10.1634/theoncologist.2015-0106. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
389. Saltz LB, Clarke S, Diaz-Rubio E, Scheithauer W, Figer A, Wong R, Koski S, Lichinitser M, Yang TS, Rivera F, et al. Bevacizumab in combination with oxaliplatin-based chemotherapy as first-line therapy in metastatic colorectal cancer: a randomized phase III study. J Clin Oncol. 2008;26(12):2013–2019. 10.1200/JCO.2007.14.9930. [Abstract] [CrossRef] [Google Scholar]
390. McKay JA, Murray LJ, Curran S, Ross VG, Clark C, Murray GI, Cassidy J, McLeod HL. Evaluation of the epidermal growth factor receptor (EGFR) in colorectal tumours and lymph node metastases. Eur J Cancer. 2002;38(17):2258–2264. 10.1016/S0959-8049(02)00234-4. [Abstract] [CrossRef] [Google Scholar]
391. Spano JP, Lagorce C, Atlan D, Milano G, Domont J, Benamouzig R, Attar A, Benichou J, Martin A, Morere JF, et al. Impact of EGFR expression on colorectal cancer patient prognosis and survival. Ann Oncol. 2005;16(1):102–108. 10.1093/annonc/mdi006. [Abstract] [CrossRef] [Google Scholar]
392. Barber TD, Vogelstein B, Kinzler KW, Velculescu VE. Somatic mutations of EGFR in colorectal cancers and glioblastomas. N Engl J Med. 2004;351(27):2883. 10.1056/NEJM200412303512724. [Abstract] [CrossRef] [Google Scholar]
393. Li S, Schmitz KR, Jeffrey PD, Wiltzius JJ, Kussie P, Ferguson KM. Structural basis for inhibition of the epidermal growth factor receptor by cetuximab. Cancer Cell. 2005;7(4):301–311. 10.1016/j.ccr.2005.03.003. [Abstract] [CrossRef] [Google Scholar]
394. Montagut C, Dalmases A, Bellosillo B, Crespo M, Pairet S, Iglesias M, Salido M, Gallen M, Marsters S, Tsai SP, et al. Identification of a mutation in the extracellular domain of the epidermal growth factor receptor conferring cetuximab resistance in colorectal cancer. Nat Med. 2012;18(2):221–223. 10.1038/nm.2609. [Abstract] [CrossRef] [Google Scholar]
395. Arena S, Bellosillo B, Siravegna G, Martinez A, Canadas I, Lazzari L, Ferruz N, Russo M, Misale S, Gonzalez I, et al. Emergence of multiple EGFR extracellular mutations during cetuximab treatment in colorectal cancer. Clin Cancer Res. 2015;21(9):2157–2166. 10.1158/1078-0432.CCR-14-2821. [Abstract] [CrossRef] [Google Scholar]
396. Sickmier EA, Kurzeja RJ, Michelsen K, Vazir M, Yang E, Tasker AS. The panitumumab EGFR complex reveals a binding mechanism that overcomes cetuximab induced resistance. PLoS One. 2016;11(9):e0163366. 10.1371/journal.pone.0163366. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
397. Price TJ, Peeters M, Kim TW, Li J, Cascinu S, Ruff P, Suresh AS, Thomas A, Tjulandin S, Zhang K, et al. Panitumumab versus cetuximab in patients with chemotherapy-refractory wild-type KRAS exon 2 metastatic colorectal cancer (ASPECCT): a randomised, multicentre, open-label, non-inferiority phase 3 study. Lancet Oncol. 2014;15(6):569–579. 10.1016/S1470-2045(14)70118-4. [Abstract] [CrossRef] [Google Scholar]
398. Ferrara N, Hillan KJ, Gerber HP, Novotny W. Discovery and development of bevacizumab, an anti-VEGF antibody for treating cancer. Nat Rev Drug Discov. 2004;3(5):391–400. 10.1038/nrd1381. [Abstract] [CrossRef] [Google Scholar]
399. Ferrara N. Vascular endothelial growth factor: basic science and clinical progress. Endocr Rev. 2004;25(4):581–611. 10.1210/er.2003-0027. [Abstract] [CrossRef] [Google Scholar]
400. Uemura A, Fruttiger M, D'Amore PA, De Falco S, Joussen AM, Sennlaub F, Brunck LR, Johnson KT, Lambrou GN, Rittenhouse KD, et al. VEGFR1 signaling in retinal angiogenesis and microinflammation. Prog Retin Eye Res. 2021 10.1016/j.preteyeres.2021. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
401. Joukov V, Pajusola K, Kaipainen A, Chilov D, Lahtinen I, Kukk E, Saksela O, Kalkkinen N, Alitalo K. A novel vascular endothelial growth factor, VEGF-C, is a ligand for the Flt4 (VEGFR-3) and KDR (VEGFR-2) receptor tyrosine kinases. EMBO J. 1996;15(2):290–298. 10.1002/j.1460-2075.1996.tb00359.x. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
402. Achen MG, Jeltsch M, Kukk E, Makinen T, Vitali A, Wilks AF, Alitalo K, Stacker SA. Vascular endothelial growth factor D (VEGF-D) is a ligand for the tyrosine kinases VEGF receptor 2 (Flk1) and VEGF receptor 3 (Flt4) Proc Natl Acad Sci U S A. 1998;95(2):548–553. 10.1073/pnas.95.2.548. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
403. Cremolini C, Loupakis F, Antoniotti C, Lupi C, Sensi E, Lonardi S, Mezi S, Tomasello G, Ronzoni M, Zaniboni A, et al. FOLFOXIRI plus bevacizumab versus FOLFIRI plus bevacizumab as first-line treatment of patients with metastatic colorectal cancer: updated overall survival and molecular subgroup analyses of the open-label, phase 3 TRIBE study. Lancet Oncol. 2015;16(13):1306–1315. 10.1016/S1470-2045(15)00122-9. [Abstract] [CrossRef] [Google Scholar]
404. McCormack PL, Keam SJ. Bevacizumab: a review of its use in metastatic colorectal cancer. Drugs. 2008;68(4):487–506. 10.2165/00003495-200868040-00009. [Abstract] [CrossRef] [Google Scholar]
405. Holash J, Davis S, Papadopoulos N, Croll SD, Ho L, Russell M, Boland P, Leidich R, Hylton D, Burova E, et al. VEGF-Trap: a VEGF blocker with potent antitumor effects. Proc Natl Acad Sci U S A. 2002;99(17):11393–11398. 10.1073/pnas.172398299. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
406. Clarke JM, Hurwitz HI. Ziv-aflibercept: binding to more than VEGF-A–does more matter? Nat Rev Clin Oncol. 2013;10(1):10–11. 10.1038/nrclinonc.2012.197. [Abstract] [CrossRef] [Google Scholar]
407. Van Cutsem E, Tabernero J, Lakomy R, Prenen H, Prausova J, Macarulla T, Ruff P, van Hazel GA, Moiseyenko V, Ferry D, et al. Addition of aflibercept to fluorouracil, leucovorin, and irinotecan improves survival in a phase III randomized trial in patients with metastatic colorectal cancer previously treated with an oxaliplatin-based regimen. J Clin Oncol. 2012;30(28):3499–3506. 10.1200/JCO.2012.42.8201. [Abstract] [CrossRef] [Google Scholar]
408. Van Cutsem E, Nordlinger B, Cervantes A, Group EGW Advanced colorectal cancer: ESMO Clinical Practice Guidelines for treatment. Ann Oncol. 2010;21(Suppl 5):v93–97. 10.1093/annonc/mdq222. [Abstract] [CrossRef] [Google Scholar]
409. Gschwind A, Fischer OM, Ullrich A. The discovery of receptor tyrosine kinases: targets for cancer therapy. Nat Rev Cancer. 2004;4(5):361–370. 10.1038/nrc1360. [Abstract] [CrossRef] [Google Scholar]
410. De Roock W, De Vriendt V, Normanno N, Ciardiello F, Tejpar S. KRAS, BRAF, PIK3CA, and PTEN mutations: implications for targeted therapies in metastatic colorectal cancer. Lancet Oncol. 2011;12(6):594–603. 10.1016/S1470-2045(10)70209-6. [Abstract] [CrossRef] [Google Scholar]
411. Wilhelm SM, Dumas J, Adnane L, Lynch M, Carter CA, Schutz G, Thierauch KH, Zopf D. Regorafenib (BAY 73–4506): a new oral multikinase inhibitor of angiogenic, stromal and oncogenic receptor tyrosine kinases with potent preclinical antitumor activity. Int J Cancer. 2011;129(1):245–255. 10.1002/ijc.25864. [Abstract] [CrossRef] [Google Scholar]
412. Ettrich TJ, Seufferlein T: Regorafenib. Small Molecules in Oncology. 2018; 45–56. [Abstract]
413. Grothey A, Van Cutsem E, Sobrero A, Siena S, Falcone A, Ychou M, Humblet Y, Bouche O, Mineur L, Barone C, et al. Regorafenib monotherapy for previously treated metastatic colorectal cancer (CORRECT): an international, multicentre, randomised, placebo-controlled, phase 3 trial. Lancet. 2013;381(9863):303–312. 10.1016/S0140-6736(12)61900-X. [Abstract] [CrossRef] [Google Scholar]
414. Dhillon S. Regorafenib: a review in metastatic colorectal cancer. Drugs. 2018;78(11):1133–1144. 10.1007/s40265-018-0938-y. [Abstract] [CrossRef] [Google Scholar]
415. Zhou J, Liu Y, Zhang Y, Li Q, Cao Y. Modeling tumor evolutionary dynamics to predict clinical outcomes for patients with metastatic colorectal cancer: a retrospective analysis. Cancer Res. 2020;80(3):591–601. 10.1158/0008-5472.CAN-19-1940. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
416. van der Velden DL, Opdam FL, Voest EE. TAS-102 for treatment of advanced colorectal cancers that are no longer responding to other therapies. Clin Cancer Res. 2016;22(12):2835–2839. 10.1158/1078-0432.CCR-15-2783. [Abstract] [CrossRef] [Google Scholar]
417. Lenz HJ, Stintzing S, Loupakis F. TAS-102, a novel antitumor agent: a review of the mechanism of action. Cancer Treat Rev. 2015;41(9):777–783. 10.1016/j.ctrv.2015.06.001. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
418. Tanaka N, Sakamoto K, Okabe H, Fujioka A, Yamamura K, Nakagawa F, Nagase H, Yokogawa T, Oguchi K, Ishida K, et al. Repeated oral dosing of TAS-102 confers high trifluridine incorporation into DNA and sustained antitumor activity in mouse models. Oncol Rep. 2014;32(6):2319–2326. 10.3892/or.2014.3487. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
419. Bijnsdorp IV, Peters GJ, Temmink OH, Fukushima M, Kruyt FA. Differential activation of cell death and autophagy results in an increased cytotoxic potential for trifluorothymidine compared to 5-fluorouracil in colon cancer cells. Int J Cancer. 2010;126(10):2457–2468. [Abstract] [Google Scholar]
420. Emura T, Murakami Y, Nakagawa F, Fukushima M, Kitazato K. A novel antimetabolite, TAS-102 retains its effect on FU-related resistant cancer cells. Int J Mol Med. 2004;13(4):545–549. [Abstract] [Google Scholar]
421. Emura T, Suzuki N, Yamaguchi M, Ohshimo H, Fukushima M. A novel combination antimetabolite, TAS-102, exhibits antitumor activity in FU-resistant human cancer cells through a mechanism involving FTD incorporation in DNA. Int J Oncol. 2004;25(3):571–578. [Abstract] [Google Scholar]
422. Burness CB, Duggan ST. Trifluridine/Tipiracil: a review in metastatic colorectal cancer. Drugs. 2016;76(14):1393–1402. 10.1007/s40265-016-0633-9. [Abstract] [CrossRef] [Google Scholar]
423. Rebello RJ, Oing C, Knudsen KE, Loeb S, Johnson DC, Reiter RE, Gillessen S, Van der Kwast T, Bristow RG. Prostate cancer Nat Rev Dis Primers. 2021;7(1):9. 10.1038/s41572-020-00243-0. [Abstract] [CrossRef] [Google Scholar]
424. Visakorpi T, Hyytinen E, Koivisto P, Tanner M, Keinanen R, Palmberg C, Palotie A, Tammela T, Isola J, Kallioniemi OP. In vivo amplification of the androgen receptor gene and progression of human prostate cancer. Nat Genet. 1995;9(4):401–406. 10.1038/ng0495-401. [Abstract] [CrossRef] [Google Scholar]
425. Urbanucci A, Sahu B, Seppala J, Larjo A, Latonen LM, Waltering KK, Tammela TL, Vessella RL, Lahdesmaki H, Janne OA, et al. Overexpression of androgen receptor enhances the binding of the receptor to the chromatin in prostate cancer. Oncogene. 2012;31(17):2153–2163. 10.1038/onc.2011.401. [Abstract] [CrossRef] [Google Scholar]
426. Tomura A, Goto K, Morinaga H, Nomura M, Okabe T, Yanase T, Takayanagi R, Nawata H. The subnuclear three-dimensional image analysis of androgen receptor fused to green fluorescence protein. J Biol Chem. 2001;276(30):28395–28401. 10.1074/jbc.M101755200. [Abstract] [CrossRef] [Google Scholar]
427. Riegman PH, Vlietstra RJ, van der Korput JA, Brinkmann AO, Trapman J. The promoter of the prostate-specific antigen gene contains a functional androgen responsive element. Mol Endocrinol. 1991;5(12):1921–1930. 10.1210/mend-5-12-1921. [Abstract] [CrossRef] [Google Scholar]
428. Cleutjens KB, van Eekelen CC, van der Korput HA, Brinkmann AO, Trapman J. Two androgen response regions cooperate in steroid hormone regulated activity of the prostate-specific antigen promoter. J Biol Chem. 1996;271(11):6379–6388. 10.1074/jbc.271.11.6379. [Abstract] [CrossRef] [Google Scholar]
429. Huang W, Shostak Y, Tarr P, Sawyers C, Carey M. Cooperative assembly of androgen receptor into a nucleoprotein complex that regulates the prostate-specific antigen enhancer. J Biol Chem. 1999;274(36):25756–25768. 10.1074/jbc.274.36.25756. [Abstract] [CrossRef] [Google Scholar]
430. Stenman UH, Leinonen J, Alfthan H, Rannikko S, Tuhkanen K, Alfthan O. A complex between prostate-specific antigen and alpha 1-antichymotrypsin is the major form of prostate-specific antigen in serum of patients with prostatic cancer: assay of the complex improves clinical sensitivity for cancer. Cancer Res. 1991;51(1):222–226. [Abstract] [Google Scholar]
431. Taplin ME, Bubley GJ, Shuster TD, Frantz ME, Spooner AE, Ogata GK, Keer HN, Balk SP. Mutation of the androgen-receptor gene in metastatic androgen-independent prostate cancer. N Engl J Med. 1995;332(21):1393–1398. 10.1056/NEJM199505253322101. [Abstract] [CrossRef] [Google Scholar]
432. Massie CE, Lynch A, Ramos-Montoya A, Boren J, Stark R, Fazli L, Warren A, Scott H, Madhu B, Sharma N, et al. The androgen receptor fuels prostate cancer by regulating central metabolism and biosynthesis. EMBO J. 2011;30(13):2719–2733. 10.1038/emboj.2011.158. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
433. Mills IG. Maintaining and reprogramming genomic androgen receptor activity in prostate cancer. Nat Rev Cancer. 2014;14(3):187–198. 10.1038/nrc3678. [Abstract] [CrossRef] [Google Scholar]
434. Rice MA, Malhotra SV, Stoyanova T. Second-generation antiandrogens: from discovery to standard of care in castration resistant prostate cancer. Front Oncol. 2019;9:801. 10.3389/fonc.2019.00801. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
435. Zhao J, Ning S, Lou W, Yang JC, Armstrong CM, Lombard AP, D'Abronzo LS, Evans CP, Gao AC, Liu C. Cross-resistance among next-generation antiandrogen drugs through the AKR1C3/AR-V7 axis in advanced prostate cancer. Mol Cancer Ther. 2020;19(8):1708–1718. 10.1158/1535-7163.MCT-20-0015. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
436. Rathkopf DE, Smith MR, Ryan CJ, Berry WR, Shore ND, Liu G, Higano CS, Alumkal JJ, Hauke R, Tutrone RF, et al. Androgen receptor mutations in patients with castration-resistant prostate cancer treated with apalutamide. Ann Oncol. 2017;28(9):2264–2271. 10.1093/annonc/mdx283. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
437. Fenton MA, Shuster TD, Fertig AM, Taplin ME, Kolvenbag G, Bubley GJ, Balk SP. Functional characterization of mutant androgen receptors from androgen-independent prostate cancer. Clin Cancer Res. 1997;3(8):1383–1388. [Abstract] [Google Scholar]
438. Hara T, Miyazaki J, Araki H, Yamaoka M, Kanzaki N, Kusaka M, Miyamoto M. Novel mutations of androgen receptor: a possible mechanism of bicalutamide withdrawal syndrome. Cancer Res. 2003;63(1):149–153. [Abstract] [Google Scholar]
439. Bohl CE, Gao W, Miller DD, Bell CE, Dalton JT. Structural basis for antagonism and resistance of bicalutamide in prostate cancer. Proc Natl Acad Sci USA. 2005;102(17):6201–6206. 10.1073/pnas.0500381102. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
440. Urushibara M, Ishioka J, Hyochi N, Kihara K, Hara S, Singh P, Isaacs JT, Kageyama Y. Effects of steroidal and non-steroidal antiandrogens on wild-type and mutant androgen receptors. Prostate. 2007;67(8):799–807. 10.1002/pros.20542. [Abstract] [CrossRef] [Google Scholar]
441. Balbas MD, Evans MJ, Hosfield DJ, Wongvipat J, Arora VK, Watson PA, Chen Y, Greene GL, Shen Y, Sawyers CL. Overcoming mutation-based resistance to antiandrogens with rational drug design. Elife. 2013;2:e00499. 10.7554/eLife.00499. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
442. Joseph JD, Lu N, Qian J, Sensintaffar J, Shao G, Brigham D, Moon M, Maneval EC, Chen I, Darimont B, et al. A clinically relevant androgen receptor mutation confers resistance to second-generation antiandrogens enzalutamide and ARN-509. Cancer Discov. 2013;3(9):1020–1029. 10.1158/2159-8290.CD-13-0226. [Abstract] [CrossRef] [Google Scholar]
443. Korpal M, Korn JM, Gao X, Rakiec DP, Ruddy DA, Doshi S, Yuan J, Kovats SG, Kim S, Cooke VG, et al. An F876L mutation in androgen receptor confers genetic and phenotypic resistance to MDV3100 (enzalutamide) Cancer Discov. 2013;3(9):1030–1043. 10.1158/2159-8290.CD-13-0142. [Abstract] [CrossRef] [Google Scholar]
444. Moilanen AM, Riikonen R, Oksala R, Ravanti L, Aho E, Wohlfahrt G, Nykanen PS, Tormakangas OP, Palvimo JJ, Kallio PJ. Discovery of ODM-201, a new-generation androgen receptor inhibitor targeting resistance mechanisms to androgen signaling-directed prostate cancer therapies. Sci Rep. 2015;5:12007. 10.1038/srep12007. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
445. Huirne JA, Lambalk CB. Gonadotropin-releasing-hormone-receptor antagonists. Lancet. 2001;358(9295):1793–1803. 10.1016/S0140-6736(01)06797-6. [Abstract] [CrossRef] [Google Scholar]
446. Moussa M, Papatsoris A, Dellis A, Chakra MA, Fragkoulis C. Current and emerging gonadotropin-releasing hormone (GnRH) antagonists for the treatment of prostate cancer. Expert Opin Pharmacother. 2021 10.1080/14656566.2021.1948012. [Abstract] [CrossRef] [Google Scholar]
447. Engel JB, Schally AV. Drug Insight: clinical use of agonists and antagonists of luteinizing-hormone-releasing hormone. Nat Clin Pract Endocrinol Metab. 2007;3(2):157–167. 10.1038/ncpendmet0399. [Abstract] [CrossRef] [Google Scholar]
448. Olivennes F, Cunha-Filho JS, Fanchin R, Bouchard P, Frydman R. The use of GnRH antagonists in ovarian stimulation. Hum Reprod Update. 2002;8(3):279–290. 10.1093/humupd/8.3.279. [Abstract] [CrossRef] [Google Scholar]
449. Pokuri VK, Nourkeyhani H, Betsy B, Herbst L, Sikorski M, Spangenthal E, Fabiano A, George S. Strategies to circumvent testosterone surge and disease flare in advanced prostate cancer: emerging treatment paradigms. J Natl Compr Canc Netw. 2015;13(7):e49–55. 10.6004/jnccn.2015.0109. [Abstract] [CrossRef] [Google Scholar]
450. Schally AV, Arimura A, Baba Y, Nair RM, Matsuo H, Redding TW, Debeljuk L. Isolation and properties of the FSH and LH-releasing hormone. Biochem Biophys Res Commun. 1971;43(2):393–399. 10.1016/0006-291X(71)90766-2. [Abstract] [CrossRef] [Google Scholar]
451. Conn PM, Crowley WF., Jr Gonadotropin-releasing hormone and its analogues. N Engl J Med. 1991;324(2):93–103. 10.1056/NEJM199101103240205. [Abstract] [CrossRef] [Google Scholar]
452. Limonta P, Manea M. Gonadotropin-releasing hormone receptors as molecular therapeutic targets in prostate cancer: Current options and emerging strategies. Cancer Treat Rev. 2013;39(6):647–663. 10.1016/j.ctrv.2012.12.003. [Abstract] [CrossRef] [Google Scholar]
453. Molineaux CJ, Lasdun A, Michaud C, Orlowski M. Endopeptidase-24.15 is the primary enzyme that degrades luteinizing hormone releasing hormone both in vitro and in vivo. J Neurochem. 1988;51(2):624–633. 10.1111/j.1471-4159.1988.tb01084.x. [Abstract] [CrossRef] [Google Scholar]
454. Redding TW, Kastin AJ, Gonzales-Barcena D, Coy DH, Coy EJ, Schalch DS, Schally AV. The half-life, metabolism and excretion of tritiated luteinizing hormone-releasing hormone (LH-RH) in man. J Clin Endocrinol Metab. 1973;37(4):626–631. 10.1210/jcem-37-4-626. [Abstract] [CrossRef] [Google Scholar]
455. Sealfon SC, Weinstein H, Millar RP. Molecular mechanisms of ligand interaction with the gonadotropin-releasing hormone receptor. Endocr Rev. 1997;18(2):180–205. 10.1210/edrv.18.2.0295. [Abstract] [CrossRef] [Google Scholar]
456. Shore N, Cookson MS, Gittelman MC. Long-term efficacy and tolerability of once-yearly histrelin acetate subcutaneous implant in patients with advanced prostate cancer. BJU Int. 2012;109(2):226–232. 10.1111/j.1464-410X.2011.10370.x. [Abstract] [CrossRef] [Google Scholar]
457. Tunn UW. A 6-month depot formulation of leuprolide acetate is safe and effective in daily clinical practice: a non-interventional prospective study in 1273 patients. BMC Urol. 2011;11:15. 10.1186/1471-2490-11-15. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
458. Yri OE, Bjoro T, Fossa SD. Failure to achieve castration levels in patients using leuprolide acetate in locally advanced prostate cancer. Eur Urol. 2006;49(1):54–58. 10.1016/j.eururo.2005.09.009. [Abstract] [CrossRef] [Google Scholar]
459. Karten MJ, Rivier JE. Gonadotropin-releasing hormone analog design Structure-function studies toward the development of agonists and antagonists: rationale and perspective. Endocr Rev. 1986;7(1):44–66. 10.1210/edrv-7-1-44. [Abstract] [CrossRef] [Google Scholar]
460. Herbst KL. Gonadotropin-releasing hormone antagonists. Curr Opin Pharmacol. 2003;3(6):660–666. 10.1016/j.coph.2003.06.009. [Abstract] [CrossRef] [Google Scholar]
461. Tomera K, Gleason D, Gittelman M, Moseley W, Zinner N, Murdoch M, Menon M, Campion M, Garnick MB. The gonadotropin-releasing hormone antagonist abarelix depot versus luteinizing hormone releasing hormone agonists leuprolide or goserelin: initial results of endocrinological and biochemical efficacies in patients with prostate cancer. J Urol. 2001;165(5):1585–1589. 10.1016/S0022-5347(05)66353-7. [Abstract] [CrossRef] [Google Scholar]
462. McLeod D, Zinner N, Tomera K, Gleason D, Fotheringham N, Campion M, Garnick MB, Abarelix Study G A phase 3, multicenter, open-label, randomized study of abarelix versus leuprolide acetate in men with prostate cancer. Urology. 2001;58(5):756–761. 10.1016/S0090-4295(01)01342-5. [Abstract] [CrossRef] [Google Scholar]
463. Koch M, Steidle C, Brosman S, Centeno A, Gaylis F, Campion M, Garnick MB, Abarelix Study G An open-label study of abarelix in men with symptomatic prostate cancer at risk of treatment with LHRH agonists. Urology. 2003;62(5):877–882. 10.1016/S0090-4295(03)00656-3. [Abstract] [CrossRef] [Google Scholar]
464. Mongiat-Artus P, Teillac P. Abarelix: the first gonadotrophin-releasing hormone antagonist for the treatment of prostate cancer. Expert Opin Pharmacother. 2004;5(10):2171–2179. 10.1517/14656566.5.10.2171. [Abstract] [CrossRef] [Google Scholar]
465. Schroder F, Crawford ED, Axcrona K, Payne H, Keane TE. Androgen deprivation therapy: past, present and future. BJU Int. 2012;109(Suppl 6):1–12. 10.1111/j.1464-410X.2012.11215.x. [Abstract] [CrossRef] [Google Scholar]
466. Huhtaniemi I, White R, McArdle CA, Persson BE. Will GnRH antagonists improve prostate cancer treatment? Trends Endocrinol Metab. 2009;20(1):43–50. 10.1016/j.tem.2008.09.003. [Abstract] [CrossRef] [Google Scholar]
467. Carter NJ, Keam SJ. Degarelix: a review of its use in patients with prostate cancer. Drugs. 2014;74(6):699–712. 10.1007/s40265-014-0211-y. [Abstract] [CrossRef] [Google Scholar]
468. Klotz L, Boccon-Gibod L, Shore ND, Andreou C, Persson BE, Cantor P, Jensen JK, Olesen TK, Schroder FH. The efficacy and safety of degarelix: a 12-month, comparative, randomized, open-label, parallel-group phase III study in patients with prostate cancer. BJU Int. 2008;102(11):1531–1538. 10.1111/j.1464-410X.2008.08183.x. [Abstract] [CrossRef] [Google Scholar]
469. Crawford ED, Tombal B, Miller K, Boccon-Gibod L, Schroder F, Shore N, Moul JW, Jensen JK, Olesen TK, Persson BE. A phase III extension trial with a 1-arm crossover from leuprolide to degarelix: comparison of gonadotropin-releasing hormone agonist and antagonist effect on prostate cancer. J Urol. 2011;186(3):889–897. 10.1016/j.juro.2011.04.083. [Abstract] [CrossRef] [Google Scholar]
470. Miwa K, Hitaka T, Imada T, Sasaki S, Yoshimatsu M, Kusaka M, Tanaka A, Nakata D, Furuya S, Endo S, et al. Discovery of 1-{4-[1-(2,6-difluorobenzyl)-5-[(dimethylamino)methyl]-3-(6-methoxypyridazin-3-yl )-2,4-dioxo-1,2,3,4-tetrahydrothieno[2,3-d]pyrimidin-6-yl]phenyl}-3-methoxyurea (TAK-385) as a potent, orally active, non-peptide antagonist of the human gonadotropin-releasing hormone receptor. J Med Chem. 2011;54(14):4998–5012. 10.1021/jm200216q. [Abstract] [CrossRef] [Google Scholar]
471. Shore ND, Saad F, Cookson MS, George DJ, Saltzstein DR, Tutrone R, Akaza H, Bossi A, van Veenhuyzen DF, Selby B, et al. Oral relugolix for androgen-deprivation therapy in advanced prostate cancer. N Engl J Med. 2020;382(23):2187–2196. 10.1056/NEJMoa2004325. [Abstract] [CrossRef] [Google Scholar]
472. Paschalis A, de Bono JS. Prostate cancer 2020: "the times they are a'changing". Cancer Cell. 2020;38(1):25–27. 10.1016/j.ccell.2020.06.008. [Abstract] [CrossRef] [Google Scholar]
473. Cheetham P, Petrylak DP. Tubulin-targeted agents including docetaxel and cabazitaxel. Cancer J. 2013;19(1):59–65. 10.1097/PPO.0b013e3182828d38. [Abstract] [CrossRef] [Google Scholar]
474. Galsky MD, Dritselis A, Kirkpatrick P, Oh WK. Cabazitaxel. Nat Rev Drug Discov. 2010;9(9):677–678. 10.1038/nrd3254. [Abstract] [CrossRef] [Google Scholar]
475. Villanueva C, Bazan F, Kim S, Demarchi M, Chaigneau L, Thiery-Vuillemin A, Nguyen T, Cals L, Dobi E, Pivot X. Cabazitaxel: a novel microtubule inhibitor. Drugs. 2011;71(10):1251–1258. 10.2165/11591390-000000000-00000. [Abstract] [CrossRef] [Google Scholar]
476. Duran GE, Derdau V, Weitz D, Philippe N, Blankenstein J, Atzrodt J, Semiond D, Gianolio DA, Mace S, Sikic BI. Cabazitaxel is more active than first-generation taxanes in ABCB1(+) cell lines due to its reduced affinity for P-glycoprotein. Cancer Chemother Pharmacol. 2018;81(6):1095–1103. 10.1007/s00280-018-3572-1. [Abstract] [CrossRef] [Google Scholar]
477. de Bono JS, Oudard S, Ozguroglu M, Hansen S, Machiels JP, Kocak I, Gravis G, Bodrogi I, Mackenzie MJ, Shen L, et al. Prednisone plus cabazitaxel or mitoxantrone for metastatic castration-resistant prostate cancer progressing after docetaxel treatment: a randomised open-label trial. Lancet. 2010;376(9747):1147–1154. 10.1016/S0140-6736(10)61389-X. [Abstract] [CrossRef] [Google Scholar]
478. Pezaro CJ, Omlin AG, Altavilla A, Lorente D, Ferraldeschi R, Bianchini D, Dearnaley D, Parker C, de Bono JS, Attard G. Activity of cabazitaxel in castration-resistant prostate cancer progressing after docetaxel and next-generation endocrine agents. Eur Urol. 2014;66(3):459–465. 10.1016/j.eururo.2013.11.044. [Abstract] [CrossRef] [Google Scholar]
479. Paller CJ, Antonarakis ES. Cabazitaxel: a novel second-line treatment for metastatic castration-resistant prostate cancer. Drug Des Devel Ther. 2011;5:117–124. [Europe PMC free article] [Abstract] [Google Scholar]
480. Hoy SM. Abiraterone acetate: a review of its use in patients with metastatic castration-resistant prostate cancer. Drugs. 2013;73(18):2077–2091. 10.1007/s40265-013-0150-z. [Abstract] [CrossRef] [Google Scholar]
481. Attard G, Reid AH, Olmos D, de Bono JS. Antitumor activity with CYP17 blockade indicates that castration-resistant prostate cancer frequently remains hormone driven. Cancer Res. 2009;69(12):4937–4940. 10.1158/0008-5472.CAN-08-4531. [Abstract] [CrossRef] [Google Scholar]
482. DeVore NM, Scott EE. Structures of cytochrome P450 17A1 with prostate cancer drugs abiraterone and TOK-001. Nature. 2012;482(7383):116–119. 10.1038/nature10743. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
483. Scott LJ. Abiraterone acetate: a review in metastatic castration-resistant prostrate cancer. Drugs. 2017;77(14):1565–1576. 10.1007/s40265-017-0799-9. [Abstract] [CrossRef] [Google Scholar]
484. de Bono JS, Logothetis CJ, Molina A, Fizazi K, North S, Chu L, Chi KN, Jones RJ, Goodman OB, Jr, Saad F, et al. Abiraterone and increased survival in metastatic prostate cancer. N Engl J Med. 2011;364(21):1995–2005. 10.1056/NEJMoa1014618. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
485. Ryan CJ, Smith MR, de Bono JS, Molina A, Logothetis CJ, de Souza P, Fizazi K, Mainwaring P, Piulats JM, Ng S, et al. Abiraterone in metastatic prostate cancer without previous chemotherapy. N Engl J Med. 2013;368(2):138–148. 10.1056/NEJMoa1209096. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
486. Fizazi K, Scher HI, Molina A, Logothetis CJ, Chi KN, Jones RJ, Staffurth JN, North S, Vogelzang NJ, Saad F, et al. Abiraterone acetate for treatment of metastatic castration-resistant prostate cancer: final overall survival analysis of the COU-AA-301 randomised, double-blind, placebo-controlled phase 3 study. Lancet Oncol. 2012;13(10):983–992. 10.1016/S1470-2045(12)70379-0. [Abstract] [CrossRef] [Google Scholar]
487. Ryan CJ, Smith MR, Fizazi K, Saad F, Mulders PF, Sternberg CN, Miller K, Logothetis CJ, Shore ND, Small EJ, et al. Abiraterone acetate plus prednisone versus placebo plus prednisone in chemotherapy-naive men with metastatic castration-resistant prostate cancer (COU-AA-302): final overall survival analysis of a randomised, double-blind, placebo-controlled phase 3 study. Lancet Oncol. 2015;16(2):152–160. 10.1016/S1470-2045(14)71205-7. [Abstract] [CrossRef] [Google Scholar]
488. Gourd E. EMA guidance on radium-223 dichloride in prostate cancer. Lancet Oncol. 2018;19(4):e190. 10.1016/S1470-2045(18)30216-X. [Abstract] [CrossRef] [Google Scholar]
489. Coleman R. Treatment of metastatic bone disease and the emerging role of radium-223. Semin Nucl Med. 2016;46(2):99–104. 10.1053/j.semnuclmed.2015.10.012. [Abstract] [CrossRef] [Google Scholar]
490. Shirley M, McCormack PL. Radium-223 dichloride: a review of its use in patients with castration-resistant prostate cancer with symptomatic bone metastases. Drugs. 2014;74(5):579–586. 10.1007/s40265-014-0198-4. [Abstract] [CrossRef] [Google Scholar]
491. Baidoo KE, Yong K, Brechbiel MW. Molecular pathways: targeted alpha-particle radiation therapy. Clin Cancer Res. 2013;19(3):530–537. 10.1158/1078-0432.CCR-12-0298. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
492. Diaz-Montero CM, Rini BI, Finke JH. The immunology of renal cell carcinoma. Nat Rev Nephrol. 2020;16(12):721–735. 10.1038/s41581-020-0316-3. [Abstract] [CrossRef] [Google Scholar]
493. Morgan DA, Ruscetti FW, Gallo R. Selective in vitro growth of T lymphocytes from normal human bone marrows. Science. 1976;193(4257):1007–1008. 10.1126/science.181845. [Abstract] [CrossRef] [Google Scholar]
494. Taniguchi T, Matsui H, Fujita T, Takaoka C, Kashima N, Yoshimoto R, Hamuro J. Structure and expression of a cloned cDNA for human interleukin-2. Nature. 1983;302(5906):305–310. 10.1038/302305a0. [Abstract] [CrossRef] [Google Scholar]
495. Paliard X, de Waal MR, Yssel H, Blanchard D, Chretien I, Abrams J, de Vries J, Spits H. Simultaneous production of IL-2, IL-4, and IFN-gamma by activated human CD4+ and CD8+ T cell clones. J Immunol. 1988;141(3):849–855. [Abstract] [Google Scholar]
496. Spolski R, Li P, Leonard WJ. Biology and regulation of IL-2: from molecular mechanisms to human therapy. Nat Rev Immunol. 2018;18(10):648–659. 10.1038/s41577-018-0046-y. [Abstract] [CrossRef] [Google Scholar]
497. Whittington R, Faulds D. Interleukin-2. A review of its pharmacological properties and therapeutic use in patients with cancer. Drugs. 1993;46(3):446–514. 10.2165/00003495-199346030-00009. [Abstract] [CrossRef] [Google Scholar]
498. Fyfe G, Fisher RI, Rosenberg SA, Sznol M, Parkinson DR, Louie AC. Results of treatment of 255 patients with metastatic renal cell carcinoma who received high-dose recombinant interleukin-2 therapy. J Clin Oncol. 1995;13(3):688–696. 10.1200/JCO.1995.13.3.688. [Abstract] [CrossRef] [Google Scholar]
499. McDermott DF, Cheng SC, Signoretti S, Margolin KA, Clark JI, Sosman JA, Dutcher JP, Logan TF, Curti BD, Ernstoff MS, et al. The high-dose aldesleukin "select" trial: a trial to prospectively validate predictive models of response to treatment in patients with metastatic renal cell carcinoma. Clin Cancer Res. 2015;21(3):561–568. 10.1158/1078-0432.CCR-14-1520. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
500. Wilhelm SM, Carter C, Tang L, Wilkie D, McNabola A, Rong H, Chen C, Zhang X, Vincent P, McHugh M, et al. BAY 43–9006 exhibits broad spectrum oral antitumor activity and targets the RAF/MEK/ERK pathway and receptor tyrosine kinases involved in tumor progression and angiogenesis. Cancer Res. 2004;64(19):7099–7109. 10.1158/0008-5472.CAN-04-1443. [Abstract] [CrossRef] [Google Scholar]
501. Wilhelm S, Carter C, Lynch M, Lowinger T, Dumas J, Smith RA, Schwartz B, Simantov R, Kelley S. Discovery and development of sorafenib: a multikinase inhibitor for treating cancer. Nat Rev Drug Discov. 2006;5(10):835–844. 10.1038/nrd2130. [Abstract] [CrossRef] [Google Scholar]
502. Wan PT, Garnett MJ, Roe SM, Lee S, Niculescu-Duvaz D, Good VM, Jones CM, Marshall CJ, Springer CJ, Barford D, et al. Mechanism of activation of the RAF-ERK signaling pathway by oncogenic mutations of B-RAF. Cell. 2004;116(6):855–867. 10.1016/S0092-8674(04)00215-6. [Abstract] [CrossRef] [Google Scholar]
503. He Y, Luo Y, Huang L, Zhang D, Wang X, Ji J, Liang S. New frontiers against sorafenib resistance in renal cell carcinoma: from molecular mechanisms to predictive biomarkers. Pharmacol Res. 2021;170:105732. 10.1016/j.phrs.2021.105732. [Abstract] [CrossRef] [Google Scholar]
504. Bruix J, Qin S, Merle P, Granito A, Huang YH, Bodoky G, Pracht M, Yokosuka O, Rosmorduc O, Breder V, et al. Regorafenib for patients with hepatocellular carcinoma who progressed on sorafenib treatment (RESORCE): a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet. 2017;389(10064):56–66. 10.1016/S0140-6736(16)32453-9. [Abstract] [CrossRef] [Google Scholar]
505. Yasuda D, Ohe T, Takahashi K, Imamura R, Kojima H, Okabe T, Ichimura Y, Komatsu M, Yamamoto M, Nagano T, et al. Inhibitors of the protein-protein interaction between phosphorylated p62 and Keap1 attenuate chemoresistance in a human hepatocellular carcinoma cell line. Free Radic Res. 2020;54(11–12):859–871. 10.1080/10715762.2020.1732955. [Abstract] [CrossRef] [Google Scholar]
506. Eichelberg C, Vervenne WL, De Santis M, Fischer von Weikersthal L, Goebell PJ, Lerchenmuller C, Zimmermann U, Bos MM, Freier W, Schirrmacher-Memmel S, et al. SWITCH: a randomised, sequential, open-label study to evaluate the efficacy and safety of sorafenib-sunitinib versus sunitinib-sorafenib in the treatment of metastatic renal cell cancer. Eur Urol. 2015;68(5):837–847. 10.1016/j.eururo.2015.04.017. [Abstract] [CrossRef] [Google Scholar]
507. Harris PA, Boloor A, Cheung M, Kumar R, Crosby RM, Davis-Ward RG, Epperly AH, Hinkle KW, Hunter RN, 3rd, Johnson JH, et al. Discovery of 5-[[4-[(2,3-dimethyl-2H-indazol-6-yl)methylamino]-2-pyrimidinyl]amino]-2-methyl-b enzenesulfonamide (Pazopanib), a novel and potent vascular endothelial growth factor receptor inhibitor. J Med Chem. 2008;51(15):4632–4640. 10.1021/jm800566m. [Abstract] [CrossRef] [Google Scholar]
508. Kumar R, Knick VB, Rudolph SK, Johnson JH, Crosby RM, Crouthamel MC, Hopper TM, Miller CG, Harrington LE, Onori JA, et al. Pharmacokinetic-pharmacodynamic correlation from mouse to human with pazopanib, a multikinase angiogenesis inhibitor with potent antitumor and antiangiogenic activity. Mol Cancer Ther. 2007;6(7):2012–2021. 10.1158/1535-7163.MCT-07-0193. [Abstract] [CrossRef] [Google Scholar]
509. Keisner SV, Shah SR. Pazopanib: the newest tyrosine kinase inhibitor for the treatment of advanced or metastatic renal cell carcinoma. Drugs. 2011;71(4):443–454. [Abstract] [Google Scholar]
510. van Geel RM, Beijnen JH, Schellens JH. Concise drug review: pazopanib and axitinib. Oncologist. 2012;17(8):1081–1089. 10.1634/theoncologist.2012-0055. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
511. Hainsworth JD, Rubin MS, Arrowsmith ER, Khatcheressian J, Crane EJ, Franco LA. Pazopanib as second-line treatment after sunitinib or bevacizumab in patients with advanced renal cell carcinoma: a Sarah Cannon Oncology Research Consortium Phase II Trial. Clin Genitourin Cancer. 2013;11(3):270–275. 10.1016/j.clgc.2013.04.006. [Abstract] [CrossRef] [Google Scholar]
512. Motzer RJ, Hutson TE, Cella D, Reeves J, Hawkins R, Guo J, Nathan P, Staehler M, de Souza P, Merchan JR, et al. Pazopanib versus sunitinib in metastatic renal-cell carcinoma. N Engl J Med. 2013;369(8):722–731. 10.1056/NEJMoa1303989. [Abstract] [CrossRef] [Google Scholar]
513. Keating GM. Axitinib: a review in advanced renal cell carcinoma. Drugs. 2015;75(16):1903–1913. 10.1007/s40265-015-0483-x. [Abstract] [CrossRef] [Google Scholar]
514. Kim ES. Tivozanib: first global approval. Drugs. 2017;77(17):1917–1923. 10.1007/s40265-017-0825-y. [Abstract] [CrossRef] [Google Scholar]
515. Rugo HS, Herbst RS, Liu G, Park JW, Kies MS, Steinfeldt HM, Pithavala YK, Reich SD, Freddo JL, Wilding G. Phase I trial of the oral antiangiogenesis agent AG-013736 in patients with advanced solid tumors: pharmacokinetic and clinical results. J Clin Oncol. 2005;23(24):5474–5483. 10.1200/JCO.2005.04.192. [Abstract] [CrossRef] [Google Scholar]
516. Rini BI, Wilding G, Hudes G, Stadler WM, Kim S, Tarazi J, Rosbrook B, Trask PC, Wood L, Dutcher JP. Phase II study of axitinib in sorafenib-refractory metastatic renal cell carcinoma. J Clin Oncol. 2009;27(27):4462–4468. 10.1200/JCO.2008.21.7034. [Abstract] [CrossRef] [Google Scholar]
517. Hutson TE, Lesovoy V, Al-Shukri S, Stus VP, Lipatov ON, Bair AH, Rosbrook B, Chen C, Kim S, Vogelzang NJ. Axitinib versus sorafenib as first-line therapy in patients with metastatic renal-cell carcinoma: a randomised open-label phase 3 trial. Lancet Oncol. 2013;14(13):1287–1294. 10.1016/S1470-2045(13)70465-0. [Abstract] [CrossRef] [Google Scholar]
518. Rini BI, Escudier B, Tomczak P, Kaprin A, Szczylik C, Hutson TE, Michaelson MD, Gorbunova VA, Gore ME, Rusakov IG, et al. Comparative effectiveness of axitinib versus sorafenib in advanced renal cell carcinoma (AXIS): a randomised phase 3 trial. Lancet. 2011;378(9807):1931–1939. 10.1016/S0140-6736(11)61613-9. [Abstract] [CrossRef] [Google Scholar]
519. Motzer RJ, Escudier B, Tomczak P, Hutson TE, Michaelson MD, Negrier S, Oudard S, Gore ME, Tarazi J, Hariharan S, et al. Axitinib versus sorafenib as second-line treatment for advanced renal cell carcinoma: overall survival analysis and updated results from a randomised phase 3 trial. Lancet Oncol. 2013;14(6):552–562. 10.1016/S1470-2045(13)70093-7. [Abstract] [CrossRef] [Google Scholar]
520. Motzer RJ, Penkov K, Haanen J, Rini B, Albiges L, Campbell MT, Venugopal B, Kollmannsberger C, Negrier S, Uemura M, et al. Avelumab plus axitinib versus sunitinib for advanced renal-cell carcinoma. N Engl J Med. 2019;380(12):1103–1115. 10.1056/NEJMoa1816047. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
521. Rini BI, Plimack ER, Stus V, Gafanov R, Hawkins R, Nosov D, Pouliot F, Alekseev B, Soulieres D, Melichar B, et al. Pembrolizumab plus axitinib versus sunitinib for advanced renal-cell carcinoma. N Engl J Med. 2019;380(12):1116–1127. 10.1056/NEJMoa1816714. [Abstract] [CrossRef] [Google Scholar]
522. Lee K, Jeong KW, Lee Y, Song JY, Kim MS, Lee GS, Kim Y. Pharmacophore modeling and virtual screening studies for new VEGFR-2 kinase inhibitors. Eur J Med Chem. 2010;45(11):5420–5427. 10.1016/j.ejmech.2010.09.002. [Abstract] [CrossRef] [Google Scholar]
523. Nakamura K, Taguchi E, Miura T, Yamamoto A, Takahashi K, Bichat F, Guilbaud N, Hasegawa K, Kubo K, Fujiwara Y, et al. KRN951, a highly potent inhibitor of vascular endothelial growth factor receptor tyrosine kinases, has antitumor activities and affects functional vascular properties. Cancer Res. 2006;66(18):9134–9142. 10.1158/0008-5472.CAN-05-4290. [Abstract] [CrossRef] [Google Scholar]
524. Rini BI, Pal SK, Escudier BJ, Atkins MB, Hutson TE, Porta C, Verzoni E, Needle MN, McDermott DF. Tivozanib versus sorafenib in patients with advanced renal cell carcinoma (TIVO-3): a phase 3, multicentre, randomised, controlled, open-label study. Lancet Oncol. 2020;21(1):95–104. 10.1016/S1470-2045(19)30735-1. [Abstract] [CrossRef] [Google Scholar]
525. Liu GY, Sabatini DM. mTOR at the nexus of nutrition, growth, ageing and disease. Nat Rev Mol Cell Biol. 2020;21(4):183–203. 10.1038/s41580-019-0199-y. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
526. Kim DH, Sarbassov DD, Ali SM, Latek RR, Guntur KV, Erdjument-Bromage H, Tempst P, Sabatini DM. GbetaL, a positive regulator of the rapamycin-sensitive pathway required for the nutrient-sensitive interaction between raptor and mTOR. Mol Cell. 2003;11(4):895–904. 10.1016/S1097-2765(03)00114-X. [Abstract] [CrossRef] [Google Scholar]
527. Kim DH, Sarbassov DD, Ali SM, King JE, Latek RR, Erdjument-Bromage H, Tempst P, Sabatini DM. mTOR interacts with raptor to form a nutrient-sensitive complex that signals to the cell growth machinery. Cell. 2002;110(2):163–175. 10.1016/S0092-8674(02)00808-5. [Abstract] [CrossRef] [Google Scholar]
528. Sarbassov DD, Ali SM, Kim DH, Guertin DA, Latek RR, Erdjument-Bromage H, Tempst P, Sabatini DM. Rictor, a novel binding partner of mTOR, defines a rapamycin-insensitive and raptor-independent pathway that regulates the cytoskeleton. Curr Biol. 2004;14(14):1296–1302. 10.1016/j.cub.2004.06.054. [Abstract] [CrossRef] [Google Scholar]
529. Jacinto E, Loewith R, Schmidt A, Lin S, Ruegg MA, Hall A, Hall MN. Mammalian TOR complex 2 controls the actin cytoskeleton and is rapamycin insensitive. Nat Cell Biol. 2004;6(11):1122–1128. 10.1038/ncb1183. [Abstract] [CrossRef] [Google Scholar]
530. Vezina C, Kudelski A, Sehgal SN. Rapamycin (AY-22,989), a new antifungal antibiotic. I. Taxonomy of the producing streptomycete and isolation of the active principle. J Antibiot. 1975;28(10):721–726. 10.7164/antibiotics.28.721. [Abstract] [CrossRef] [Google Scholar]
531. Rini BI. Temsirolimus, an inhibitor of mammalian target of rapamycin. Clin Cancer Res. 2008;14(5):1286–1290. 10.1158/1078-0432.CCR-07-4719. [Abstract] [CrossRef] [Google Scholar]
532. Houghton PJ. Everolimus. Clin Cancer Res. 2010;16(5):1368–1372. 10.1158/1078-0432.CCR-09-1314. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
533. Bukowski RM. Temsirolimus: a safety and efficacy review. Expert Opin Drug Saf. 2012;11(5):861–879. 10.1517/14740338.2012.713344. [Abstract] [CrossRef] [Google Scholar]
534. Boni JP, Hug B, Leister C, Sonnichsen D. Intravenous temsirolimus in cancer patients: clinical pharmacology and dosing considerations. Semin Oncol. 2009;36(Suppl 3):S18–25. 10.1053/j.seminoncol.2009.10.009. [Abstract] [CrossRef] [Google Scholar]
535. Hudes G, Carducci M, Tomczak P, Dutcher J, Figlin R, Kapoor A, Staroslawska E, Sosman J, McDermott D, Bodrogi I, et al. Temsirolimus, interferon alfa, or both for advanced renal-cell carcinoma. N Engl J Med. 2007;356(22):2271–2281. 10.1056/NEJMoa066838. [Abstract] [CrossRef] [Google Scholar]
536. Motzer RJ, Escudier B, Oudard S, Hutson TE, Porta C, Bracarda S, Grunwald V, Thompson JA, Figlin RA, Hollaender N, et al. Efficacy of everolimus in advanced renal cell carcinoma: a double-blind, randomised, placebo-controlled phase III trial. Lancet. 2008;372(9637):449–456. 10.1016/S0140-6736(08)61039-9. [Abstract] [CrossRef] [Google Scholar]
537. Iacovelli R, Santoni M, Verzoni E, Grassi P, Testa I, de Braud F, Cascinu S, Procopio G. Everolimus and temsirolimus are not the same second-line in metastatic renal cell carcinoma. A systematic review and meta-analysis of literature data. Clin Genitourin Cancer. 2015;13(2):137–141. 10.1016/j.clgc.2014.07.006. [Abstract] [CrossRef] [Google Scholar]
538. Wong MK, Yang H, Signorovitch JE, Wang X, Liu Z, Liu NS, Qi CZ, George DJ. Comparative outcomes of everolimus, temsirolimus and sorafenib as second targeted therapies for metastatic renal cell carcinoma: a US medical record review. Curr Med Res Opin. 2014;30(4):537–545. 10.1185/03007995.2013.871243. [Abstract] [CrossRef] [Google Scholar]
539. Garcia JA, Rini BI. Recent progress in the management of advanced renal cell carcinoma. CA Cancer J Clin. 2007;57(2):112–125. 10.3322/canjclin.57.2.112. [Abstract] [CrossRef] [Google Scholar]
540. Bukowski RM, Yasothan U, Kirkpatrick P. Pazopanib. Nat Rev Drug Discov. 2010;9(1):17–18. 10.1038/nrd3073. [Abstract] [CrossRef] [Google Scholar]
541. Motzer R, Alekseev B, Rha SY, Porta C, Eto M, Powles T, Grunwald V, Hutson TE, Kopyltsov E, Mendez-Vidal MJ, et al. Lenvatinib plus pembrolizumab or everolimus for advanced renal cell carcinoma. N Engl J Med. 2021;384(14):1289–1300. 10.1056/NEJMoa2035716. [Abstract] [CrossRef] [Google Scholar]
542. Choueiri TK, Powles T, Burotto M, Escudier B, Bourlon MT, Zurawski B, Oyervides Juarez VM, Hsieh JJ, Basso U, Shah AY, et al. Nivolumab plus cabozantinib versus sunitinib for advanced renal-cell carcinoma. N Engl J Med. 2021;384(9):829–841. 10.1056/NEJMoa2026982. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
543. Chehrazi-Raffle A, Dorff TB, Pal SK, Lyou Y. Wnt/beta-catenin signaling and immunotherapy resistance: lessons for the treatment of urothelial carcinoma. Cancers. 2021 10.3390/cancers13040889. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
544. Bellmunt J, Mullane SA, Werner L, Fay AP, Callea M, Leow JJ, Taplin ME, Choueiri TK, Hodi FS, Freeman GJ, et al. Association of PD-L1 expression on tumor-infiltrating mononuclear cells and overall survival in patients with urothelial carcinoma. Ann Oncol. 2015;26(4):812–817. 10.1093/annonc/mdv009. [Abstract] [CrossRef] [Google Scholar]
545. Doroshow DB, Bhalla S, Beasley MB, Sholl LM, Kerr KM, Gnjatic S, Wistuba II, Rimm DL, Tsao MS, Hirsch FR. PD-L1 as a biomarker of response to immune-checkpoint inhibitors. Nat Rev Clin Oncol. 2021;18(6):345–362. 10.1038/s41571-021-00473-5. [Abstract] [CrossRef] [Google Scholar]
546. Deng R, Bumbaca D, Pastuskovas CV, Boswell CA, West D, Cowan KJ, Chiu H, McBride J, Johnson C, Xin Y, et al. Preclinical pharmacokinetics, pharmacodynamics, tissue distribution, and tumor penetration of anti-PD-L1 monoclonal antibody, an immune checkpoint inhibitor. MAbs. 2016;8(3):593–603. 10.1080/19420862.2015.1136043. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
547. Lee HT, Lee JY, Lim H, Lee SH, Moon YJ, Pyo HJ, Ryu SE, Shin W, Heo YS. Molecular mechanism of PD-1/PD-L1 blockade via anti-PD-L1 antibodies atezolizumab and durvalumab. Sci Rep. 2017;7(1):5532. 10.1038/s41598-017-06002-8. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
548. Markham A. Atezolizumab: first global approval. Drugs. 2016;76(12):1227–1232. 10.1007/s40265-016-0618-8. [Abstract] [CrossRef] [Google Scholar]
549. Syed YY. Durvalumab: first global approval. Drugs. 2017;77(12):1369–1376. 10.1007/s40265-017-0782-5. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
550. Powles T, Eder JP, Fine GD, Braiteh FS, Loriot Y, Cruz C, Bellmunt J, Burris HA, Petrylak DP, Teng SL, et al. MPDL3280A (anti-PD-L1) treatment leads to clinical activity in metastatic bladder cancer. Nature. 2014;515(7528):558–562. 10.1038/nature13904. [Abstract] [CrossRef] [Google Scholar]
551. Rosenberg JE, Hoffman-Censits J, Powles T, van der Heijden MS, Balar AV, Necchi A, Dawson N, O'Donnell PH, Balmanoukian A, Loriot Y, et al. Atezolizumab in patients with locally advanced and metastatic urothelial carcinoma who have progressed following treatment with platinum-based chemotherapy: a single-arm, multicentre, phase 2 trial. Lancet. 2016;387(10031):1909–1920. 10.1016/S0140-6736(16)00561-4. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
552. Balar AV, Galsky MD, Rosenberg JE, Powles T, Petrylak DP, Bellmunt J, Loriot Y, Necchi A, Hoffman-Censits J, Perez-Gracia JL, et al. Atezolizumab as first-line treatment in cisplatin-ineligible patients with locally advanced and metastatic urothelial carcinoma: a single-arm, multicentre, phase 2 trial. Lancet. 2017;389(10064):67–76. 10.1016/S0140-6736(16)32455-2. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
553. Powles T, Duran I, van der Heijden MS, Loriot Y, Vogelzang NJ, De Giorgi U, Oudard S, Retz MM, Castellano D, Bamias A, et al. Atezolizumab versus chemotherapy in patients with platinum-treated locally advanced or metastatic urothelial carcinoma (IMvigor211): a multicentre, open-label, phase 3 randomised controlled trial. Lancet. 2018;391(10122):748–757. 10.1016/S0140-6736(17)33297-X. [Abstract] [CrossRef] [Google Scholar]
554. Powles T, O'Donnell PH, Massard C, Arkenau H-T, Friedlander TW, Hoimes C, Lee J-L, Ong M, Sridhar SS, Vogelzang NJ, et al. Updated efficacy and tolerability of durvalumab in locally advanced or metastatic urothelial carcinoma. J Clin Oncol. 2017;35(6_suppl):286–286. 10.1200/JCO.2017.35.6_suppl.286. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
555. Hahn NM, Powles T, Massard C, Arkenau H-T, Friedlander TW, Hoimes CJ, Lee J-L, Ong M, Sridhar SS, Vogelzang NJ, et al. Updated efficacy and tolerability of durvalumab in locally advanced or metastatic urothelial carcinoma (UC) J Clin Oncol. 2017;35(15_suppl):4525–4525. 10.1200/JCO.2017.35.15_suppl.4525. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
556. Powles T, O'Donnell PH, Massard C, Arkenau HT, Friedlander TW, Hoimes CJ, Lee JL, Ong M, Sridhar SS, Vogelzang NJ, et al. Efficacy and safety of durvalumab in locally advanced or metastatic urothelial carcinoma: updated results from a phase 1/2 open-label study. JAMA Oncol. 2017;3(9):e172411. 10.1001/jamaoncol.2017.2411. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
557. Patel MR, Ellerton J, Infante JR, Agrawal M, Gordon M, Aljumaily R, Britten CD, Dirix L, Lee KW, Taylor M, et al. Avelumab in metastatic urothelial carcinoma after platinum failure (JAVELIN Solid Tumor): pooled results from two expansion cohorts of an open-label, phase 1 trial. Lancet Oncol. 2018;19(1):51–64. 10.1016/S1470-2045(17)30900-2. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
558. Powles T, Park SH, Voog E, Caserta C, Valderrama BP, Gurney H, Kalofonos H, Radulovic S, Demey W, Ullen A, et al. Avelumab maintenance therapy for advanced or metastatic urothelial carcinoma. N Engl J Med. 2020;383(13):1218–1230. 10.1056/NEJMoa2002788. [Abstract] [CrossRef] [Google Scholar]
559. Plimack ER, Bellmunt J, Gupta S, Berger R, Chow LQ, Juco J, Lunceford J, Saraf S, Perini RF, O'Donnell PH. Safety and activity of pembrolizumab in patients with locally advanced or metastatic urothelial cancer (KEYNOTE-012): a non-randomised, open-label, phase 1b study. Lancet Oncol. 2017;18(2):212–220. 10.1016/S1470-2045(17)30007-4. [Abstract] [CrossRef] [Google Scholar]
560. Balar AV, Castellano D, O'Donnell PH, Grivas P, Vuky J, Powles T, Plimack ER, Hahn NM, de Wit R, Pang L, et al. First-line pembrolizumab in cisplatin-ineligible patients with locally advanced and unresectable or metastatic urothelial cancer (KEYNOTE-052): a multicentre, single-arm, phase 2 study. Lancet Oncol. 2017;18(11):1483–1492. 10.1016/S1470-2045(17)30616-2. [Abstract] [CrossRef] [Google Scholar]
561. Sharma P, Callahan MK, Bono P, Kim J, Spiliopoulou P, Calvo E, Pillai RN, Ott PA, de Braud F, Morse M, et al. Nivolumab monotherapy in recurrent metastatic urothelial carcinoma (CheckMate 032): a multicentre, open-label, two-stage, multi-arm, phase 1/2 trial. Lancet Oncol. 2016;17(11):1590–1598. 10.1016/S1470-2045(16)30496-X. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
562. Sharma P, Retz M, Siefker-Radtke A, Baron A, Necchi A, Bedke J, Plimack ER, Vaena D, Grimm MO, Bracarda S, et al. Nivolumab in metastatic urothelial carcinoma after platinum therapy (CheckMate 275): a multicentre, single-arm, phase 2 trial. Lancet Oncol. 2017;18(3):312–322. 10.1016/S1470-2045(17)30065-7. [Abstract] [CrossRef] [Google Scholar]
563. Challita-Eid PM, Satpayev D, Yang P, An Z, Morrison K, Shostak Y, Raitano A, Nadell R, Liu W, Lortie DR, et al. Enfortumab vedotin antibody-drug conjugate targeting nectin-4 is a highly potent therapeutic agent in multiple preclinical cancer models. Cancer Res. 2016;76(10):3003–3013. 10.1158/0008-5472.CAN-15-1313. [Abstract] [CrossRef] [Google Scholar]
564. Heath EI, Rosenberg JE. The biology and rationale of targeting nectin-4 in urothelial carcinoma. Nat Rev Urol. 2021;18(2):93–103. 10.1038/s41585-020-00394-5. [Abstract] [CrossRef] [Google Scholar]
565. Takano A, Ishikawa N, Nishino R, Masuda K, Yasui W, Inai K, Nishimura H, Ito H, Nakayama H, Miyagi Y, et al. Identification of nectin-4 oncoprotein as a diagnostic and therapeutic target for lung cancer. Cancer Res. 2009;69(16):6694–6703. 10.1158/0008-5472.CAN-09-0016. [Abstract] [CrossRef] [Google Scholar]
566. Siddharth S, Goutam K, Das S, Nayak A, Nayak D, Sethy C, Wyatt MD, Kundu CN. Nectin-4 is a breast cancer stem cell marker that induces WNT/beta-catenin signaling via Pi3k/Akt axis. Int J Biochem Cell Biol. 2017;89:85–94. 10.1016/j.biocel.2017.06.007. [Abstract] [CrossRef] [Google Scholar]
567. Zhang Y, Liu S, Wang L, Wu Y, Hao J, Wang Z, Lu W, Wang XA, Zhang F, Cao Y, et al. A novel PI3K/AKT signaling axis mediates Nectin-4-induced gallbladder cancer cell proliferation, metastasis and tumor growth. Cancer Lett. 2016;375(1):179–189. 10.1016/j.canlet.2016.02.049. [Abstract] [CrossRef] [Google Scholar]
568. Zhang Y, Chen P, Yin W, Ji Y, Shen Q, Ni Q. Nectin-4 promotes gastric cancer progression via the PI3K/AKT signaling pathway. Hum Pathol. 2018;72:107–116. 10.1016/j.humpath.2017.10.034. [Abstract] [CrossRef] [Google Scholar]
569. Kedashiro S, Sugiura A, Mizutani K, Takai Y. Nectin-4 cis-interacts with ErbB2 and its trastuzumab-resistant splice variants, enhancing their activation and DNA synthesis. Sci Rep. 2019;9(1):18997. 10.1038/s41598-019-55460-9. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
570. Tomiyama E, Fujita K, Rodriguez Pena MDC, Taheri D, Banno E, Kato T, Hatano K, Kawashima A, Ujike T, Uemura M, et al. Expression of nectin-4 and PD-L1 in upper tract urothelial carcinoma. Int J Mol Sci. 2020 10.3390/ijms21155390. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
571. Powles T, Rosenberg JE, Sonpavde GP, Loriot Y, Duran I, Lee JL, Matsubara N, Vulsteke C, Castellano D, Wu C, et al. Enfortumab vedotin in previously treated advanced urothelial carcinoma. N Engl J Med. 2021;384(12):1125–1135. 10.1056/NEJMoa2035807. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
572. Yu EY, Petrylak DP, O'Donnell PH, Lee JL, van der Heijden MS, Loriot Y, Stein MN, Necchi A, Kojima T, Harrison MR, et al. Enfortumab vedotin after PD-1 or PD-L1 inhibitors in cisplatin-ineligible patients with advanced urothelial carcinoma (EV201): a multicentre, single-arm, phase 2 trial. Lancet Oncol. 2021;22(6):872–882. 10.1016/S1470-2045(21)00094-2. [Abstract] [CrossRef] [Google Scholar]
573. Chow KC, Macdonald TL, Ross WE. DNA binding by epipodophyllotoxins and N-acyl anthracyclines: implications for mechanism of topoisomerase II inhibition. Mol Pharmacol. 1988;34(4):467–473. [Abstract] [Google Scholar]
574. Blum RH, Garnick MB, Israel M, Panellos GP, Henderson IC, Frei E 3rd. Preclinical rationale and phase I clinical trial of the adriamycin analog, AD 32. Recent Results Cancer Res 1981; 76:7–15. [Abstract]
575. Onrust SV, Lamb HM. Valrubicin. Drugs Aging. 1999;15(1):69–75. 10.2165/00002512-199915010-00006. [Abstract] [CrossRef] [Google Scholar]
576. Steinberg G, Bahnson R, Brosman S, Middleton R, Wajsman Z, Wehle M. Efficacy and safety of valrubicin for the treatment of Bacillus Calmette-Guerin refractory carcinoma in situ of the bladder. The Valrubicin Study Group. J Urol. 2000;163(3):761–767. 10.1016/S0022-5347(05)67799-3. [Abstract] [CrossRef] [Google Scholar]
577. Helsten T, Elkin S, Arthur E, Tomson BN, Carter J, Kurzrock R. The FGFR Landscape in cancer: analysis of 4,853 tumors by next-generation sequencing. Clin Cancer Res. 2016;22(1):259–267. 10.1158/1078-0432.CCR-14-3212. [Abstract] [CrossRef] [Google Scholar]
578. Karkera JD, Cardona GM, Bell K, Gaffney D, Portale JC, Santiago-Walker A, Moy CH, King P, Sharp M, Bahleda R, et al. Oncogenic characterization and pharmacologic sensitivity of activating fibroblast growth factor receptor (FGFR) genetic alterations to the selective FGFR inhibitor erdafitinib. Mol Cancer Ther. 2017;16(8):1717–1726. 10.1158/1535-7163.MCT-16-0518. [Abstract] [CrossRef] [Google Scholar]
579. Markham A. Erdafitinib: first global approval. Drugs. 2019;79(9):1017–1021. 10.1007/s40265-019-01142-9. [Abstract] [CrossRef] [Google Scholar]
580. Roskoski R., Jr The role of fibroblast growth factor receptor (FGFR) protein-tyrosine kinase inhibitors in the treatment of cancers including those of the urinary bladder. Pharmacol Res. 2020;151:104567. 10.1016/j.phrs.2019.104567. [Abstract] [CrossRef] [Google Scholar]
581. Joerger M, Cassier PA, Penel N, Cathomas R, Richly H, Schostak M, Janitzky A, Wermke M, Nogova L, Tai DW-M, et al. Rogaratinib in patients with advanced urothelial carcinomas prescreened for tumor FGFR mRNA expression and effects of mutations in the FGFR signaling pathway. J Clin Oncol. 2018;36(15):4513–4513. 10.1200/JCO.2018.36.15_suppl.4513. [CrossRef] [Google Scholar]
582. Necchi A, Pouessel D, Leibowitz-Amit R, Flechon A, Gupta S, Barthelemy P, Maio M, Zhu X, Asatiani E, Serbest G, et al. Interim results of fight-201, a phase II, open-label, multicenter study of INCB054828 in patients (pts) with metastatic or surgically unresectable urothelial carcinoma (UC) harboring fibroblast growth factor (FGF)/FGF receptor (FGFR) genetic alterations (GA) Annals of Oncology. 2018;29:319–320. 10.1093/annonc/mdy283.109. [CrossRef] [Google Scholar]
583. Pal SK, Rosenberg JE, Hoffman-Censits JH, Berger R, Quinn DI, Galsky MD, Wolf J, Dittrich C, Keam B, Delord JP, et al. Efficacy of BGJ398, a fibroblast growth factor receptor 1–3 inhibitor, in patients with previously treated advanced urothelial carcinoma with fgfr3 alterations. Cancer Discov. 2018;8(7):812–821. 10.1158/2159-8290.CD-18-0229. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
584. Loriot Y, Necchi A, Park SH, Garcia-Donas J, Huddart R, Burgess E, Fleming M, Rezazadeh A, Mellado B, Varlamov S, et al. Erdafitinib in locally advanced or metastatic urothelial carcinoma. N Engl J Med. 2019;381(4):338–348. 10.1056/NEJMoa1817323. [Abstract] [CrossRef] [Google Scholar]
585. Weber CK, Slupsky JR, Kalmes HA, Rapp UR. Active Ras induces heterodimerization of cRaf and BRaf. Cancer Res. 2001;61(9):3595–3598. [Abstract] [Google Scholar]
586. Aplin AE, Kaplan FM, Shao Y. Mechanisms of resistance to RAF inhibitors in melanoma. J Invest Dermatol. 2011;131(9):1817–1820. 10.1038/jid.2011.147. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
587. Jakob JA, Bassett RL, Jr, Ng CS, Curry JL, Joseph RW, Alvarado GC, Rohlfs ML, Richard J, Gershenwald JE, Kim KB, et al. NRAS mutation status is an independent prognostic factor in metastatic melanoma. Cancer. 2012;118(16):4014–4023. 10.1002/cncr.26724. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
588. Karasarides M, Chiloeches A, Hayward R, Niculescu-Duvaz D, Scanlon I, Friedlos F, Ogilvie L, Hedley D, Martin J, Marshall CJ, et al. B-RAF is a therapeutic target in melanoma. Oncogene. 2004;23(37):6292–6298. 10.1038/sj.onc.1207785. [Abstract] [CrossRef] [Google Scholar]
589. Hoeflich KP, Gray DC, Eby MT, Tien JY, Wong L, Bower J, Gogineni A, Zha J, Cole MJ, Stern HM, et al. Oncogenic BRAF is required for tumor growth and maintenance in melanoma models. Cancer Res. 2006;66(2):999–1006. 10.1158/0008-5472.CAN-05-2720. [Abstract] [CrossRef] [Google Scholar]
590. Bollag G, Tsai J, Zhang J, Zhang C, Ibrahim P, Nolop K, Hirth P. Vemurafenib: the first drug approved for BRAF-mutant cancer. Nat Rev Drug Discov. 2012;11(11):873–886. 10.1038/nrd3847. [Abstract] [CrossRef] [Google Scholar]
591. Chapman PB, Hauschild A, Robert C, Haanen JB, Ascierto P, Larkin J, Dummer R, Garbe C, Testori A, Maio M, et al. Improved survival with vemurafenib in melanoma with BRAF V600E mutation. N Engl J Med. 2011;364(26):2507–2516. 10.1056/NEJMoa1103782. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
592. Hauschild A, Grob JJ, Demidov LV, Jouary T, Gutzmer R, Millward M, Rutkowski P, Blank CU, Miller WH, Jr, Kaempgen E, et al. Dabrafenib in BRAF-mutated metastatic melanoma: a multicentre, open-label, phase 3 randomised controlled trial. Lancet. 2012;380(9839):358–365. 10.1016/S0140-6736(12)60868-X. [Abstract] [CrossRef] [Google Scholar]
593. Waizenegger IC, Baum A, Steurer S, Stadtmuller H, Bader G, Schaaf O, Garin-Chesa P, Schlattl A, Schweifer N, Haslinger C, et al. A novel RAF kinase inhibitor with DFG-out-binding mode: high efficacy in BRAF-mutant tumor xenograft models in the absence of normal tissue hyperproliferation. Mol Cancer Ther. 2016;15(3):354–365. 10.1158/1535-7163.MCT-15-0617. [Abstract] [CrossRef] [Google Scholar]
594. Grob JJ, Amonkar MM, Karaszewska B, Schachter J, Dummer R, Mackiewicz A, Stroyakovskiy D, Drucis K, Grange F, Chiarion-Sileni V, et al. Comparison of dabrafenib and trametinib combination therapy with vemurafenib monotherapy on health-related quality of life in patients with unresectable or metastatic cutaneous BRAF Val600-mutation-positive melanoma (COMBI-v): results of a phase 3, open-label, randomised trial. Lancet Oncol. 2015;16(13):1389–1398. 10.1016/S1470-2045(15)00087-X. [Abstract] [CrossRef] [Google Scholar]
595. Sullivan RJ, Flaherty KT. Resistance to BRAF-targeted therapy in melanoma. Eur J Cancer. 2013;49(6):1297–1304. 10.1016/j.ejca.2012.11.019. [Abstract] [CrossRef] [Google Scholar]
596. Shirley M. Encorafenib and binimetinib: first global approvals. Drugs. 2018;78(12):1277–1284. 10.1007/s40265-018-0963-x. [Abstract] [CrossRef] [Google Scholar]
597. Yao Z, Torres NM, Tao A, Gao Y, Luo L, Li Q, de Stanchina E, Abdel-Wahab O, Solit DB, Poulikakos PI, et al. BRAF mutants evade ERK-dependent feedback by different mechanisms that determine their sensitivity to pharmacologic inhibition. Cancer Cell. 2015;28(3):370–383. 10.1016/j.ccell.2015.08.001. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
598. Karoulia Z, Gavathiotis E, Poulikakos PI. New perspectives for targeting RAF kinase in human cancer. Nat Rev Cancer. 2017;17(11):676–691. 10.1038/nrc.2017.79. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
599. Dummer R, Ascierto PA, Gogas HJ, Arance A, Mandala M, Liszkay G, Garbe C, Schadendorf D, Krajsova I, Gutzmer R, et al. Encorafenib plus binimetinib versus vemurafenib or encorafenib in patients with BRAF-mutant melanoma (COLUMBUS): a multicentre, open-label, randomised phase 3 trial. Lancet Oncol. 2018;19(5):603–615. 10.1016/S1470-2045(18)30142-6. [Abstract] [CrossRef] [Google Scholar]
600. Dummer R, Ascierto PA, Gogas HJ, Arance A, Mandala M, Liszkay G, Garbe C, Schadendorf D, Krajsova I, Gutzmer R, et al. Overall survival in patients with BRAF-mutant melanoma receiving encorafenib plus binimetinib versus vemurafenib or encorafenib (COLUMBUS): a multicentre, open-label, randomised, phase 3 trial. Lancet Oncol. 2018;19(10):1315–1327. 10.1016/S1470-2045(18)30497-2. [Abstract] [CrossRef] [Google Scholar]
601. Solit DB, Garraway LA, Pratilas CA, Sawai A, Getz G, Basso A, Ye Q, Lobo JM, She Y, Osman I, et al. BRAF mutation predicts sensitivity to MEK inhibition. Nature. 2006;439(7074):358–362. 10.1038/nature04304. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
602. Curti BD, Faries MB. Recent advances in the treatment of melanoma. N Engl J Med. 2021;384(23):2229–2240. 10.1056/NEJMra2034861. [Abstract] [CrossRef] [Google Scholar]
603. Wright CJ, McCormack PL. Trametinib: first global approval. Drugs. 2013;73(11):1245–1254. 10.1007/s40265-013-0096-1. [Abstract] [CrossRef] [Google Scholar]
604. Gilmartin AG, Bleam MR, Groy A, Moss KG, Minthorn EA, Kulkarni SG, Rominger CM, Erskine S, Fisher KE, Yang J, et al. GSK1120212 (JTP-74057) is an inhibitor of MEK activity and activation with favorable pharmacokinetic properties for sustained in vivo pathway inhibition. Clin Cancer Res. 2011;17(5):989–1000. 10.1158/1078-0432.CCR-10-2200. [Abstract] [CrossRef] [Google Scholar]
605. Garnock-Jones KP. Cobimetinib: first global approval. Drugs. 2015;75(15):1823–1830. 10.1007/s40265-015-0477-8. [Abstract] [CrossRef] [Google Scholar]
606. Ascierto PA, Schadendorf D, Berking C, Agarwala SS, van Herpen CM, Queirolo P, Blank CU, Hauschild A, Beck JT, St-Pierre A, et al. MEK162 for patients with advanced melanoma harbouring NRAS or Val600 BRAF mutations: a non-randomised, open-label phase 2 study. Lancet Oncol. 2013;14(3):249–256. 10.1016/S1470-2045(13)70024-X. [Abstract] [CrossRef] [Google Scholar]
607. Long GV, Hauschild A, Santinami M, Atkinson V, Mandala M, Chiarion-Sileni V, Larkin J, Nyakas M, Dutriaux C, Haydon A, et al. Adjuvant dabrafenib plus trametinib in stage III BRAF-mutated melanoma. N Engl J Med. 2017;377(19):1813–1823. 10.1056/NEJMoa1708539. [Abstract] [CrossRef] [Google Scholar]
608. Dummer R, Hauschild A, Santinami M, Atkinson V, Mandala M, Kirkwood JM, Chiarion Sileni V, Larkin J, Nyakas M, Dutriaux C, et al. Five-year analysis of adjuvant dabrafenib plus trametinib in stage III melanoma. N Engl J Med. 2020;383(12):1139–1148. 10.1056/NEJMoa2005493. [Abstract] [CrossRef] [Google Scholar]
609. Larkin J, Ascierto PA, Dreno B, Atkinson V, Liszkay G, Maio M, Mandala M, Demidov L, Stroyakovskiy D, Thomas L, et al. Combined vemurafenib and cobimetinib in BRAF-mutated melanoma. N Engl J Med. 2014;371(20):1867–1876. 10.1056/NEJMoa1408868. [Abstract] [CrossRef] [Google Scholar]
610. Ascierto PA, McArthur GA, Dreno B, Atkinson V, Liszkay G, Di Giacomo AM, Mandala M, Demidov L, Stroyakovskiy D, Thomas L, et al. Cobimetinib combined with vemurafenib in advanced BRAF(V600)-mutant melanoma (coBRIM): updated efficacy results from a randomised, double-blind, phase 3 trial. Lancet Oncol. 2016;17(9):1248–1260. 10.1016/S1470-2045(16)30122-X. [Abstract] [CrossRef] [Google Scholar]
611. Walunas TL, Lenschow DJ, Bakker CY, Linsley PS, Freeman GJ, Green JM, Thompson CB, Bluestone JA. CTLA-4 can function as a negative regulator of T cell activation. Immunity. 1994;1(5):405–413. 10.1016/1074-7613(94)90071-X. [Abstract] [CrossRef] [Google Scholar]
612. Leach DR, Krummel MF, Allison JP. Enhancement of antitumor immunity by CTLA-4 blockade. Science. 1996;271(5256):1734–1736. 10.1126/science.271.5256.1734. [Abstract] [CrossRef] [Google Scholar]
613. Lipson EJ, Drake CG. Ipilimumab: an anti-CTLA-4 antibody for metastatic melanoma. Clin Cancer Res. 2011;17(22):6958–6962. 10.1158/1078-0432.CCR-11-1595. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
614. Hodi FS, O'Day SJ, McDermott DF, Weber RW, Sosman JA, Haanen JB, Gonzalez R, Robert C, Schadendorf D, Hassel JC, et al. Improved survival with ipilimumab in patients with metastatic melanoma. N Engl J Med. 2010;363(8):711–723. 10.1056/NEJMoa1003466. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
615. Freeman GJ, Long AJ, Iwai Y, Bourque K, Chernova T, Nishimura H, Fitz LJ, Malenkovich N, Okazaki T, Byrne MC, et al. Engagement of the PD-1 immunoinhibitory receptor by a novel B7 family member leads to negative regulation of lymphocyte activation. J Exp Med. 2000;192(7):1027–1034. 10.1084/jem.192.7.1027. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
616. Latchman Y, Wood CR, Chernova T, Chaudhary D, Borde M, Chernova I, Iwai Y, Long AJ, Brown JA, Nunes R, et al. PD-L2 is a second ligand for PD-1 and inhibits T cell activation. Nat Immunol. 2001;2(3):261–268. 10.1038/85330. [Abstract] [CrossRef] [Google Scholar]
617. Kleffel S, Posch C, Barthel SR, Mueller H, Schlapbach C, Guenova E, Elco CP, Lee N, Juneja VR, Zhan Q, et al. Melanoma cell-intrinsic PD-1 receptor functions promote tumor growth. Cell. 2015;162(6):1242–1256. 10.1016/j.cell.2015.08.052. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
618. Lee JY, Lee HT, Shin W, Chae J, Choi J, Kim SH, Lim H, Won Heo T, Park KY, Lee YJ, et al. Structural basis of checkpoint blockade by monoclonal antibodies in cancer immunotherapy. Nat Commun. 2016;7:13354. 10.1038/ncomms13354. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
619. Na Z, Yeo SP, Bharath SR, Bowler MW, Balikci E, Wang CI, Song H. Structural basis for blocking PD-1-mediated immune suppression by therapeutic antibody pembrolizumab. Cell Res. 2017;27(1):147–150. 10.1038/cr.2016.77. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
620. Fessas P, Lee H, Ikemizu S, Janowitz T. A molecular and preclinical comparison of the PD-1-targeted T-cell checkpoint inhibitors nivolumab and pembrolizumab. Semin Oncol. 2017;44(2):136–140. 10.1053/j.seminoncol.2017.06.002. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
621. Robert C, Schachter J, Long GV, Arance A, Grob JJ, Mortier L, Daud A, Carlino MS, McNeil C, Lotem M, et al. Pembrolizumab versus Ipilimumab in advanced melanoma. N Engl J Med. 2015;372(26):2521–2532. 10.1056/NEJMoa1503093. [Abstract] [CrossRef] [Google Scholar]
622. Schachter J, Ribas A, Long GV, Arance A, Grob JJ, Mortier L, Daud A, Carlino MS, McNeil C, Lotem M, et al. Pembrolizumab versus ipilimumab for advanced melanoma: final overall survival results of a multicentre, randomised, open-label phase 3 study (KEYNOTE-006) Lancet. 2017;390(10105):1853–1862. 10.1016/S0140-6736(17)31601-X. [Abstract] [CrossRef] [Google Scholar]
623. Robert C, Ribas A, Wolchok JD, Hodi FS, Hamid O, Kefford R, Weber JS, Joshua AM, Hwu WJ, Gangadhar TC, et al. Anti-programmed-death-receptor-1 treatment with pembrolizumab in ipilimumab-refractory advanced melanoma: a randomised dose-comparison cohort of a phase 1 trial. Lancet. 2014;384(9948):1109–1117. 10.1016/S0140-6736(14)60958-2. [Abstract] [CrossRef] [Google Scholar]
624. Tan S, Zhang H, Chai Y, Song H, Tong Z, Wang Q, Qi J, Wong G, Zhu X, Liu WJ, et al. An unexpected N-terminal loop in PD-1 dominates binding by nivolumab. Nat Commun. 2017;8:14369. 10.1038/ncomms14369. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
625. Wolchok JD, Kluger H, Callahan MK, Postow MA, Rizvi NA, Lesokhin AM, Segal NH, Ariyan CE, Gordon RA, Reed K, et al. Nivolumab plus ipilimumab in advanced melanoma. N Engl J Med. 2013;369(2):122–133. 10.1056/NEJMoa1302369. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
626. Tawbi HA, Forsyth PA, Algazi A, Hamid O, Hodi FS, Moschos SJ, Khushalani NI, Lewis K, Lao CD, Postow MA, et al. Combined nivolumab and ipilimumab in melanoma metastatic to the brain. N Engl J Med. 2018;379(8):722–730. 10.1056/NEJMoa1805453. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
627. Wolchok JD, Chiarion-Sileni V, Gonzalez R, Rutkowski P, Grob JJ, Cowey CL, Lao CD, Wagstaff J, Schadendorf D, Ferrucci PF, et al. Overall survival with combined nivolumab and ipilimumab in advanced melanoma. N Engl J Med. 2017;377(14):1345–1356. 10.1056/NEJMoa1709684. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
628. Larkin J, Chiarion-Sileni V, Gonzalez R, Grob JJ, Rutkowski P, Lao CD, Cowey CL, Schadendorf D, Wagstaff J, Dummer R, et al. Five-year survival with combined nivolumab and ipilimumab in advanced melanoma. N Engl J Med. 2019;381(16):1535–1546. 10.1056/NEJMoa1910836. [Abstract] [CrossRef] [Google Scholar]
629. Nehal KS, Bichakjian CK. Update on keratinocyte carcinomas. N Engl J Med. 2018;379(4):363–374. 10.1056/NEJMra1708701. [Abstract] [CrossRef] [Google Scholar]
630. Rogers HW, Weinstock MA, Feldman SR, Coldiron BM. Incidence estimate of nonmelanoma skin cancer (Keratinocyte Carcinomas) in the US population, 2012. JAMA Dermatol. 2015;151(10):1081–1086. 10.1001/jamadermatol.2015.1187. [Abstract] [CrossRef] [Google Scholar]
631. Zelin E, Zalaudek I, Agozzino M, Dianzani C, Dri A, Di Meo N, Giuffrida R, Marangi GF, Neagu N, Persichetti P, et al. Neoadjuvant therapy for non-melanoma skin cancer: updated therapeutic approaches for basal, squamous, and merkel cell carcinoma. Curr Treat Options Oncol. 2021;22(4):35. 10.1007/s11864-021-00826-3. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
632. Madan V, Lear JT, Szeimies RM. Non-melanoma skin cancer. Lancet. 2010;375(9715):673–685. 10.1016/S0140-6736(09)61196-X. [Abstract] [CrossRef] [Google Scholar]
633. Rubin AI, Chen EH, Ratner D. Basal-cell carcinoma. N Engl J Med. 2005;353(21):2262–2269. 10.1056/NEJMra044151. [Abstract] [CrossRef] [Google Scholar]
634. Sanchez-Danes A, Hannezo E, Larsimont JC, Liagre M, Youssef KK, Simons BD, Blanpain C. Defining the clonal dynamics leading to mouse skin tumour initiation. Nature. 2016;536(7616):298–303. 10.1038/nature19069. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
635. Gailani MR, Stahle-Backdahl M, Leffell DJ, Glynn M, Zaphiropoulos PG, Pressman C, Unden AB, Dean M, Brash DE, Bale AE, et al. The role of the human homologue of Drosophila patched in sporadic basal cell carcinomas. Nat Genet. 1996;14(1):78–81. 10.1038/ng0996-78. [Abstract] [CrossRef] [Google Scholar]
636. Dlugosz A, Agrawal S, Kirkpatrick P. Vismodegib. Nat Rev Drug Discov. 2012;11(6):437–438. 10.1038/nrd3753. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
637. Casey D, Demko S, Shord S, Zhao H, Chen H, He K, Putman A, Helms W, Keegan P, Pazdur R. FDA approval summary: sonidegib for locally advanced basal cell carcinoma. Clin Cancer Res. 2017;23(10):2377–2381. 10.1158/1078-0432.CCR-16-2051. [Abstract] [CrossRef] [Google Scholar]
638. Wang C, Wu H, Evron T, Vardy E, Han GW, Huang XP, Hufeisen SJ, Mangano TJ, Urban DJ, Katritch V, et al. Structural basis for Smoothened receptor modulation and chemoresistance to anticancer drugs. Nat Commun. 2014;5:4355. 10.1038/ncomms5355. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
639. Byrne EFX, Sircar R, Miller PS, Hedger G, Luchetti G, Nachtergaele S, Tully MD, Mydock-McGrane L, Covey DF, Rambo RP, et al. Structural basis of smoothened regulation by its extracellular domains. Nature. 2016;535(7613):517–522. 10.1038/nature18934. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
640. Rudin CM. Vismodegib. Clin Cancer Res. 2012;18(12):3218–3222. 10.1158/1078-0432.CCR-12-0568. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
641. Chang AL, Oro AE. Initial assessment of tumor regrowth after vismodegib in advanced basal cell carcinoma. Arch Dermatol. 2012;148(11):1324–1325. 10.1001/archdermatol.2012.2354. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
642. Yauch RL, Dijkgraaf GJ, Alicke B, Januario T, Ahn CP, Holcomb T, Pujara K, Stinson J, Callahan CA, Tang T, et al. Smoothened mutation confers resistance to a Hedgehog pathway inhibitor in medulloblastoma. Science. 2009;326(5952):572–574. 10.1126/science.1179386. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
643. Sharpe HJ, Pau G, Dijkgraaf GJ, Basset-Seguin N, Modrusan Z, Januario T, Tsui V, Durham AB, Dlugosz AA, Haverty PM, et al. Genomic analysis of smoothened inhibitor resistance in basal cell carcinoma. Cancer Cell. 2015;27(3):327–341. 10.1016/j.ccell.2015.02.001. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
644. Atwood SX, Sarin KY, Whitson RJ, Li JR, Kim G, Rezaee M, Ally MS, Kim J, Yao C, Chang AL, et al. Smoothened variants explain the majority of drug resistance in basal cell carcinoma. Cancer Cell. 2015;27(3):342–353. 10.1016/j.ccell.2015.02.002. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
645. Atwood SX, Sarin KY, Li JR, Yao CY, Urman NM, Chang ALS, Tang JY, Oro AE. Rolling the genetic dice: neutral and deleterious smoothened mutations in drug-resistant basal cell carcinoma. J Invest Dermatol. 2015;135(8):2138–2141. 10.1038/jid.2015.115. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
646. Dong X, Wang C, Chen Z, Zhao W. Overcoming the resistance mechanisms of smoothened inhibitors. Drug Discov Today. 2018;23(3):704–710. 10.1016/j.drudis.2018.01.012. [Abstract] [CrossRef] [Google Scholar]
647. Burness CB. Sonidegib: first global approval. Drugs. 2015;75(13):1559–1566. 10.1007/s40265-015-0458-y. [Abstract] [CrossRef] [Google Scholar]
648. Danial C, Sarin KY, Oro AE, Chang AL. An investigator-initiated open-label trial of sonidegib in advanced basal cell carcinoma patients resistant to vismodegib. Clin Cancer Res. 2016;22(6):1325–1329. 10.1158/1078-0432.CCR-15-1588. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
649. Becker JC, Stang A, DeCaprio JA, Cerroni L, Lebbe C, Veness M, Nghiem P. Merkel cell carcinoma. Nat Rev Dis Primers. 2017;3:17077. 10.1038/nrdp.2017.77. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
650. Guenole M, Benigni P, Bourbonne V, Lucia F, Legoupil D, Pradier O, Misery L, Uguen A, Schick U. The prognostic significance of PD-L1 expression on tumor and immune cells in Merkel cell carcinoma. J Cancer Res Clin Oncol. 2021;147(9):2569–2578. 10.1007/s00432-021-03676-6. [Abstract] [CrossRef] [Google Scholar]
651. Lipson EJ, Vincent JG, Loyo M, Kagohara LT, Luber BS, Wang H, Xu H, Nayar SK, Wang TS, Sidransky D, et al. PD-L1 expression in the Merkel cell carcinoma microenvironment: association with inflammation, Merkel cell polyomavirus and overall survival. Cancer Immunol Res. 2013;1(1):54–63. 10.1158/2326-6066.CIR-13-0034. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
652. Liu K, Tan S, Chai Y, Chen D, Song H, Zhang CW, Shi Y, Liu J, Tan W, Lyu J, et al. Structural basis of anti-PD-L1 monoclonal antibody avelumab for tumor therapy. Cell Res. 2017;27(1):151–153. 10.1038/cr.2016.102. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
653. Kaufman HL, Russell J, Hamid O, Bhatia S, Terheyden P, D'Angelo SP, Shih KC, Lebbe C, Linette GP, Milella M, et al. Avelumab in patients with chemotherapy-refractory metastatic Merkel cell carcinoma: a multicentre, single-group, open-label, phase 2 trial. Lancet Oncol. 2016;17(10):1374–1385. 10.1016/S1470-2045(16)30364-3. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
654. D'Angelo SP, Russell J, Lebbe C, Chmielowski B, Gambichler T, Grob JJ, Kiecker F, Rabinowits G, Terheyden P, Zwiener I, et al. Efficacy and safety of first-line avelumab treatment in patients with stage iv metastatic merkel cell carcinoma: a preplanned interim analysis of a clinical trial. JAMA Oncol. 2018;4(9):e180077. 10.1001/jamaoncol.2018.0077. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
655. Nghiem P, Bhatia S, Lipson EJ, Sharfman WH, Kudchadkar RR, Brohl AS, Friedlander PA, Daud A, Kluger HM, Reddy SA, et al. Durable tumor regression and overall survival in patients with advanced merkel cell carcinoma receiving pembrolizumab as first-line therapy. J Clin Oncol. 2019;37(9):693–702. 10.1200/JCO.18.01896. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
656. Nghiem PT, Bhatia S, Lipson EJ, Kudchadkar RR, Miller NJ, Annamalai L, Berry S, Chartash EK, Daud A, Fling SP, et al. PD-1 Blockade with pembrolizumab in advanced merkel-cell carcinoma. N Engl J Med. 2016;374(26):2542–2552. 10.1056/NEJMoa1603702. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
657. Topalian SL, Bhatia S, Amin A, Kudchadkar RR, Sharfman WH, Lebbe C, Delord JP, Dunn LA, Shinohara MM, Kulikauskas R, et al. Neoadjuvant nivolumab for patients with resectable Merkel cell carcinoma in the CheckMate 358 Trial. J Clin Oncol. 2020;38(22):2476–2487. 10.1200/JCO.20.00201. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
658. Alam M, Ratner D. Cutaneous squamous-cell carcinoma. N Engl J Med. 2001;344(13):975–983. 10.1056/NEJM200103293441306. [Abstract] [CrossRef] [Google Scholar]
659. Stern RS, Laird N, Melski J, Parrish JA, Fitzpatrick TB, Bleich HL. Cutaneous squamous-cell carcinoma in patients treated with PUVA. N Engl J Med. 1984;310(18):1156–1161. 10.1056/NEJM198405033101805. [Abstract] [CrossRef] [Google Scholar]
660. Varki V, Ioffe OB, Bentzen SM, Heath J, Cellini A, Feliciano J, Zandberg DP. PD-L1, B7–H3, and PD-1 expression in immunocompetent vs. immunosuppressed patients with cutaneous squamous cell carcinoma. Cancer Immunol Immunother. 2018;67(5):805–814. 10.1007/s00262-018-2138-8. [Abstract] [CrossRef] [Google Scholar]
661. Liu K, Tan S, Jin W, Guan J, Wang Q, Sun H, Qi J, Yan J, Chai Y, Wang Z, et al. N-glycosylation of PD-1 promotes binding of camrelizumab. EMBO Rep. 2020;21(12):e51444. 10.15252/embr.202051444. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
662. Migden MR, Rischin D, Schmults CD, Guminski A, Hauschild A, Lewis KD, Chung CH, Hernandez-Aya L, Lim AM, Chang ALS, et al. PD-1 Blockade with cemiplimab in advanced cutaneous squamous-cell carcinoma. N Engl J Med. 2018;379(4):341–351. 10.1056/NEJMoa1805131. [Abstract] [CrossRef] [Google Scholar]
663. Migden MR, Khushalani NI, Chang ALS, Lewis KD, Schmults CD, Hernandez-Aya L, Meier F, Schadendorf D, Guminski A, Hauschild A, et al. Cemiplimab in locally advanced cutaneous squamous cell carcinoma: results from an open-label, phase 2, single-arm trial. Lancet Oncol. 2020;21(2):294–305. 10.1016/S1470-2045(19)30728-4. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
664. Markham A, Duggan S. Cemiplimab: first global approval. Drugs. 2018;78(17):1841–1846. 10.1007/s40265-018-1012-5. [Abstract] [CrossRef] [Google Scholar]
665. Cabanillas ME, McFadden DG, Durante C. Thyroid cancer. Lancet. 2016;388(10061):2783–2795. 10.1016/S0140-6736(16)30172-6. [Abstract] [CrossRef] [Google Scholar]
666. Amdur RJ, Mazzaferri EL: Recombinant Human TSH: Background information and standard protocol. In: Essentials of thyroid cancer management. 2005; 233–237.
667. Mallick U, Harmer C, Yap B, Wadsley J, Clarke S, Moss L, Nicol A, Clark PM, Farnell K, McCready R, et al. Ablation with low-dose radioiodine and thyrotropin alfa in thyroid cancer. N Engl J Med. 2012;366(18):1674–1685. 10.1056/NEJMoa1109589. [Abstract] [CrossRef] [Google Scholar]
668. Juweid M, O'Dorisio T, Milhem M. Diagnosis of poorly differentiated thyroid cancer with radioiodine scanning after thyrotropin alfa stimulation. N Engl J Med. 2008;359(12):1295–1297. 10.1056/NEJMc0803606. [Abstract] [CrossRef] [Google Scholar]
669. Mazzaferri EL, Kloos RT. Clinical review 128: current approaches to primary therapy for papillary and follicular thyroid cancer. J Clin Endocrinol Metab. 2001;86(4):1447–1463. 10.1210/jcem.86.4.7407. [Abstract] [CrossRef] [Google Scholar]
670. Spitzweg C, Bible KC, Hofbauer LC, Morris JC. Advanced radioiodine-refractory differentiated thyroid cancer: the sodium iodide symporter and other emerging therapeutic targets. Lancet Diabetes Endocrinol. 2014;2(10):830–842. 10.1016/S2213-8587(14)70051-8. [Abstract] [CrossRef] [Google Scholar]
671. Riesco-Eizaguirre G, Rodriguez I, De la Vieja A, Costamagna E, Carrasco N, Nistal M, Santisteban P. The BRAFV600E oncogene induces transforming growth factor beta secretion leading to sodium iodide symporter repression and increased malignancy in thyroid cancer. Cancer Res. 2009;69(21):8317–8325. 10.1158/0008-5472.CAN-09-1248. [Abstract] [CrossRef] [Google Scholar]
672. Romei C, Ciampi R, Faviana P, Agate L, Molinaro E, Bottici V, Basolo F, Miccoli P, Pacini F, Pinchera A, et al. BRAFV600E mutation, but not RET/PTC rearrangements, is correlated with a lower expression of both thyroperoxidase and sodium iodide symporter genes in papillary thyroid cancer. Endocr Relat Cancer. 2008;15(2):511–520. 10.1677/ERC-07-0130. [Abstract] [CrossRef] [Google Scholar]
673. Garcia B, Santisteban P. PI3K is involved in the IGF-I inhibition of TSH-induced sodium/iodide symporter gene expression. Mol Endocrinol. 2002;16(2):342–352. 10.1210/mend.16.2.0774. [Abstract] [CrossRef] [Google Scholar]
674. Matsui J, Yamamoto Y, Funahashi Y, Tsuruoka A, Watanabe T, Wakabayashi T, Uenaka T, Asada M. E7080, a novel inhibitor that targets multiple kinases, has potent antitumor activities against stem cell factor producing human small cell lung cancer H146, based on angiogenesis inhibition. Int J Cancer. 2008;122(3):664–671. 10.1002/ijc.23131. [Abstract] [CrossRef] [Google Scholar]
675. Matsui J, Funahashi Y, Uenaka T, Watanabe T, Tsuruoka A, Asada M. Multi-kinase inhibitor E7080 suppresses lymph node and lung metastases of human mammary breast tumor MDA-MB-231 via inhibition of vascular endothelial growth factor-receptor (VEGF-R) 2 and VEGF-R3 kinase. Clin Cancer Res. 2008;14(17):5459–5465. 10.1158/1078-0432.CCR-07-5270. [Abstract] [CrossRef] [Google Scholar]
676. Okamoto K, Ikemori-Kawada M, Jestel A, von Konig K, Funahashi Y, Matsushima T, Tsuruoka A, Inoue A, Matsui J. Distinct binding mode of multikinase inhibitor lenvatinib revealed by biochemical characterization. ACS Med Chem Lett. 2015;6(1):89–94. 10.1021/ml500394m. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
677. Schlumberger M, Tahara M, Wirth LJ, Robinson B, Brose MS, Elisei R, Habra MA, Newbold K, Shah MH, Hoff AO, et al. Lenvatinib versus placebo in radioiodine-refractory thyroid cancer. N Engl J Med. 2015;372(7):621–630. 10.1056/NEJMoa1406470. [Abstract] [CrossRef] [Google Scholar]
678. Commander H, Whiteside G, Perry C. Vandetanib: first global approval. Drugs. 2011;71(10):1355–1365. 10.2165/11595310-000000000-00000. [Abstract] [CrossRef] [Google Scholar]
679. Romei C, Cosci B, Renzini G, Bottici V, Molinaro E, Agate L, Passannanti P, Viola D, Biagini A, Basolo F, et al. RET genetic screening of sporadic medullary thyroid cancer (MTC) allows the preclinical diagnosis of unsuspected gene carriers and the identification of a relevant percentage of hidden familial MTC (FMTC) Clin Endocrinol. 2011;74(2):241–247. 10.1111/j.1365-2265.2010.03900.x. [Abstract] [CrossRef] [Google Scholar]
680. Hu MI, Cote GJ. Medullary thyroid carcinoma: who's on first? Thyroid. 2012;22(5):451–453. 10.1089/thy.2012.2205.ed. [Abstract] [CrossRef] [Google Scholar]
681. Frampton JE. Vandetanib: in medullary thyroid cancer. Drugs. 2012;72(10):1423–1436. 10.2165/11209300-000000000-00000. [Abstract] [CrossRef] [Google Scholar]
682. Rodriguez-Antona C, Pallares J, Montero-Conde C, Inglada-Perez L, Castelblanco E, Landa I, Leskela S, Leandro-Garcia LJ, Lopez-Jimenez E, Leton R, et al. Overexpression and activation of EGFR and VEGFR2 in medullary thyroid carcinomas is related to metastasis. Endocr Relat Cancer. 2010;17(1):7–16. 10.1677/ERC-08-0304. [Abstract] [CrossRef] [Google Scholar]
683. Ivan M, Bond JA, Prat M, Comoglio PM, Wynford-Thomas D. Activated ras and ret oncogenes induce over-expression of c-met (hepatocyte growth factor receptor) in human thyroid epithelial cells. Oncogene. 1997;14(20):2417–2423. 10.1038/sj.onc.1201083. [Abstract] [CrossRef] [Google Scholar]
684. Hoy SM. Cabozantinib: a review of its use in patients with medullary thyroid cancer. Drugs. 2014;74(12):1435–1444. 10.1007/s40265-014-0265-x. [Abstract] [CrossRef] [Google Scholar]
685. Knowles PP, Murray-Rust J, Kjaer S, Scott RP, Hanrahan S, Santoro M, Ibanez CF, McDonald NQ. Structure and chemical inhibition of the RET tyrosine kinase domain. J Biol Chem. 2006;281(44):33577–33587. 10.1074/jbc.M605604200. [Abstract] [CrossRef] [Google Scholar]
686. Giovannetti E, Zucali PA, Assaraf YG, Leon LG, Smid K, Alecci C, Giancola F, Destro A, Gianoncelli L, Lorenzi E, et al. Preclinical emergence of vandetanib as a potent antitumour agent in mesothelioma: molecular mechanisms underlying its synergistic interaction with pemetrexed and carboplatin. Br J Cancer. 2011;105(10):1542–1553. 10.1038/bjc.2011.400. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
687. Inoue K, Torimura T, Nakamura T, Iwamoto H, Masuda H, Abe M, Hashimoto O, Koga H, Ueno T, Yano H, et al. Vandetanib, an inhibitor of VEGF receptor-2 and EGF receptor, suppresses tumor development and improves prognosis of liver cancer in mice. Clin Cancer Res. 2012;18(14):3924–3933. 10.1158/1078-0432.CCR-11-2041. [Abstract] [CrossRef] [Google Scholar]
688. Carlomagno F, Guida T, Anaganti S, Vecchio G, Fusco A, Ryan AJ, Billaud M, Santoro M. Disease associated mutations at valine 804 in the RET receptor tyrosine kinase confer resistance to selective kinase inhibitors. Oncogene. 2004;23(36):6056–6063. 10.1038/sj.onc.1207810. [Abstract] [CrossRef] [Google Scholar]
689. Nakaoku T, Kohno T, Araki M, Niho S, Chauhan R, Knowles PP, Tsuchihara K, Matsumoto S, Shimada Y, Mimaki S, et al. A secondary RET mutation in the activation loop conferring resistance to vandetanib. Nat Commun. 2018;9(1):625. 10.1038/s41467-018-02994-7. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
690. Naresh G, Guruprasad L. Enhanced metastable state models of TAM kinase binding to cabozantinib explains the dynamic nature of receptor tyrosine kinases. J Biomol Struct Dyn. 2021;39(4):1213–1235. 10.1080/07391102.2020.1730968. [Abstract] [CrossRef] [Google Scholar]
691. Yakes FM, Chen J, Tan J, Yamaguchi K, Shi Y, Yu P, Qian F, Chu F, Bentzien F, Cancilla B, et al. Cabozantinib (XL184), a novel MET and VEGFR2 inhibitor, simultaneously suppresses metastasis, angiogenesis, and tumor growth. Mol Cancer Ther. 2011;10(12):2298–2308. 10.1158/1535-7163.MCT-11-0264. [Abstract] [CrossRef] [Google Scholar]
692. Kurzrock R, Sherman SI, Ball DW, Forastiere AA, Cohen RB, Mehra R, Pfister DG, Cohen EE, Janisch L, Nauling F, et al. Activity of XL184 (Cabozantinib), an oral tyrosine kinase inhibitor, in patients with medullary thyroid cancer. J Clin Oncol. 2011;29(19):2660–2666. 10.1200/JCO.2010.32.4145. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
693. Elisei R, Schlumberger MJ, Muller SP, Schoffski P, Brose MS, Shah MH, Licitra L, Jarzab B, Medvedev V, Kreissl MC, et al. Cabozantinib in progressive medullary thyroid cancer. J Clin Oncol. 2013;31(29):3639–3646. 10.1200/JCO.2012.48.4659. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
694. Davare MA, Vellore NA, Wagner JP, Eide CA, Goodman JR, Drilon A, Deininger MW, O'Hare T, Druker BJ. Structural insight into selectivity and resistance profiles of ROS1 tyrosine kinase inhibitors. Proc Natl Acad Sci U S A. 2015;112(39):E5381–5390. 10.1073/pnas.1515281112. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
695. Drilon A, Somwar R, Wagner JP, Vellore NA, Eide CA, Zabriskie MS, Arcila ME, Hechtman JF, Wang L, Smith RS, et al. A novel crizotinib-resistant solvent-front mutation responsive to cabozantinib therapy in a patient with ROS1-rearranged lung cancer. Clin Cancer Res. 2016;22(10):2351–2358. 10.1158/1078-0432.CCR-15-2013. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
696. Gamboa AC, Gronchi A, Cardona K. Soft-tissue sarcoma in adults: An update on the current state of histiotype-specific management in an era of personalized medicine. CA Cancer J Clin. 2020;70(3):200–229. 10.3322/caac.21605. [Abstract] [CrossRef] [Google Scholar]
697. Cesarman E, Chadburn A, Rubinstein PG. KSHV/HHV8-mediated hematologic diseases. Blood 2021. [Europe PMC free article] [Abstract]
698. Ioachim HL, Adsay V, Giancotti FR, Dorsett B, Melamed J. Kaposi's sarcoma of internal organs. A multiparameter study of 86 cases. Cancer. 1995;75(6):1376–1385. 10.1002/1097-0142(19950315)75:6<1376::AID-CNCR2820750621>3.0.CO;2-Y. [Abstract] [CrossRef] [Google Scholar]
699. Zelent A, Krust A, Petkovich M, Kastner P, Chambon P. Cloning of murine alpha and beta retinoic acid receptors and a novel receptor gamma predominantly expressed in skin. Nature. 1989;339(6227):714–717. 10.1038/339714a0. [Abstract] [CrossRef] [Google Scholar]
700. Garnock-Jones KP, Perry CM. Alitretinoin: in severe chronic hand eczema. Drugs. 2009;69(12):1625–1634. 10.2165/11202200-000000000-00000. [Abstract] [CrossRef] [Google Scholar]
701. Zhang XK, Lehmann J, Hoffmann B, Dawson MI, Cameron J, Graupner G, Hermann T, Tran P, Pfahl M. Homodimer formation of retinoid X receptor induced by 9-cis retinoic acid. Nature. 1992;358(6387):587–591. 10.1038/358587a0. [Abstract] [CrossRef] [Google Scholar]
702. Redfern CP, Lovat PE, Malcolm AJ, Pearson AD. Gene expression and neuroblastoma cell differentiation in response to retinoic acid: differential effects of 9-cis and all-trans retinoic acid. Eur J Cancer. 1995;31A(4):486–494. 10.1016/0959-8049(95)00066-R. [Abstract] [CrossRef] [Google Scholar]
703. Brodowicz T, Wiltschke C, Kandioler-Eckersberger D, Grunt TW, Rudas M, Schneider SM, Hejna M, Budinsky A, Zielinski CC. Inhibition of proliferation and induction of apoptosis in soft tissue sarcoma cells by interferon-alpha and retinoids. Br J Cancer. 1999;80(9):1350–1358. 10.1038/sj.bjc.6690528. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
704. Walmsley S, Northfelt DW, Melosky B, Conant M, Friedman-Kien AE, Wagner B. Treatment of AIDS-related cutaneous Kaposi's sarcoma with topical alitretinoin (9-cis-retinoic acid) gel. Panretin Gel North American Study Group. J Acquir Immune Defic Syndr. 1999;22(3):235–246. 10.1097/00126334-199911010-00004. [Abstract] [CrossRef] [Google Scholar]
705. Bodsworth NJ, Bloch M, Bower M, Donnell D, Yocum R. International Panretin Gel KSSG: Phase III vehicle-controlled, multi-centered study of topical alitretinoin gel 0.1% in cutaneous AIDS-related Kaposi's sarcoma. Am J Clin Dermatol. 2001;2(2):77–87. 10.2165/00128071-200102020-00004. [Abstract] [CrossRef] [Google Scholar]
706. Lee ATJ, Thway K, Huang PH, Jones RL. Clinical and molecular spectrum of liposarcoma. J Clin Oncol. 2018;36(2):151–159. 10.1200/JCO.2017.74.9598. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
707. George S, Serrano C, Hensley ML, Ray-Coquard I. Soft tissue and uterine leiomyosarcoma. J Clin Oncol. 2018;36(2):144–150. 10.1200/JCO.2017.75.9845. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
708. El-Rifai W, Sarlomo-Rikala M, Knuutila S, Miettinen M. DNA copy number changes in development and progression in leiomyosarcomas of soft tissues. Am J Pathol. 1998;153(3):985–990. 10.1016/S0002-9440(10)65640-4. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
709. Carter NJ, Keam SJ. Trabectedin : a review of its use in the management of soft tissue sarcoma and ovarian cancer. Drugs. 2007;67(15):2257–2276. 10.2165/00003495-200767150-00009. [Abstract] [CrossRef] [Google Scholar]
710. Carter NJ, Keam SJ. Trabectedin: a review of its use in soft tissue sarcoma and ovarian cancer. Drugs. 2010;70(3):355–376. 10.2165/11202860-000000000-00000. [Abstract] [CrossRef] [Google Scholar]
711. D'Incalci M, Galmarini CM. A review of trabectedin (ET-743): a unique mechanism of action. Mol Cancer Ther. 2010;9(8):2157–2163. 10.1158/1535-7163.MCT-10-0263. [Abstract] [CrossRef] [Google Scholar]
712. Barone A, Chi DC, Theoret MR, Chen H, He K, Kufrin D, Helms WS, Subramaniam S, Zhao H, Patel A, et al. FDA approval summary: trabectedin for unresectable or metastatic liposarcoma or leiomyosarcoma following an anthracycline-containing regimen. Clin Cancer Res. 2017;23(24):7448–7453. 10.1158/1078-0432.CCR-17-0898. [Abstract] [CrossRef] [Google Scholar]
713. Yoon SS, Segal NH, Park PJ, Detwiller KY, Fernando NT, Ryeom SW, Brennan MF, Singer S. Angiogenic profile of soft tissue sarcomas based on analysis of circulating factors and microarray gene expression. J Surg Res. 2006;135(2):282–290. 10.1016/j.jss.2006.01.023. [Abstract] [CrossRef] [Google Scholar]
714. Antoniou G, Lee ATJ, Huang PH, Jones RL. Olaratumab in soft tissue sarcoma-current status and future perspectives. Eur J Cancer. 2018;92:33–39. 10.1016/j.ejca.2017.12.026. [Abstract] [CrossRef] [Google Scholar]
715. van der Graaf WT. Olaratumab in soft-tissue sarcomas. Lancet. 2016;388(10043):442–444. 10.1016/S0140-6736(16)30788-7. [Abstract] [CrossRef] [Google Scholar]
716. Loizos N, Xu Y, Huber J, Liu M, Lu D, Finnerty B, Rolser R, Malikzay A, Persaud A, Corcoran E, et al. Targeting the platelet-derived growth factor receptor alpha with a neutralizing human monoclonal antibody inhibits the growth of tumor xenografts: implications as a potential therapeutic target. Mol Cancer Ther. 2005;4(3):369–379. 10.1158/1535-7163.MCT-04-0114. [Abstract] [CrossRef] [Google Scholar]
717. Tap WD, Jones RL, Van Tine BA, Chmielowski B, Elias AD, Adkins D, Agulnik M, Cooney MM, Livingston MB, Pennock G, et al. Olaratumab and doxorubicin versus doxorubicin alone for treatment of soft-tissue sarcoma: an open-label phase 1b and randomised phase 2 trial. Lancet. 2016;388(10043):488–497. 10.1016/S0140-6736(16)30587-6. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
718. Tap WD, Gelderblom H, Palmerini E, Desai J, Bauer S, Blay JY, Alcindor T, Ganjoo K, Martin-Broto J, Ryan CW, et al. Pexidartinib versus placebo for advanced tenosynovial giant cell tumour (ENLIVEN): a randomised phase 3 trial. Lancet. 2019;394(10197):478–487. 10.1016/S0140-6736(19)30764-0. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
719. West RB, Rubin BP, Miller MA, Subramanian S, Kaygusuz G, Montgomery K, Zhu S, Marinelli RJ, De Luca A, Downs-Kelly E, et al. A landscape effect in tenosynovial giant-cell tumor from activation of CSF1 expression by a translocation in a minority of tumor cells. Proc Natl Acad Sci U S A. 2006;103(3):690–695. 10.1073/pnas.0507321103. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
720. Guilbert LJ, Stanley ER. Specific interaction of murine colony-stimulating factor with mononuclear phagocytic cells. J Cell Biol. 1980;85(1):153–159. 10.1083/jcb.85.1.153. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
721. Shi G, Yang Q, Zhang Y, Jiang Q, Lin Y, Yang S, Wang H, Cheng L, Zhang X, Li Y, et al. Modulating the tumor microenvironment via oncolytic viruses and CSF-1R inhibition synergistically enhances anti-PD-1 immunotherapy. Mol Ther. 2019;27(1):244–260. 10.1016/j.ymthe.2018.11.010. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
722. Murga-Zamalloa C, Rolland DCM, Polk A, Wolfe A, Dewar H, Chowdhury P, Onder O, Dewar R, Brown NA, Bailey NG, et al. Colony-stimulating factor 1 receptor (CSF1R) activates AKT/mTOR signaling and promotes T-cell lymphoma viability. Clin Cancer Res. 2020;26(3):690–703. 10.1158/1078-0432.CCR-19-1486. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
723. Lamb YN. Pexidartinib: first approval. Drugs. 2019;79(16):1805–1812. 10.1007/s40265-019-01210-0. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
724. Tap WD, Wainberg ZA, Anthony SP, Ibrahim PN, Zhang C, Healey JH, Chmielowski B, Staddon AP, Cohn AL, Shapiro GI, et al. Structure-guided blockade of CSF1R kinase in tenosynovial giant-cell tumor. N Engl J Med. 2015;373(5):428–437. 10.1056/NEJMoa1411366. [Abstract] [CrossRef] [Google Scholar]
725. Smith CC, Zhang C, Lin KC, Lasater EA, Zhang Y, Massi E, Damon LE, Pendleton M, Bashir A, Sebra R, et al. Characterizing and overriding the structural mechanism of the quizartinib-resistant FLT3 "Gatekeeper" F691L mutation with PLX3397. Cancer Discov. 2015;5(6):668–679. 10.1158/2159-8290.CD-15-0060. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
726. Spillane AJ, Thomas JM, Fisher C. Epithelioid sarcoma: the clinicopathological complexities of this rare soft tissue sarcoma. Ann Surg Oncol. 2000;7(3):218–225. 10.1007/BF02523657. [Abstract] [CrossRef] [Google Scholar]
727. Rothbart SB, Baylin SB. Epigenetic therapy for epithelioid sarcoma. Cell. 2020;181(2):211. 10.1016/j.cell.2020.03.042. [Abstract] [CrossRef] [Google Scholar]
728. Modena P, Lualdi E, Facchinetti F, Galli L, Teixeira MR, Pilotti S, Sozzi G. SMARCB1/INI1 tumor suppressor gene is frequently inactivated in epithelioid sarcomas. Cancer Res. 2005;65(10):4012–4019. 10.1158/0008-5472.CAN-04-3050. [Abstract] [CrossRef] [Google Scholar]
729. Mittal P, Roberts CWM. The SWI/SNF complex in cancer - biology, biomarkers and therapy. Nat Rev Clin Oncol. 2020;17(7):435–448. 10.1038/s41571-020-0357-3. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
730. Papp G, Changchien YC, Peterfia B, Pecsenka L, Krausz T, Stricker TP, Khoor A, Donner L, Sapi Z. SMARCB1 protein and mRNA loss is not caused by promoter and histone hypermethylation in epithelioid sarcoma. Mod Pathol. 2013;26(3):393–403. 10.1038/modpathol.2012.190. [Abstract] [CrossRef] [Google Scholar]
731. Hoy SM. Tazemetostat: first approval. Drugs. 2020;80(5):513–521. 10.1007/s40265-020-01288-x. [Abstract] [CrossRef] [Google Scholar]
732. Stacchiotti S, Schoffski P, Jones R, Agulnik M, Villalobos VM, Jahan TM, Chen TW-W, Italiano A, Demetri GD, Cote GM et al. Safety and efficacy of tazemetostat, a first-in-class EZH2 inhibitor, in patients (pts) with epithelioid sarcoma (ES) (NCT02601950). J Clin Oncol 2019; 37(15_suppl):11003-11003.
733. Gounder M, Schoffski P, Jones RL, Agulnik M, Cote GM, Villalobos VM, Attia S, Chugh R, Chen TW, Jahan T, et al. Tazemetostat in advanced epithelioid sarcoma with loss of INI1/SMARCB1: an international, open-label, phase 2 basket study. Lancet Oncol. 2020;21(11):1423–1432. 10.1016/S1470-2045(20)30451-4. [Abstract] [CrossRef] [Google Scholar]
734. Miller KD, Ostrom QT, Kruchko C, Patil N, Tihan T, Cioffi G, Fuchs HE, Waite KA, Jemal A, Siegel RL, et al. Brain and other central nervous system tumor statistics, 2021. CA Cancer J Clin. 2021;71(5):381–406. 10.3322/caac.21693. [Abstract] [CrossRef] [Google Scholar]
735. Parsons DW, Jones S, Zhang X, Lin JC, Leary RJ, Angenendt P, Mankoo P, Carter H, Siu IM, Gallia GL, et al. An integrated genomic analysis of human glioblastoma multiforme. Science. 2008;321(5897):1807–1812. 10.1126/science.1164382. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
736. Rottenberg DA, Ginos JZ, Kearfott KJ, Junck L, Bigner DD. In vivo measurement of regional brain tissue pH using positron emission tomography. Ann Neurol. 1984;15(Suppl):S98–102. 10.1002/ana.410150718. [Abstract] [CrossRef] [Google Scholar]
737. Zhang J, Stevens MF, Bradshaw TD. Temozolomide: mechanisms of action, repair and resistance. Curr Mol Pharmacol. 2012;5(1):102–114. 10.2174/1874467211205010102. [Abstract] [CrossRef] [Google Scholar]
738. Abe H, Natsumeda M, Kanemaru Y, Watanabe J, Tsukamoto Y, Okada M, Yoshimura J, Oishi M, Fujii Y. MGMT expression contributes to temozolomide resistance in H3K27M-mutant diffuse midline gliomas and MGMT silencing to temozolomide sensitivity in IDH-mutant gliomas. Neurol Med Chir. 2018;58(7):290–295. 10.2176/nmc.ra.2018-0044. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
739. Karachi A, Dastmalchi F, Mitchell DA, Rahman M. Temozolomide for immunomodulation in the treatment of glioblastoma. Neuro Oncol. 2018;20(12):1566–1572. 10.1093/neuonc/noy072. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
740. Ochs K, Kaina B. Apoptosis induced by DNA damage O6-methylguanine is Bcl-2 and caspase-9/3 regulated and Fas/caspase-8 independent. Cancer Res. 2000;60(20):5815–5824. [Abstract] [Google Scholar]
741. Yung WK, Albright RE, Olson J, Fredericks R, Fink K, Prados MD, Brada M, Spence A, Hohl RJ, Shapiro W, et al. A phase II study of temozolomide vs. procarbazine in patients with glioblastoma multiforme at first relapse. Br J Cancer. 2000;83(5):588–593. 10.1054/bjoc.2000.1316. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
742. Stupp R, Dietrich PY, Ostermann Kraljevic S, Pica A, Maillard I, Maeder P, Meuli R, Janzer R, Pizzolato G, Miralbell R, et al. Promising survival for patients with newly diagnosed glioblastoma multiforme treated with concomitant radiation plus temozolomide followed by adjuvant temozolomide. J Clin Oncol. 2002;20(5):1375–1382. 10.1200/JCO.2002.20.5.1375. [Abstract] [CrossRef] [Google Scholar]
743. Chinot OL, Wick W, Mason W, Henriksson R, Saran F, Nishikawa R, Carpentier AF, Hoang-Xuan K, Kavan P, Cernea D, et al. Bevacizumab plus radiotherapy-temozolomide for newly diagnosed glioblastoma. N Engl J Med. 2014;370(8):709–722. 10.1056/NEJMoa1308345. [Abstract] [CrossRef] [Google Scholar]
744. Hegi ME, Diserens AC, Godard S, Dietrich PY, Regli L, Ostermann S, Otten P, Van Melle G, de Tribolet N, Stupp R. Clinical trial substantiates the predictive value of O-6-methylguanine-DNA methyltransferase promoter methylation in glioblastoma patients treated with temozolomide. Clin Cancer Res. 2004;10(6):1871–1874. 10.1158/1078-0432.CCR-03-0384. [Abstract] [CrossRef] [Google Scholar]
745. Hegi ME, Diserens AC, Gorlia T, Hamou MF, de Tribolet N, Weller M, Kros JM, Hainfellner JA, Mason W, Mariani L, et al. MGMT gene silencing and benefit from temozolomide in glioblastoma. N Engl J Med. 2005;352(10):997–1003. 10.1056/NEJMoa043331. [Abstract] [CrossRef] [Google Scholar]
746. Park JA, Cheung NV. Targets and antibody formats for immunotherapy of neuroblastoma. J Clin Oncol. 2020;38(16):1836–1848. 10.1200/JCO.19.01410. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
747. Brodeur GM. Neuroblastoma: biological insights into a clinical enigma. Nat Rev Cancer. 2003;3(3):203–216. 10.1038/nrc1014. [Abstract] [CrossRef] [Google Scholar]
748. Suzuki M, Cheung NK. Disialoganglioside GD2 as a therapeutic target for human diseases. Expert Opin Ther Targets. 2015;19(3):349–362. 10.1517/14728222.2014.986459. [Abstract] [CrossRef] [Google Scholar]
749. Maris JM. Recent advances in neuroblastoma. N Engl J Med. 2010;362(23):2202–2211. 10.1056/NEJMra0804577. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
750. Schulz G, Cheresh DA, Varki NM, Yu A, Staffileno LK, Reisfeld RA. Detection of ganglioside GD2 in tumor tissues and sera of neuroblastoma patients. Cancer Res. 1984;44(12 Pt 1):5914–5920. [Abstract] [Google Scholar]
751. Mujoo K, Cheresh DA, Yang HM, Reisfeld RA. Disialoganglioside GD2 on human neuroblastoma cells: target antigen for monoclonal antibody-mediated cytolysis and suppression of tumor growth. Cancer Res. 1987;47(4):1098–1104. [Abstract] [Google Scholar]
752. Horwacik I, Golik P, Grudnik P, Kolinski M, Zdzalik M, Rokita H, Dubin G. Structural basis of GD2 ganglioside and mimetic peptide recognition by 14G2a antibody. Mol Cell Proteomics. 2015;14(10):2577–2590. 10.1074/mcp.M115.052720. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
753. Dhillon S. Dinutuximab: first global approval. Drugs. 2015;75(8):923–927. 10.1007/s40265-015-0399-5. [Abstract] [CrossRef] [Google Scholar]
754. Yu AL, Gilman AL, Ozkaynak MF, London WB, Kreissman SG, Chen HX, Smith M, Anderson B, Villablanca JG, Matthay KK, et al. Anti-GD2 antibody with GM-CSF, interleukin-2, and isotretinoin for neuroblastoma. N Engl J Med. 2010;363(14):1324–1334. 10.1056/NEJMoa0911123. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
755. Cheung NK, Guo H, Hu J, Tassev DV, Cheung IY. Humanizing murine IgG3 anti-GD2 antibody m3F8 substantially improves antibody-dependent cell-mediated cytotoxicity while retaining targeting in vivo. Oncoimmunology. 2012;1(4):477–486. 10.4161/onci.19864. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
756. Kushner BH, Cheung IY, Modak S, Basu EM, Roberts SS, Cheung NK. Humanized 3F8 anti-GD2 monoclonal antibody dosing with granulocyte-macrophage colony-stimulating factor in patients with resistant neuroblastoma: a phase 1 Clinical trial. JAMA Oncol. 2018;4(12):1729–1735. 10.1001/jamaoncol.2018.4005. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
757. Mora J, Kushner BH, Flores MA, Santa-María V, Garraus M, Basu EM, Roberts SS, Castañeda A, Gorostegui M, Cheung N-KV, et al. Naxitamab-based chemoimmunotherapy for resistant high-risk neuroblastoma: Preliminary results of HITS pilot/phase II study. J Clin Oncol. 2019;37(15_suppl):10025–10025. 10.1200/JCO.2019.37.15_suppl.10025. [CrossRef] [Google Scholar]
758. Mora J, Chan GC, Morgenstern DA, Nysom K, Bear M, Dalby LW, Lisby S, Kushner BH. 75P Efficacy and updated safety results from pivotal phase II trial 201 of naxitamab (Hu3F8): A humanized GD2-targeted immunotherapy for the treatment of refractory/relapsed (R/R) high-risk (HR) neuroblastoma (NB) Annals Oncol. 2020 10.1016/j.annonc.2020.10.563. [CrossRef] [Google Scholar]
759. Janes SM, Alrifai D, Fennell DA. Perspectives on the treatment of malignant pleural mesothelioma. N Engl J Med. 2021;385(13):1207–1218. 10.1056/NEJMra1912719. [Abstract] [CrossRef] [Google Scholar]
760. Milano MT, Zhang H. Malignant pleural mesothelioma: a population-based study of survival. J Thorac Oncol. 2010;5(11):1841–1848. 10.1097/JTO.0b013e3181f1cf2b. [Abstract] [CrossRef] [Google Scholar]
761. Chen C, Ke J, Zhou XE, Yi W, Brunzelle JS, Li J, Yong EL, Xu HE, Melcher K. Structural basis for molecular recognition of folic acid by folate receptors. Nature. 2013;500(7463):486–489. 10.1038/nature12327. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
762. Macfarlane AJ, Perry CA, McEntee MF, Lin DM, Stover PJ. Shmt1 heterozygosity impairs folate-dependent thymidylate synthesis capacity and modifies risk of Apc(min)-mediated intestinal cancer risk. Cancer Res. 2011;71(6):2098–2107. 10.1158/0008-5472.CAN-10-1886. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
763. Bueno R, Appasani K, Mercer H, Lester S, Sugarbaker D. The alpha folate receptor is highly activated in malignant pleural mesothelioma. J Thorac Cardiovasc Surg. 2001;121(2):225–233. 10.1067/mtc.2001.111176. [Abstract] [CrossRef] [Google Scholar]
764. Adjei AA. Pemetrexed: a multitargeted antifolate agent with promising activity in solid tumors. Ann Oncol. 2000;11(10):1335–1341. 10.1023/A:1008379101017. [Abstract] [CrossRef] [Google Scholar]
765. Hazarika M, White RM, Jr, Booth BP, Wang YC, Ham DY, Liang CY, Rahman A, Gobburu JV, Li N, Sridhara R, et al. Pemetrexed in malignant pleural mesothelioma. Clin Cancer Res. 2005;11(3):982–992. 10.1158/1078-0432.982.11.3. [Abstract] [CrossRef] [Google Scholar]
766. Forde PM, Anagnostou V, Sun Z, Dahlberg SE, Kindler HL, Niknafs N, Purcell T, Santana-Davila R, Dudek AZ, Borghaei H, et al. Durvalumab with platinum-pemetrexed for unresectable pleural mesothelioma: survival, genomic and immunologic analyses from the phase 2 PrE0505 trial. Nat Med. 2021 10.1038/s41591-021-01541-0. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
767. Amatu A, Sartore-Bianchi A, Siena S. NTRK gene fusions as novel targets of cancer therapy across multiple tumour types. ESMO Open. 2016;1(2):e000023. 10.1136/esmoopen-2015-000023. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
768. Doebele RC, Drilon A, Paz-Ares L, Siena S, Shaw AT, Farago AF, Blakely CM, Seto T, Cho BC, Tosi D, et al. Entrectinib in patients with advanced or metastatic NTRK fusion-positive solid tumours: integrated analysis of three phase 1–2 trials. Lancet Oncol. 2020;21(2):271–282. 10.1016/S1470-2045(19)30691-6. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
769. Bhangoo MS, Sigal D. TRK inhibitors: clinical development of larotrectinib. Curr Oncol Rep. 2019;21(2):14. 10.1007/s11912-019-0761-y. [Abstract] [CrossRef] [Google Scholar]
770. Scott LJ. Larotrectinib: first global approval. Drugs. 2019;79(2):201–206. 10.1007/s40265-018-1044-x. [Abstract] [CrossRef] [Google Scholar]
771. Doebele RC, Davis LE, Vaishnavi A, Le AT, Estrada-Bernal A, Keysar S, Jimeno A, Varella-Garcia M, Aisner DL, Li Y, et al. An oncogenic NTRK fusion in a patient with soft-tissue sarcoma with response to the tropomyosin-related kinase inhibitor LOXO-101. Cancer Discov. 2015;5(10):1049–1057. 10.1158/2159-8290.CD-15-0443. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
772. Drilon A, Laetsch TW, Kummar S, DuBois SG, Lassen UN, Demetri GD, Nathenson M, Doebele RC, Farago AF, Pappo AS, et al. Efficacy of larotrectinib in TRK fusion-positive cancers in adults and children. N Engl J Med. 2018;378(8):731–739. 10.1056/NEJMoa1714448. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
773. Hong DS, DuBois SG, Kummar S, Farago AF, Albert CM, Rohrberg KS, van Tilburg CM, Nagasubramanian R, Berlin JD, Federman N, et al. Larotrectinib in patients with TRK fusion-positive solid tumours: a pooled analysis of three phase 1/2 clinical trials. Lancet Oncol. 2020;21(4):531–540. 10.1016/S1470-2045(19)30856-3. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
774. Fuse MJ, Okada K, Oh-Hara T, Ogura H, Fujita N, Katayama R. Mechanisms of resistance to NTRK inhibitors and therapeutic strategies in NTRK1-rearranged cancers. Mol Cancer Ther. 2017;16(10):2130–2143. 10.1158/1535-7163.MCT-16-0909. [Abstract] [CrossRef] [Google Scholar]
775. Drilon A, Zhai D, Deng W, Zhang X, Lee D, Rogers E, Whitten J, Huang Z, Graber A, Liu J et al: Abstract 442: Repotrectinib, a next generation TRK inhibitor, overcomes TRK resistance mutations including solvent front, gatekeeper and compound mutations. In: Clinical Research (Excluding Clinical Trials). 2019; 442–442.
776. Lemmon MA, Schlessinger J. Cell signaling by receptor tyrosine kinases. Cell. 2010;141(7):1117–1134. 10.1016/j.cell.2010.06.011. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
777. Berns K, Horlings HM, Hennessy BT, Madiredjo M, Hijmans EM, Beelen K, Linn SC, Gonzalez-Angulo AM, Stemke-Hale K, Hauptmann M, et al. A functional genetic approach identifies the PI3K pathway as a major determinant of trastuzumab resistance in breast cancer. Cancer Cell. 2007;12(4):395–402. 10.1016/j.ccr.2007.08.030. [Abstract] [CrossRef] [Google Scholar]
778. Copin MC, Lesaffre M, Berbon M, Doublet L, Leroy C, Tresch E, Porte H, Vicogne J, Cortot AB, Dansin E, et al. High-MET status in non-small cell lung tumors correlates with receptor phosphorylation but not with the serum level of soluble form. Lung Cancer. 2016;101:59–67. 10.1016/j.lungcan.2016.09.009. [Abstract] [CrossRef] [Google Scholar]
779. Slastnikova TA, Ulasov AV, Rosenkranz AA, Sobolev AS. Targeted intracellular delivery of antibodies: the state of the art. Front Pharmacol. 2018;9:1208. 10.3389/fphar.2018.01208. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
780. Razavi P, Dickler MN, Shah PD, Toy W, Brown DN, Won HH, Li BT, Shen R, Vasan N, Modi S, et al. Alterations in PTEN and ESR1 promote clinical resistance to alpelisib plus aromatase inhibitors. Nat Cancer. 2020;1(4):382–393. 10.1038/s43018-020-0047-1. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
781. Huang AC, Zappasodi R. A decade of checkpoint blockade immunotherapy in melanoma: understanding the molecular basis for immune sensitivity and resistance. Nat Immunol. 2022;23(5):660–670. 10.1038/s41590-022-01141-1. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
782. Johnson DB, Nebhan CA, Moslehi JJ, Balko JM. Immune-checkpoint inhibitors: long-term implications of toxicity. Nat Rev Clin Oncol. 2022;19(4):254–267. 10.1038/s41571-022-00600-w. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
783. Morad G, Helmink BA, Sharma P, Wargo JA. Hallmarks of response, resistance, and toxicity to immune checkpoint blockade. Cell. 2021;184(21):5309–5337. 10.1016/j.cell.2021.09.020. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
784. Chen DS, Mellman I. Elements of cancer immunity and the cancer-immune set point. Nature. 2017;541(7637):321–330. 10.1038/nature21349. [Abstract] [CrossRef] [Google Scholar]
785. Schumacher TN, Schreiber RD. Neoantigens in cancer immunotherapy. Science. 2015;348(6230):69–74. 10.1126/science.aaa4971. [Abstract] [CrossRef] [Google Scholar]
786. Herbst RS, Soria JC, Kowanetz M, Fine GD, Hamid O, Gordon MS, Sosman JA, McDermott DF, Powderly JD, Gettinger SN, et al. Predictive correlates of response to the anti-PD-L1 antibody MPDL3280A in cancer patients. Nature. 2014;515(7528):563–567. 10.1038/nature14011. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
787. Tumeh PC, Harview CL, Yearley JH, Shintaku IP, Taylor EJ, Robert L, Chmielowski B, Spasic M, Henry G, Ciobanu V, et al. PD-1 blockade induces responses by inhibiting adaptive immune resistance. Nature. 2014;515(7528):568–571. 10.1038/nature13954. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
788. Jing Y, Yang J, Johnson DB, Moslehi JJ, Han L. Harnessing big data to characterize immune-related adverse events. Nat Rev Clin Oncol. 2022;19(4):269–280. 10.1038/s41571-021-00597-8. [Abstract] [CrossRef] [Google Scholar]
789. Ferrara R, Mezquita L, Texier M, Lahmar J, Audigier-Valette C, Tessonnier L, Mazieres J, Zalcman G, Brosseau S, Le Moulec S, et al. Hyperprogressive disease in patients with advanced non-small cell lung cancer treated with PD-1/PD-L1 inhibitors or with single-agent chemotherapy. JAMA Oncol. 2018;4(11):1543–1552. 10.1001/jamaoncol.2018.3676. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
790. Karlsson P, Sun Z, Braun D, Price KN, Castiglione-Gertsch M, Rabaglio M, Gelber RD, Crivellari D, Collins J, Murray E, et al. Long-term results of International Breast Cancer Study Group Trial VIII: adjuvant chemotherapy plus goserelin compared with either therapy alone for premenopausal patients with node-negative breast cancer. Ann Oncol. 2011;22(10):2216–2226. 10.1093/annonc/mdq735. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
791. Halabi S, Lin CY, Kelly WK, Fizazi KS, Moul JW, Kaplan EB, Morris MJ, Small EJ. Updated prognostic model for predicting overall survival in first-line chemotherapy for patients with metastatic castration-resistant prostate cancer. J Clin Oncol. 2014;32(7):671–677. 10.1200/JCO.2013.52.3696. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
792. Beer TM, Armstrong AJ, Rathkopf D, Loriot Y, Sternberg CN, Higano CS, Iversen P, Evans CP, Kim CS, Kimura G, et al. Enzalutamide in men with chemotherapy-naive metastatic castration-resistant prostate cancer: extended analysis of the phase 3 PREVAIL study. Eur Urol. 2017;71(2):151–154. 10.1016/j.eururo.2016.07.032. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
793. Teply BA, Wang H, Luber B, Sullivan R, Rifkind I, Bruns A, Spitz A, DeCarli M, Sinibaldi V, Pratz CF, et al. Bipolar androgen therapy in men with metastatic castration-resistant prostate cancer after progression on enzalutamide: an open-label, phase 2, multicohort study. Lancet Oncol. 2018;19(1):76–86. 10.1016/S1470-2045(17)30906-3. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
794. Markowski MC, Wang H, Sullivan R, Rifkind I, Sinibaldi V, Schweizer MT, Teply BA, Ngomba N, Fu W, Carducci MA, et al. A multicohort open-label phase II trial of bipolar androgen therapy in men with metastatic castration-resistant prostate cancer (RESTORE): a comparison of post-abiraterone versus post-enzalutamide cohorts. Eur Urol. 2021;79(5):692–699. 10.1016/j.eururo.2020.06.042. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
795. Paik J. Nivolumab plus relatlimab: first approval. Drugs. 2022;82(8):925–931. 10.1007/s40265-022-01723-1. [Abstract] [CrossRef] [Google Scholar]
796. Sakamoto KM, Kim KB, Kumagai A, Mercurio F, Crews CM, Deshaies RJ. Protacs: chimeric molecules that target proteins to the Skp1-Cullin-F box complex for ubiquitination and degradation. Proc Natl Acad Sci U S A. 2001;98(15):8554–8559. 10.1073/pnas.141230798. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
797. Paul CP, Good PD, Winer I, Engelke DR. Effective expression of small interfering RNA in human cells. Nat Biotechnol. 2002;20(5):505–508. 10.1038/nbt0502-505. [Abstract] [CrossRef] [Google Scholar]
798. Burslem GM, Crews CM. Proteolysis-targeting chimeras as therapeutics and tools for biological discovery. Cell. 2020;181(1):102–114. 10.1016/j.cell.2019.11.031. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
799. Lamb YN. Inclisiran: first approval. Drugs. 2021;81(3):389–395. 10.1007/s40265-021-01473-6. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
800. Garber K. The PROTAC gold rush. Nat Biotechnol. 2022;40(1):12–16. 10.1038/s41587-021-01173-2. [Abstract] [CrossRef] [Google Scholar]
801. Fan X, Jin WY, Lu J, Wang J, Wang YT. Rapid and reversible knockdown of endogenous proteins by peptide-directed lysosomal degradation. Nat Neurosci. 2014;17(3):471–480. 10.1038/nn.3637. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
802. Clift D, McEwan WA, Labzin LI, Konieczny V, Mogessie B, James LC, Schuh M. A method for the acute and rapid degradation of endogenous proteins. Cell. 2017;171(7):1692–1706 e1618. 10.1016/j.cell.2017.10.033. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
803. Nabet B, Roberts JM, Buckley DL, Paulk J, Dastjerdi S, Yang A, Leggett AL, Erb MA, Lawlor MA, Souza A, et al. The dTAG system for immediate and target-specific protein degradation. Nat Chem Biol. 2018;14(5):431–441. 10.1038/s41589-018-0021-8. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
804. Banik SM, Pedram K, Wisnovsky S, Ahn G, Riley NM, Bertozzi CR. Lysosome-targeting chimaeras for degradation of extracellular proteins. Nature. 2020;584(7820):291–297. 10.1038/s41586-020-2545-9. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
805. Elshiaty M, Schindler H, Christopoulos P. Principles and current clinical landscape of multispecific antibodies against cancer. Int J Mol Sci. 2021;22(11):5632. 10.3390/ijms22115632. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
806. June CH, Sadelain M. Chimeric antigen receptor therapy. N Engl J Med. 2018;379(1):64–73. 10.1056/NEJMra1706169. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
807. The Lancet O. CAR T-cell therapy for solid tumours. Lancet Oncol. 2021;22(7):893. 10.1016/S1470-2045(21)00353-3. [Abstract] [CrossRef] [Google Scholar]
808. Yarmarkovich M, Marshall QF, Warrington JM, Premaratne R, Farrel A, Groff D, Li W, di Marco M, Runbeck E, Truong H, et al. Cross-HLA targeting of intracellular oncoproteins with peptide-centric CARs. Nature. 2021;599(7885):477–484. 10.1038/s41586-021-04061-6. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar] Retracted
809. Wei R, Zhou Y, Li C, Rychahou P, Zhang S, Titlow WB, Bauman G, Wu Y, Liu J, Wang C, et al. Ketogenesis attenuates KLF5-dependent production of CXCL12 to overcome the immunosuppressive tumor microenvironment in colorectal cancer. Cancer Res. 2022;82(8):1575–1588. 10.1158/0008-5472.CAN-21-2778. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
810. Norelli M, Camisa B, Barbiera G, Falcone L, Purevdorj A, Genua M, Sanvito F, Ponzoni M, Doglioni C, Cristofori P, et al. Monocyte-derived IL-1 and IL-6 are differentially required for cytokine-release syndrome and neurotoxicity due to CAR T cells. Nat Med. 2018;24(6):739–748. 10.1038/s41591-018-0036-4. [Abstract] [CrossRef] [Google Scholar]
811. Sterner RM, Sakemura R, Cox MJ, Yang N, Khadka RH, Forsman CL, Hansen MJ, Jin F, Ayasoufi K, Hefazi M, et al. GM-CSF inhibition reduces cytokine release syndrome and neuroinflammation but enhances CAR-T cell function in xenografts. Blood. 2019;133(7):697–709. 10.1182/blood-2018-10-881722. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
812. Brown CE, Alizadeh D, Starr R, Weng L, Wagner JR, Naranjo A, Ostberg JR, Blanchard MS, Kilpatrick J, Simpson J, et al. Regression of glioblastoma after chimeric antigen receptor T-cell therapy. N Engl J Med. 2016;375(26):2561–2569. 10.1056/NEJMoa1610497. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
813. Vitanza NA, Johnson AJ, Wilson AL, Brown C, Yokoyama JK, Kunkele A, Chang CA, Rawlings-Rhea S, Huang W, Seidel K, et al. Locoregional infusion of HER2-specific CAR T cells in children and young adults with recurrent or refractory CNS tumors: an interim analysis. Nat Med. 2021;27(9):1544–1552. 10.1038/s41591-021-01404-8. [Abstract] [CrossRef] [Google Scholar]
814. Adusumilli PS, Zauderer MG, Riviere I, Solomon SB, Rusch VW, O'Cearbhaill RE, Zhu A, Cheema W, Chintala NK, Halton E, et al. A phase I trial of regional mesothelin-targeted CAR T-cell therapy in patients with malignant pleural disease, in combination with the anti-PD-1 agent pembrolizumab. Cancer Discov. 2021;11(11):2748–2763. 10.1158/2159-8290.CD-21-0407. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
815. Narayan V, Barber-Rotenberg JS, Jung IY, Lacey SF, Rech AJ, Davis MM, Hwang WT, Lal P, Carpenter EL, Maude SL, et al. PSMA-targeting TGFbeta-insensitive armored CAR T cells in metastatic castration-resistant prostate cancer: a phase 1 trial. Nat Med. 2022;28(4):724–734. 10.1038/s41591-022-01726-1. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
816. Kaufman HL, Kohlhapp FJ, Zloza A. Oncolytic viruses: a new class of immunotherapy drugs. Nat Rev Drug Discov. 2015;14(9):642–662. 10.1038/nrd4663. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
817. Melcher A, Harrington K, Vile R. Oncolytic virotherapy as immunotherapy. Science. 2021;374(6573):1325–1326. 10.1126/science.abk3436. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
818. Dummer R, Gyorki DE, Hyngstrom J, Berger AC, Conry R, Demidov L, Sharma A, Treichel SA, Radcliffe H, Gorski KS, et al. Neoadjuvant talimogene laherparepvec plus surgery versus surgery alone for resectable stage IIIB-IVM1a melanoma: a randomized, open-label, phase 2 trial. Nat Med. 2021;27(10):1789–1796. 10.1038/s41591-021-01510-7. [Abstract] [CrossRef] [Google Scholar]
819. Cloughesy TF, Petrecca K, Walbert T, Butowski N, Salacz M, Perry J, Damek D, Bota D, Bettegowda C, Zhu JJ, et al. Effect of vocimagene amiretrorepvec in combination with flucytosine vs standard of care on survival following tumor resection in patients with recurrent high-grade glioma: a randomized clinical trial. JAMA Oncol. 2020;6(12):1939–1946. 10.1001/jamaoncol.2020.3161. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
820. Mahalingam D, Wilkinson GA, Eng KH, Fields P, Raber P, Moseley JL, Cheetham K, Coffey M, Nuovo G, Kalinski P, et al. Pembrolizumab in combination with the oncolytic virus pelareorep and chemotherapy in patients with advanced pancreatic adenocarcinoma: a phase Ib study. Clin Cancer Res. 2020;26(1):71–81. 10.1158/1078-0432.CCR-19-2078. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
821. Ribas A, Dummer R, Puzanov I, VanderWalde A, Andtbacka RHI, Michielin O, Olszanski AJ, Malvehy J, Cebon J, Fernandez E, et al. Oncolytic virotherapy promotes intratumoral T cell infiltration and improves anti-PD-1 immunotherapy. Cell. 2017;170(6):1109–1119. 10.1016/j.cell.2017.08.027. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
822. Kelly CM, Antonescu CR, Bowler T, Munhoz R, Chi P, Dickson MA, Gounder MM, Keohan ML, Movva S, Dholakia R, et al. Objective response rate among patients with locally advanced or metastatic sarcoma treated with talimogene laherparepvec in combination with pembrolizumab: a phase 2 clinical trial. JAMA Oncol. 2020;6(3):402–408. 10.1001/jamaoncol.2019.6152. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
823. Bommareddy PK, Shettigar M, Kaufman HL. Integrating oncolytic viruses in combination cancer immunotherapy. Nat Rev Immunol. 2018;18(8):498–513. 10.1038/s41577-018-0014-6. [Abstract] [CrossRef] [Google Scholar]
824. Greig SL. Talimogene laherparepvec: first global approval. Drugs. 2016;76(1):147–154. 10.1007/s40265-015-0522-7. [Abstract] [CrossRef] [Google Scholar]
825. Russell SJ, Barber GN. Oncolytic viruses as antigen-agnostic cancer vaccines. Cancer Cell. 2018;33(4):599–605. 10.1016/j.ccell.2018.03.011. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
826. Phelps MP, Yang H, Patel S, Rahman MM, McFadden G, Chen E. Oncolytic virus-mediated RAS targeting in rhabdomyosarcoma. Mol Ther Oncolytics. 2018;11:52–61. 10.1016/j.omto.2018.09.001. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
827. Pallerla S, Abdul A, Comeau J, Jois S. Cancer vaccines, treatment of the future: with emphasis on her2-positive breast cancer. Int J Mol Sci. 2021;22(2):12213. 10.3390/ijms22020779. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
828. Saxena M, van der Burg SH, Melief CJM, Bhardwaj N. Therapeutic cancer vaccines. Nat Rev Cancer. 2021;21(6):360–378. 10.1038/s41568-021-00346-0. [Abstract] [CrossRef] [Google Scholar]
829. Roden RBS, Stern PL. Opportunities and challenges for human papillomavirus vaccination in cancer. Nat Rev Cancer. 2018;18(4):240–254. 10.1038/nrc.2018.13. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
830. Kao JH. Hepatitis B vaccination and prevention of hepatocellular carcinoma. Best Pract Res Clin Gastroenterol. 2015;29(6):907–917. 10.1016/j.bpg.2015.09.011. [Abstract] [CrossRef] [Google Scholar]
831. Tsao AS, Kim ES, Hong WK. Chemoprevention of cancer. CA Cancer J Clin. 2004;54(3):150–180. 10.3322/canjclin.54.3.150. [Abstract] [CrossRef] [Google Scholar]
832. Dejea CM, Fathi P, Craig JM, Boleij A, Taddese R, Geis AL, Wu X, DeStefano Shields CE, Hechenbleikner EM, Huso DL, et al. Patients with familial adenomatous polyposis harbor colonic biofilms containing tumorigenic bacteria. Science. 2018;359(6375):592–597. 10.1126/science.aah3648. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
833. Groden J, Thliveris A, Samowitz W, Carlson M, Gelbert L, Albertsen H, Joslyn G, Stevens J, Spirio L, Robertson M, et al. Identification and characterization of the familial adenomatous polyposis coli gene. Cell. 1991;66(3):589–600. 10.1016/0092-8674(81)90021-0. [Abstract] [CrossRef] [Google Scholar]
834. Vasen HF, Moslein G, Alonso A, Aretz S, Bernstein I, Bertario L, Blanco I, Bulow S, Burn J, Capella G, et al. Guidelines for the clinical management of familial adenomatous polyposis (FAP) Gut. 2008;57(5):704–713. 10.1136/gut.2007.136127. [Abstract] [CrossRef] [Google Scholar]
835. Smith KJ, Johnson KA, Bryan TM, Hill DE, Markowitz S, Willson JK, Paraskeva C, Petersen GM, Hamilton SR, Vogelstein B, et al. The APC gene product in normal and tumor cells. Proc Natl Acad Sci U S A. 1993;90(7):2846–2850. 10.1073/pnas.90.7.2846. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
836. Nagase H, Miyoshi Y, Horii A, Aoki T, Ogawa M, Utsunomiya J, Baba S, Sasazuki T, Nakamura Y. Correlation between the location of germ-line mutations in the APC gene and the number of colorectal polyps in familial adenomatous polyposis patients. Cancer Res. 1992;52(14):4055–4057. [Abstract] [Google Scholar]
837. Oshima M, Oshima H, Kitagawa K, Kobayashi M, Itakura C, Taketo M. Loss of Apc heterozygosity and abnormal tissue building in nascent intestinal polyps in mice carrying a truncated Apc gene. Proc Natl Acad Sci U S A. 1995;92(10):4482–4486. 10.1073/pnas.92.10.4482. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
838. Bienz M, Clevers H. Linking colorectal cancer to Wnt signaling. Cell. 2000;103(2):311–320. 10.1016/S0092-8674(00)00122-7. [Abstract] [CrossRef] [Google Scholar]
839. Oshima M, Dinchuk JE, Kargman SL, Oshima H, Hancock B, Kwong E, Trzaskos JM, Evans JF, Taketo MM. Suppression of intestinal polyposis in Apc delta716 knockout mice by inhibition of cyclooxygenase 2 (COX-2) Cell. 1996;87(5):803–809. 10.1016/S0092-8674(00)81988-1. [Abstract] [CrossRef] [Google Scholar]
840. Castellone MD, Teramoto H, Williams BO, Druey KM, Gutkind JS. Prostaglandin E2 promotes colon cancer cell growth through a Gs-axin-beta-catenin signaling axis. Science. 2005;310(5753):1504–1510. 10.1126/science.1116221. [Abstract] [CrossRef] [Google Scholar]
841. Bresalier RS. Prevention of colorectal cancer: tumor progression, chemoprevention, and COX-2 inhibition. Gastroenterology. 2000;119(1):267–268. 10.1016/S0016-5085(00)70061-8. [Abstract] [CrossRef] [Google Scholar]
842. Bertagnolli MM. Chemoprevention of colorectal cancer with cyclooxygenase-2 inhibitors: two steps forward, one step back. Lancet Oncol. 2007;8(5):439–443. 10.1016/S1470-2045(07)70139-0. [Abstract] [CrossRef] [Google Scholar]
843. Peng F, Liao M, Qin R, Zhu S, Peng C, Fu L, Chen Y, Han B. Regulated cell death (RCD) in cancer: key pathways and targeted therapies. Signal Transduct Target Ther. 2022;7(1):286. 10.1038/s41392-022-01110-y. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
844. Phillips DC, Buchanan FG, Cheng D, Solomon LR, Xiao Y, Xue J, Tahir SK, Smith ML, Zhang H, Widomski D, et al. Hexavalent TRAIL fusion protein eftozanermin alfa optimally clusters apoptosis-inducing TRAIL receptors to induce on-target antitumor activity in solid tumors. Cancer Res. 2021;81(12):3402–3414. 10.1158/0008-5472.CAN-20-2178. [Abstract] [CrossRef] [Google Scholar]
845. Ishizawa J, Zarabi SF, Davis RE, Halgas O, Nii T, Jitkova Y, Zhao R, St-Germain J, Heese LE, Egan G, et al. Mitochondrial ClpP-mediated proteolysis induces selective cancer cell lethality. Cancer Cell. 2019;35(5):721–737 e729. 10.1016/j.ccell.2019.03.014. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
846. Morel D, Jeffery D, Aspeslagh S, Almouzni G, Postel-Vinay S. Combining epigenetic drugs with other therapies for solid tumours - past lessons and future promise. Nat Rev Clin Oncol. 2020;17(2):91–107. 10.1038/s41571-019-0267-4. [Abstract] [CrossRef] [Google Scholar]
847. Moreno V, Sepulveda JM, Vieito M, Hernandez-Guerrero T, Doger B, Saavedra O, Ferrero O, Sarmiento R, Arias M, De Alvaro J, et al. Phase I study of CC-90010, a reversible, oral BET inhibitor in patients with advanced solid tumors and relapsed/refractory non-Hodgkin's lymphoma. Ann Oncol. 2020;31(6):780–788. 10.1016/j.annonc.2020.03.294. [Abstract] [CrossRef] [Google Scholar]
848. Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, Tunyasuvunakool K, Bates R, Zidek A, Potapenko A, et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596(7873):583–589. 10.1038/s41586-021-03819-2. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
849. Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, Charpentier E. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science. 2012;337(6096):816–821. 10.1126/science.1225829. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
850. Katti A, Diaz BJ, Caragine CM, Sanjana NE, Dow LE. CRISPR in cancer biology and therapy. Nat Rev Cancer. 2022;22(5):259–279. 10.1038/s41568-022-00441-w. [Abstract] [CrossRef] [Google Scholar]
851. Stadtmauer EA, Fraietta JA, Davis MM, Cohen AD, Weber KL, Lancaster E, Mangan PA, Kulikovskaya I, Gupta M, Chen F, et al. CRISPR-engineered T cells in patients with refractory cancer. Science. 2020;367(6481):eaba7365. 10.1126/science.aba7365. [Abstract] [CrossRef] [Google Scholar]
852. Corsten MF, Shah K. Therapeutic stem-cells for cancer treatment: hopes and hurdles in tactical warfare. Lancet Oncol. 2008;9(4):376–384. 10.1016/S1470-2045(08)70099-8. [Abstract] [CrossRef] [Google Scholar]
853. Shi Y, Du L, Lin L, Wang Y. Tumour-associated mesenchymal stem/stromal cells: emerging therapeutic targets. Nat Rev Drug Discov. 2017;16(1):35–52. 10.1038/nrd.2016.193. [Abstract] [CrossRef] [Google Scholar]
854. Zhang C, Hu Y, Xiao W, Tian Z. Chimeric antigen receptor- and natural killer cell receptor-engineered innate killer cells in cancer immunotherapy. Cell Mol Immunol. 2021;18(9):2083–2100. 10.1038/s41423-021-00732-6. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
855. Tahara H, Zeh HJ, 3rd, Storkus WJ, Pappo I, Watkins SC, Gubler U, Wolf SF, Robbins PD, Lotze MT. Fibroblasts genetically engineered to secrete interleukin 12 can suppress tumor growth and induce antitumor immunity to a murine melanoma in vivo. Cancer Res. 1994;54(1):182–189. [Abstract] [Google Scholar]
856. Reinshagen C, Bhere D, Choi SH, Hutten S, Nesterenko I, Wakimoto H, Le Roux E, Rizvi A, Du W, Minicucci C, et al. CRISPR-enhanced engineering of therapy-sensitive cancer cells for self-targeting of primary and metastatic tumors. Sci Transl Med. 2018;10(449):eaao3240. 10.1126/scitranslmed.aao3240. [Abstract] [CrossRef] [Google Scholar]
857. Bashor CJ, Hilton IB, Bandukwala H, Smith DM, Veiseh O. Engineering the next generation of cell-based therapeutics. Nat Rev Drug Discov. 2022 10.1038/s41573-022-00476-6. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
858. Cullin N, Azevedo Antunes C, Straussman R, Stein-Thoeringer CK, Elinav E. Microbiome and cancer. Cancer Cell. 2021;39(10):1317–1341. 10.1016/j.ccell.2021.08.006. [Abstract] [CrossRef] [Google Scholar]
859. Poore GD, Kopylova E, Zhu Q, Carpenter C, Fraraccio S, Wandro S, Kosciolek T, Janssen S, Metcalf J, Song SJ, et al. Microbiome analyses of blood and tissues suggest cancer diagnostic approach. Nature. 2020;579(7800):567–574. 10.1038/s41586-020-2095-1. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
860. Helmink BA, Khan MAW, Hermann A, Gopalakrishnan V, Wargo JA. The microbiome, cancer, and cancer therapy. Nat Med. 2019;25(3):377–388. 10.1038/s41591-019-0377-7. [Abstract] [CrossRef] [Google Scholar]
861. Lim ZF, Ma PC. Emerging insights of tumor heterogeneity and drug resistance mechanisms in lung cancer targeted therapy. J Hematol Oncol. 2019;12(1):134. 10.1186/s13045-019-0818-2. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
862. McGrail DJ, Pilie PG, Rashid NU, Voorwerk L, Slagter M, Kok M, Jonasch E, Khasraw M, Heimberger AB, Lim B, et al. High tumor mutation burden fails to predict immune checkpoint blockade response across all cancer types. Ann Oncol. 2021;32(5):661–672. 10.1016/j.annonc.2021.02.006. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
863. Lim B, Woodward WA, Wang X, Reuben JM, Ueno NT. Inflammatory breast cancer biology: the tumour microenvironment is key. Nat Rev Cancer. 2018;18(8):485–499. 10.1038/s41568-018-0010-y. [Abstract] [CrossRef] [Google Scholar]
864. Karnezis AN, Cho KR, Gilks CB, Pearce CL, Huntsman DG. The disparate origins of ovarian cancers: pathogenesis and prevention strategies. Nat Rev Cancer. 2017;17(1):65–74. 10.1038/nrc.2016.113. [Abstract] [CrossRef] [Google Scholar]
865. Marquardt JU, Andersen JB, Thorgeirsson SS. Functional and genetic deconstruction of the cellular origin in liver cancer. Nat Rev Cancer. 2015;15(11):653–667. 10.1038/nrc4017. [Abstract] [CrossRef] [Google Scholar]
866. Tilg H, Zmora N, Adolph TE, Elinav E. The intestinal microbiota fuelling metabolic inflammation. Nat Rev Immunol. 2020;20(1):40–54. 10.1038/s41577-019-0198-4. [Abstract] [CrossRef] [Google Scholar]
867. Wagenlehner FME, Bjerklund Johansen TE, Cai T, Koves B, Kranz J, Pilatz A, Tandogdu Z. Epidemiology, definition and treatment of complicated urinary tract infections. Nat Rev Urol. 2020;17(10):586–600. 10.1038/s41585-020-0362-4. [Abstract] [CrossRef] [Google Scholar]
868. Ho AW, Kupper TS. T cells and the skin: from protective immunity to inflammatory skin disorders. Nat Rev Immunol. 2019;19(8):490–502. 10.1038/s41577-019-0162-3. [Abstract] [CrossRef] [Google Scholar]

Articles from Journal of Hematology & Oncology are provided here courtesy of BMC

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/136961253
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/136961253

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.1186/s13045-022-01362-9

Supporting
Mentioning
Contrasting
0
15
0

Article citations


Go to all (46) article citations

Data 


Data behind the article

This data has been text mined from the article, or deposited into data resources.

Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.


Funding 


Funders who supported this work.

Basic Research Project of Hangzhou Medical College (1)

Zhejiang Provincial Natural Science Foundation of China (1)