Skip to main content
Therapeutic Advances in Medical Oncology logoLink to Therapeutic Advances in Medical Oncology
. 2015 Sep;7(5):274–290. doi: 10.1177/1758834015590593

Treating patients with ALK-positive non-small cell lung cancer: latest evidence and management strategy

Bin-Chi Liao 1, Chia-Chi Lin 2,, Jin-Yuan Shih 3, James Chih-Hsin Yang 4
PMCID: PMC4543853  PMID: 26327925

Abstract

Rearrangements in anaplastic lymphoma kinase (ALK) gene and echinoderm microtubule-associated protein-like 4 (EML4) gene were first described in a small portion of patients with non-small cell lung cancer (NSCLC) in 2007. Fluorescence in situ hybridization is used as the diagnostic test for detecting an EML4–ALK rearrangement. Crizotinib, an ALK inhibitor, is effective in treating advanced ALK-positive NSCLC, and the US Food and Drug Administration approved it for treating ALK-positive NSCLC in 2011. Several mechanisms of acquired resistance to crizotinib have recently been reported. Second-generation ALK inhibitors were designed to overcome these resistance mechanisms. Two of them, ceritinib and alectinib, were approved in 2014 for advanced ALK-positive NSCLC in the US and Japan, respectively. Heat shock protein 90 (Hsp90) inhibitors also showed activity against ALK-positive NSCLC. Here we review the recent development of crizotinib, ceritinib, alectinib and other second-generation ALK inhibitors as well as Hsp90 inhibitors. We also discuss management strategies for advanced ALK-positive NSCLC.

Keywords: alectinib, ceritinib, crizotinib, EML4–ALK rearrangement, heat shock protein 90 inhibitor, non-small cell lung cancer

Introduction

In the era of molecular targeted therapy, epidermal growth factor receptor tyrosine kinase inhibitors (EGFR-TKIs) such as gefitinib, erlotinib and afatinib have been demonstrated to prolong progression-free survival (PFS) and preserve the quality of life when used as the first-line treatment for patients with EGFR-mutated advanced non-small cell lung cancer (NSCLC) [Mok et al. 2009; Maemondo et al. 2010; Sequist et al. 2013; Yang et al. 2013].

In 2007, Soda and colleagues first described the rearrangements in the anaplastic lymphoma kinase (ALK) gene and the echinoderm microtubule-associated protein-like 4 (EML4) gene in a small proportion of patients with NSCLC [Soda et al. 2007, 2008]. Around 2–5% of NSCLC patients harboured this gene rearrangement [Soda et al. 2007; Kwak et al. 2010]. EML4–ALK rearrangement was identified as an oncogene and EML4–ALK fusion protein was found to possess transforming activity and oncogenic potential [Soda et al. 2007]. ALK inhibitors were effective in vitro for cell lines and in vivo for mouse models of tumours harbouring the EML4–ALK rearrangement [Koivunen et al. 2008; Soda et al. 2008]. The clinicopathological features of these patients included younger age, never/light smokers and adenocarcinoma histology (predominantly signet-ring cell subtype) [Inamura et al. 2008, 2009; Rodig et al. 2009; Shaw et al. 2009]. EML4–ALK rearrangements are typically mutually exclusive with EGFR mutations or K-RAS mutations [Wong et al. 2009; Gainor et al. 2013b].

In a retrospective study, malignant pleural effusions of patients with EGFR wildtype lung adenocarcinoma were tested for ALK rearrangement. All patients were not treated with any ALK inhibitor. The survival of patients with EML4–ALK rearrangement was better than that of patients without EML4–ALK rearrangement, although this study enrolled only patients with malignant pleural effusion, which may potentially lead to biases [Wu et al. 2012]. There were several genes reported to be fused to ALK, but the incidence was quite low [Takeuchi et al. 2009; Togashi et al. 2012].

Crizotinib, a multitargeted tyrosine kinase (including ALK, MET and ROS1) inhibitor, is effective in treating patients with ALK-positive (referred to as EML4–ALK rearrangement) advanced NSCLC [Cui et al. 2011]. In 2011, this drug received approval from the US Food and Drug Administration (FDA) in treating advanced ALK-positive NSCLC using the Vysis ALK Break-Apart FISH Probe Kit (Abbott Molecular, Inc.) based on the results of early phase clinical trials. Phase III clinical trials had been conducted to compare crizotinib monotherapy with standard second-line and first-line cytotoxic chemotherapy for advanced ALK-positive NSCLC patients [Shaw et al. 2013; Solomon et al. 2014]. Because of the rarity of ALK-positive NSCLC, immunohistochemical staining is a reliable screening tool to identify ALK-positive NSCLC and to reduce the cost of general screening using standard methods [Jokoji et al. 2010; Yi et al. 2011; Park et al. 2012; Conklin et al. 2013]. Crizotinib is effective in treating some subsets of ALK-negative NSCLC, such as tumours harbouring ROS1 rearrangement and tumours with de novo MET amplification [Ou et al. 2011b; Bergethon et al. 2012; Yasuda et al. 2012]. Subjects without ALK-positive NSCLC are not covered in detail in this review.

Ceritinib, a second-generation ALK inhibitor, which is more potent against ALK than crizotinib but not active against MET, was approved by the FDA in 2014 for advanced ALK-positive NSCLC pretreated with crizotinib [Shaw et al. 2014]. Alectinib, another second-generation ALK inhibitor, was approved in Japan in 2014 for advanced ALK-positive NSCLC. Heat shock protein 90 (Hsp90) inhibitors were also effective in treating ALK-positive NSCLC [Normant et al. 2011]. However, drugs in this class have not yet been approved by any regulatory agencies. Here, we review the recent development of crizotinib, ceritinib, alectinib and other second-generation ALK inhibitors, as well as Hsp90 inhibitors. We also discuss management strategies for advanced ALK-positive NSCLC.

A literature review of clinical studies published between January 2007 and July 2014 was conducted using PubMed and MEDLINE, with keyword entry of ‘non-small cell lung cancer’, ‘EML4–ALK rearrangement’, ‘crizotinib’, ‘ceritinib’, ‘alectinib’ and ‘heat shock protein 90 inhibitor’. A manual search of abstracts presented at major oncology meetings was also performed.

First-generation ALK inhibitor: crizotinib

Overview of clinical development of crizotinib

Crizotinib was approved under the FDA’s accelerated approval programme in 2011 based on the results of two single-arm clinical trials mentioned below [Kwak et al. 2010; Kim et al. 2012]. In a phase I study (PROFILE 1001), 149 patients with advanced ALK-positive NSCLC underwent treatment with crizotinib at a dose of 250 mg twice daily. Most patients (71%) were never-smokers and 97% of the patients had adenocarcinoma histology. The overall response rate (ORR) was 60.8% and the median PFS was 9.7 months. The estimated overall survival (OS) rates at 6 months and 12 months were 87.9% and 74.8%, respectively [Camidge et al. 2012]. Visual effects, nausea and diarrhoea were the most commonly reported adverse effects (AEs). The global single-arm phase II study (PROFILE 1005) of crizotinib in treating advanced ALK-positive NSCLC after the progression of at least one line of cytotoxic chemotherapy revealed an ORR of 59.8% and a median PFS of 8.1 months [Kim et al. 2012].

A phase III study (PROFILE 1007) compared crizotinib with pemetrexed or docetaxel chemotherapy after failing one prior platinum-based chemotherapy. A total of 347 patients with advanced ALK-positive NSCLC were randomised, and it was observed that the median PFS (primary endpoint) was significantly longer in the crizotinib group (7.7 months) compared with the chemotherapy group (3.0 months) [hazard ratio (HR): 0.49, 95% confidence interval (CI) 0.37–0.64; p < 0.001). ORRs were 65% in the crizotinib group and 20% in the chemotherapy group (p < 0.001). Patients in the crizotinib group reported greater reduction of lung cancer related symptoms and improvement in the overall quality of life compared with the chemotherapy group [Shaw et al. 2013]. The common AEs are listed in Table 1. The most common grade 3 or 4 AE was elevated aminotransferase levels, which developed in 16% of patients who underwent crizotinib treatment. Grade 3 or 4 neutropenia developed in 13% of patients in the crizotinib group. Interestingly, ORRs and PFS were different between the pemetrexed (29% and 4.2 months, respectively) and docetaxel groups (7% and 2.6 months, respectively). ORR was higher than that of the general population who underwent second-line chemotherapy with pemetrexed [Hanna et al. 2004]. Other retrospective studies also demonstrated that ALK positivity was a predictive factor of pemetrexed efficacy [Camidge et al. 2011; Lee et al. 2011].

Table 1.

Selected clinical studies of ALK inhibitors for advanced ALK-positive NSCLC.

Drug Study name or ClinicalTrials.gov identifier Phase Population Comparator n ORR PFS Adverse effects or remarks
Crizotinib PROFILE 1007 Shaw et al. [2013] III Platinum-based chemotherapy pretreated Pemetrexed or docetaxel 347 65% versus. 20% (p < .001) 7.7 versus 3.0 months (HR: 0.49, 95% CI 0.37–0.64; p < .001) Visual disorder (60%), diarrhoea (60%), nausea (55%), vomiting (47%), constipation (42%), elevated aminotransferase levels (38%), oedema (31%), fatigue (27%)
PROFILE 1014 Solomon et al. [2014] III Treatment naïve, nonsquamous NSCLC Platinum plus pemetrexed 343 74% versus. 45% (p < .001) 10.9 versus 7.0 months (HR 0.45, 95% CI 0.35–0.60; p < .001) Visual disorder (71%), diarrhoea (61%), oedema (49%), vomiting (46%), constipation (43%), elevated aminotransferase levels (36%), upper respiratory infection (32%), abdominal pain (26%)
Ceritinib ASCEND-1 Shaw et al. [2014] I Advanced malignancy* No 114 58% (95% CI 0.48–0.67) 7.0 months (95% CI 5.6–9.5) Nausea (82%), diarrhoea (75%), vomiting (65%), fatigue (47%), and increased ALT level (35%)
NCT01685060 II Chemotherapy and crizotinib pretreated No NA NA NA Ongoing clinical trial
NCT01685138 II Crizotinib-naïve No NA NA NA Ongoing clinical trial
ASCEND-4 III Treatment naïve, nonsquamous NSCLC Platinum plus pemetrexed NA NA NA Ongoing clinical trial
ASCEND-5 III Platinum-based chemotherapy and crizotinib pretreated Pemetrexed or docetaxel NA NA NA Ongoing clinical trial
Alectinib$ AF-001JP Seto et al. [2013] I/II ALK-inhibitor-naïve No 46 93.5%, (95% CI 82.1–98.6) NA Dysgeusia (30%), increased AST level (28%), increased blood bilirubin level (28%), increased blood creatinine level (26%), rash (26%), constipation (24%), stomatitis (15%), and myalgia (13%).
AF-002JGGadgeel et al. [2014] I/II Crizotinib pretreated No 47 55% NA Fatigue (30%), myalgia (17%), and peripheral oedema (15%)
NCT01871805 II Chemotherapy and crizotinib pretreated No NA NA NA Ongoing clinical trial
ALEX III Treatment naïve Crizotinib NA NA NA Ongoing clinical trial
AP26113 NCT01449461 Gettinger et al. [2014b] I/II Crizotinib pretreated or ALK inhibitor naïve No 57 Crizotinib pretreated: 69% (95% CI 54–81%) Crizotinib pretreated: 10.9 months Nausea (34%), diarrhoea (26%), fatigue (22%), and vomiting (16%)
ALTA II Crizotinib pretreated No NA NA NA Ongoing clinical trial
*

Advanced malignancies were not restricted to NSCLC.

$

This drug was approved only in Japan as of September 2014.

This study was conducted in Japan with Japan Pharmaceutical Information Center, number JapicCTI-101264.

ALK, anaplastic lymphoma kinase; ALT, alanine aminotransferase; AST, aspartate aminotransferase; CI, confidence interval; HR, hazard ratio; NA, not available; ORR, overall response rate; PFS, progression-free survival.

Another phase III study (PROFILE 1014) compared crizotinib monotherapy with pemetrexed plus cisplatin chemotherapy as the first-line of treatment for advanced ALK-positive nonsquamous NSCLC. A total of 343 patients were randomised and the PFS (primary endpoint) was observed to be significantly longer in the crizotinib group (median 10.9 versus 7.0 months; HR 0.45, 95% CI 0.35–0.60; p < .001). ORR was 74% in the crizotinib group and 45% in the chemotherapy group [Solomon et al. 2014].

The OS results were immature in the aforementioned two phase III studies. The common AEs of crizotinib therapy are listed in Table 1. Other less common AEs had been reported, such as QTc prolongation, bradycardia, hypogonadism, renal cysts and interstitial pneumonitis [Ou et al. 2011, 2013; Weickhardt et al. 2012, 2013; Tamiya et al. 2013; Lin et al. 2014]. The management of these AEs has been reviewed elsewhere [Rothenstein and Letarte, 2014].

Activity against central nervous system metastases

The activity of crizotinib in central nervous system (CNS) metastases is still debatable [Gainor et al. 2013a; Maillet et al. 2013; Costa et al. 2015]. Many treatment strategies, such as high-dose crizotinib and high-dose crizotinib in combination with high-dose pemetrexed, had been reported to treat CNS disease [Gandhi et al. 2013; Kim et al. 2013c]. Some experts suggested that isolated CNS disease progression after crizotinib therapy should be treated with radiotherapy while continuing crizotinib therapy [Takeda et al. 2013]. A retrospective analysis of patients who developed RECIST-defined progressive disease (PD) in PROFILE 1001 and PROFILE 1005 studies revealed that 62% (120/194) of the patients continued crizotinib therapy after PD. Most of them had good performance status (ECOG PS 0-1) and 51% of them had brain metastases as the sole site of PD. The authors concluded that this treatment strategy may provide survival benefit [Ou et al. 2014a].

Mechanisms of crizotinib resistance

Despite the initial treatment response of crizotinib, PD inevitably develops after a period of treatment. The mechanisms of crizotinib resistance had been studied, and one of these mechanisms is a secondary mutation in the kinase domain of ALK, for example, the ‘gatekeeper mutation’ of substitution of leucine with methionine at position 1196 (L1196M) [Choi et al. 2010]. Other resistance mutations had been reported (Table 2) [Choi et al. 2010; Sasaki et al. 2010, 2011; Zhang et al. 2011; Doebele et al. 2012; Katayama et al. 2012; Lovly and Pau, 2012; Huang et al. 2013; Kim et al. 2013b; Ou et al. 2014b]. In addition to the secondary mutation, activation of alternative pathway (EGFR and KIT), ALK amplification, epithelial–mesenchymal transition (EMT) and insulin-like growth factor 1 receptor (IGF-1R) pathway activation also resulted in crizotinib resistance [Katayama et al. 2012; Tanizaki et al. 2012; Yamada et al. 2012; Kim et al. 2013; Kobayashi et al. 2013a; Lovly et al. 2014; Yamaguchi et al. 2014]. In some patients, the mechanism of acquired resistance remains unknown [Costa and Kobayashi,. 2012].

Table 2.

Characteristics of second-generation ALK inhibitors.

Drugs Molecular targets other than ALK Activity against L1196M Activity against C1156Y Activity against G1202R Activity against other crizotinib-resistant mutations No activity Mechanism of acquired resistance
Ceritinib IGF1-R, InsR, STK22D Yes No No G1269A, S1206Y, I1171T, V1180L 1151T-ins, L1152P, F1174C F1174C, G1202R
Alectinib GAK, LTK Yes Yes No G1269A, S1206Y, L1152R, F1174L, 1151T-ins F1174V I1171T, V1180L, G1202R
AP26113 ROS1, EGFRdel19/T790M Yes Yes Yes G1269A, S1206Y, D1203N, F1174C, 1151T-ins, G1269S, I1171T, E1210K, F1245C, S1206R* NA NA
ASP3026 ROS1, ACK Yes NA NA NA NA NA
PF-06463922 ROS1 Yes NA Yes G1269A NA NA
TSR-011 TRK-A, TRK-B, TRK-C Yes NA NA NA NA NA
RXDX-101 ROS1, TRK-A, TRK-B, TRK-C Yes Yes NA NA NA NA
X-396 MET Yes Yes NA NA NA NA
CEP-37440 FAK NA NA NA NA NA NA
*

S1206R confers to greatest resistance to AP26113 among the tested crizotinib-resistant mutations.

ALK, anaplastic lymphoma kinase; NA, not available.

Novel approaches to overcome crizotinib-acquired resistance had been reported. The most common strategy is to use second-generation ALK inhibitors to overcome resistance mediated by secondary ALK mutations (discussed in detail below). Other strategies, such as combination therapy with Hsp90 inhibitors, EGFR inhibitors, KIT inhibitors (e.g. imatinib) or IGF-1R inhibitors, had been reported [Sasaki et al. 2010, 2011; Katayama et al. 2012; Tanizaki et al. 2012; Yamada et al. 2012; Kim et al. 2013; Lovly et al. 2014; Yamaguchi et al. 2014]. Other strategies of crizotinib therapy are under development. A phase Ib trial of crizotinib in combination with immunotherapeutic agent ipilimumab (an anti-CTLA-4 monoclonal antibody) is ongoing [ClinicalTrials.gov identifier: NCT01998126]. Continuous treatment with crizotinib beyond disease progression is another strategy to extend the benefit of crizotinib therapy [Ou et al. 2014a]. A randomised phase II trial of crizotinib plus pemetrexed versus pemetrexed alone in patients with ALK-positive nonsquamous NSCLC who have progressed after previous benefit from crizotinib therapy is ongoing [ClinicalTrials.gov identifier: NCT02134912].

Second-generation ALK inhibitors

Second-generation ALK inhibitors were designed to have more potent activity against ALK, to overcome crizotinib-resistant mutations and to have better activity in CNS disease. Many novel agents are under development as mentioned below. The characteristics of second-generation ALK inhibitors are listed in Table 2. Those drugs that were in phase II/III development are listed in Table 1 and those in phase I/II development are listed in Table 3.

Table 3.

Novel ALK inhibitors in early phase clinical study.

Drugs ClinicalTrials.gov identifier Phase n ORR Adverse effects
ASP3026 NCT01401504 I 16 50% (crizotinib pretreated ALK-positive NSCLC) Fatigue, vomiting, nausea and constipation
PF-06463922 NCT01970865 I/II NA NA NA
TSR-011 NCT02048488 I/II 5 60% (crizotinib pretreated ALK-positive NSCLC) NA
RXDX-101 NCT02097810 I/II NA NA Asthenia, paresthesia, nausea, and diarrhea
X-396 NCT01625234 I/II 11 55% (ALK-positive NSCLC) Nausea, rash, vomiting, fatigue, oedema, and pruritus
CEP-37440 NCT01922752 I NA NA NA

ALK, anaplastic lymphoma kinase; NA, not available; ORR, overall response rate.

Ceritinib

Ceritinib (LDK378) is an orally administered potent ALK inhibitor derived from the compound NVP-TAE684 [Galkin et al. 2007; Marsilje et al. 2013]. In preclinical studies, this drug demonstrated greater antitumour potency than crizotinib, and displayed activity against some crizotinib-resistant mutations (Table 2) [Friboulet et al. 2014]. In a phase I study (ASCEND-1) [ClinicalTrials.gov identifier: NCT01283516] (Table 1), 750 mg once daily was determined as the maximum tolerated dose (MTD) with the dose-limiting toxicities (DLT) of diarrhoea, vomiting, dehydration, elevated aminotransferase levels and hypophosphataemia. Among the 114 patients who received ceritinib at least 400 mg/day, ORR was 58% and the PFS was 7.0 months. Among the 80 patients who had been previously treated with crizotinib, ORR was 56% and 19 patients with crizotinib-resistant disease underwent a tumour biopsy before ceritinib therapy. ALK mutation and ALK amplification were detected in some of the responders, but other responders had neither ALK mutation nor ALK amplification. Among the patients who were crizotinib-naïve and treated with ceritinib at least 400 mg/day, ORR was 62%. The common AEs are listed in Table 1. The most common grade 3 or 4 AEs were increased ALT level (21%), increased aspartate aminotransferase (AST) level (11%) and diarrhoea (7%), All of these AEs were reversible after discontinuation of ceritinib therapy [Shaw et al. 2014].

Ceritinib received accelerated approval from the FDA in April 2014 for patients with metastatic ALK-positive NSCLC who were previously treated with crizotinib. An updated report of ASCEND-1 study disclosed the efficacy data of the expansion cohort, in which all of the patients were treated with a starting dose of 750 mg/day. A total of 246 patients with ALK-positive NSCLC were enrolled, including 163 ALK inhibitor-pretreated (crizotinib or alectinib) and 83 ALK inhibitor-naïve patients. ORRs were 58.5%, 54.6% and 66.3% in the overall population, ALK inhibitor-pretreated group and ALK inhibitor-naïve group, respectively. The median PFS periods were 8.2 and 6.9 months in the overall population and ALK inhibitor-pretreated group, respectively. The median PFS in the ALK inhibitor-naïve group was not reached and the PFS rate at 12 months was 61.3%. In addition to the AEs mentioned above, around 4% of the patients developed interstitial lung disease/pneumonitis, and 9.4% (24/255, including 9 non-NSCLC patients) of the patients discontinued the study drug because of the AEs. However, regarding its activity in CNS disease, 124 patients had brain metastases at baseline, and 10 and 4 patients had measurable lesions in the ALK inhibitor-pretreated and ALK inhibitor-naïve groups, respectively. The intracranial ORRs were 40% and 75%, respectively. The authors concluded that ceritinib therapy had a high rate of durable responses and prolonged PFS in both ALK inhibitor-pretreated group and ALK inhibitor-naïve patients [Kim et al. 2014]. Ceritinib treatment showed activity in patients with brain metastases. Ongoing phase II and phase III clinical trials of ceritinib therapy for patients with advanced ALK-positive NSCLC are listed in Table 1.

Alectinib

Alectinib (RO5424802/CH5424802) is a highly selective and orally administered ALK inhibitor. Preclinical data revealed that alectinib had activity against some crizotinib-resistant mutations (Table 2) [Sakamoto et al. 2011; Kinoshita et al. 2012; Kodama et al. 2014a]. Alectinib showed potent efficacy against intracranial tumour in mouse models [Kodama et al. 2014b]. It had no activity against MET, ROS1 and some crizotinib-resistant mutations (Table 2) [Ou et al. 2014b].

In a phase I/II study (AF-001JP study) conducted in Japan (Table 1), patients with ALK-positive and ALK inhibitor-naïve NSCLC were treated with alectinib. In the phase I portion, there were no DLT or AEs of grade 4 up to the highest dose, and 300 mg twice daily was the recommended phase II dose. In the phase II portion, 43 out of 46 (93.5%) treated patients had objective responses. The common AEs are listed in Table 1. Grade 3 AEs were recorded in 26% of the patients, and the most common grade 3 AE was decreased neutrophil count and increased blood creatine phosphokinase (4%). No grade 4 AEs were recorded [Seto et al. 2013; Yang, 2013]. Based on this study, alectinib was approved in Japan in July 2014 for advanced ALK-positive NSCLC.

In another study conducted in Japan, 35 patients were enrolled and 29 patients (83%) had been pretreated with one or more than one ALK inhibitors. Among 28 patients pretreated with crizotinib, the ORR was 58.3% (95% CI 36.6–77.9). Two patients with measurable brain tumours were observed to have complete response at the first assessment [Nakagawa et al. 2014].

In the US, a phase I/II study (AF-002JG study) [ClinicalTrials.gov identifier: NCT01588028] of 600 mg twice a day was selected as the recommended phase II dose in patients with crizotinib-resistant ALK-positive NSCLC. Among the 47 enrolled patients, 44 patients could be assessed for drug activity and the ORR was 55%. Of 32 patients with CNS metastases, ORR was 52% [Gadgeel et al. 2014].

V1180L (gatekeeper) and I1171T mutations conferred resistance to alectinib, and ceritinib might overcome these alectinib-resistant mutations [Katayama et al. 2014]. Ongoing studies of alectinib therapy for patients with advanced ALK-positive NSCLC are listed in Table 1. One of these ongoing studies (ALEX) [ClinicalTrials.gov identifier: NCT02075840], directly compares alectinib with crizotinib in treatment-naïve advanced ALK-positive NSCLC patients.

AP26113

AP26113 is a potent and orally active inhibitor of ALK. In preclinical data, this drug demonstrated activity against ROS1 and most crizotinib-resistant mutations, including G1202R [Zhang et al. 2010; Squillace et al. 2013]. This drug has activity against activating EGFR mutations and EGFRT790M resistance mutation while sparing native EGFR (Table 2) [Rivera et al. 2012; Camidge et al. 2013]. In a phase I/II study [ClinicalTrials.gov identifier: NCT01449461], early onset pulmonary symptoms developed in 14% of patients with a starting dose of 180 mg/day. In 28 patients with a starting dose of 90 mg/day for 7 days followed by an escalation to 180 mg/day, none of them developed pulmonary symptoms during the first 7 days at 90 mg or the first 7 days at 180 mg. This dosing strategy was chosen as the phase II dosage. In the phase II part, 57 patients were enrolled and ORR was 69% among the 51 crizotinib-pretreated patients and 100% among the 6 ALK inhibitor-naïve patients. The median PFS of the 49 crizotinib-pretreated patients with follow-up scans was 10.9 months. A total of 9 out of 13 patients (69%) with brain metastases responded to AP26113 therapy. The common AEs are listed in Table 1 [Gettinger et al. 2014a, 2014b]. A phase II trial of AP26113 (ALTA) [ClinicalTrials.gov identifier: NCT02094573] is ongoing for advanced crizotinib-pretreated ALK-positive NSCLC.

ASP3026

ASP3026 is a potent ALK and ROS1 inhibitor. In mouse models, this drug showed potent antitumour activity against ALK-positive NSCLC cells expressing crizotinib-resistant L1196M mutation [Mori et al. 2014]. In a phase I study [ClinicalTrials.gov identifier: NCT01401504], a dose level of 525 mg/day was chosen as a recommended phase II dose. The common AEs are listed in Table 3. ORR was 50% among 16 patients and the median PFS was 5.5 months (95% CI 3.7–11) [Maitland et al. 2014].

PF-06463922

PF-06463922 is a potent, macrocyclic inhibitor of ALK and ROS1. This drug has low propensity for P-glycoprotein 1 mediated efflux and good passive permeability in order to facilitate its CNS penetration [Johnson et al. 2014]. In preclinical studies, this drug demonstrated activity in mice with tumour xenograft that expressed ALK and some crizotinib-resistant mutations, including G1202R (Table 2). This drug achieved brain exposure of 20–30% of its plasma level in mice and regressed the brain tumours [Johnson et al. 2013; Zou et al. 2013a, 2013b]. A phase I/II study of PF-06463922 in patients with advanced ALK- or ROS1-positive NSCLC is ongoing [ClinicalTrials.gov identifier: NCT01970865].

TSR-011

TSR-011 is a potent ALK and tropomyosin-related kinase (TRK) A, B and C (encoded by NTRK1, NTRK2 and NTRK3, respectively) inhibitor. In preclinical studies, this drug showed activity against crizotinib-resistant L1196M mutation (Table 2) [Wilcoxen et al. 2012]. In a phase I/IIa study [ClinicalTrials.gov identifier: NCT02048488], 23 patients were enrolled (including five ALK-positive NSCLC). The DLT included dysaesthesia and QTc prolongation [Weiss et al. 2014]. Rearrangement in NTRK1 had been reported in a small portion of NSCLC patients without known oncogenic alterations. Treatment with inhibitors of TRKA kinase inhibited cell growth [Vaishnavi et al. 2013]. TSR-011 might have activity in this patient group.

RXDX-101

RXDX-101 (former name: NMS-E628) is an inhibitor of ALK, ROS1 and TRK A, B and C. In preclinical studies, this drug demonstrated in vitro and in vivo activity against ALK-positive NSCLC and some crizotinib-resistant mutations (Table 2) [Ardini et al. 2009, 2011]. This drug passed through the blood–brain barrier in animal models and controlled intracranial tumours [Ardini et al. 2011]. In a phase I study, RXDX-101 was administered to patients with advanced solid tumours with relevant molecular alterations. One patient with ALK-positive NSCLC achieved PR and another patient with ALK-positive NSCLC had prolonged stable decease (SD). No DLT had been observed in all the tested dose levels. The common AEs (mainly grade 1–2) are listed in Table 3 [De Braud et al. 2014a, 2014b]. A global phase I/II study was initiated in patients with cancer confirmed to be positive for ALK, ROS1 and TRK A, B and C molecular alterations, especially NSCLC, colorectal cancer, prostate cancer, papillary thyroid cancer, pancreatic cancer and neuroblastoma [ClinicalTrials.gov identifier: NCT02097810].

X-396

X-396 is an ALK and MET inhibitor. It was more potent inhibitor of ALK and less potent inhibitor of MET compared with crizotinib. This drug also had activity against some crizotinib-resistant mutations (Table 2). In combination with the mTOR inhibitor, rapamycin, X-396 displayed synergistic growth inhibition [Lovly et al. 2011]. The efficacy results and the common AEs of a phase I/II study [ClinicalTrials.gov identifier: NCT01625234] are listed in Table 3. Doses up to 225 mg were well tolerated. The enrolment is ongoing in the expansion cohort [Horn et al. 2014].

CEP-37440

CEP-37440, an analogue of CEP-28122, is an ALK and FAK inhibitor [Shanthi et al. 2014]. Preclinical studies demonstrated activity of CEP-28122 against ALK-positive human cancer cells and tumour xenograft mouse models [Cheng et al. 2012]. The development of CEP-28122 was terminated because severe lung toxicity was developed in animal studies [Wang et al. 2014]. A phase I trial of CEP-37440 [ClinicalTrials.gov identifier: NCT01922752] for patients with advanced and metastatic solid tumours is ongoing.

Summary of ALK inhibitors

Crizotinib provides benefits to patients with advanced ALK-positive NSCLC in the first-line, second-line or heavily pretreated settings, with the PFS being around 7–10 months. Resistance disease inevitably developed after crizotinib therapy. Crizotinib-resistant ALK mutations (e.g. L1196M and G1269A) were one of the resistance mechanisms. Brain metastasis was another cause of PD. Novel ALK inhibitors were active against various crizotinib-resistant ALK mutations and brain metastases. Ceritinib is approved by the FDA for crizotinib-pretreated ALK-positive NSCLC. In Japan, alectinib is also available for ALK-positive NSCLC in the setting of either crizotinib-naïve or crizotinib-pretreated disease. These two novel ALK inhibitors gained accelerated approvals based on phase I/II studies, and confirmatory phase III studies are needed to determine the efficacy and AEs in different clinical setting. However, these aforementioned clinical trials enrolled crizotinib-pretreated patients, but the mechanisms of crizotinib resistance in these patients were not all the same. Some of these patients underwent cytotoxic chemotherapy before entering the clinical trials and clonal repopulation of crizotinib-sensitive cells might develop during chemotherapy. This situation cannot be classified as a mechanism of crizotinib resistance. Tumours that develop resistance under an ALK inhibitor therapy warrant repeat biopsies to identify the mechanism of drug resistance, although this procedure is not current standard of care and should be performed under clinical trial settings.

Even the resistance mechanisms of ceritinib (F1174C and G1202R) and alectinib (I1171T, V1180L and G1202R) had been identified (Table 2) [Friboulet et al. 2014; Ou et al. 2014b; Katayama et al. 2014]. Ceritinib was active against alectinib-resistant I1171T and V1180L mutations, but both ceritinib and alectinib were ineffective against G1202R mutation. Other novel ALK inhibitors, such as PF-06463922 and AP26113, have activity against G1202R and could be effective in treating patients who developed this mutation [Politi and Gettinger, 2014]. We encourage patients with ALK-positive NSCLC to participate in clinical trials to address the best sequence, combination, intercalation or novel therapeutic strategy for ALK inhibitors and to extend survival for patients [Gainor and Shaw 2013].

Hsp90 inhibitors in ALK-positive NSCLC

Hsp90 is a molecular chaperone that guides the normal folding, proteolytic turnover, intracellular disposition of regulators of cell growth and survival [Whitesell and Lindquist, 2005]. Hsp90 is also a chaperone involved in the stabilisation of many oncoproteins and has been implicated in tumourigenesis [Trepel et al. 2010]. In ALK-positive NSCLC, the tumour is addicted to the fusion protein resulting from chromosomal rearrangements and this protein is thought to be a client of Hsp90 [Neckers and Workman, 2012]. Hsp90 inhibitors are effective in treating these patients in early clinical studies and four Hsp90 inhibitors (IPI-504, ganetespib, AUY922 and AT13387) are currently under clinical investigation in ALK-positive NSCLC patients [Pillai and Ramalingam, 2014].

IPI-504

Retaspimycin hydrochloride (IPI-504) is an Hsp90 inhibitor. In preclinical studies, the degradation of EML4–ALK fusion protein was induced by IPI-504 therapy and it resulted in the inhibition of downstream signalling pathways, induction of growth arrest and apoptosis [Normant et al. 2011]. In a phase II trial of IPI-504 monotherapy for patients with molecularly defined NSCLC, two out of three patients with ALK-positive NSCLC responded to IPI-504 therapy and the remaining one patient had prolonged SD. The most common AEs were fatigue, nausea and diarrhoea (grades 1 and 2). Grade 3 or higher liver function abnormality was observed in 11.8% of the patients [Sequist et al. 2010]. A phase II study of IPI-504 for ALK-positive NSCLC [ClinicalTrials.gov identifier: NCT01228435] was terminated early because of slow patient recruitment and competing studies.

Ganetespib

Ganetespib (STA-9090) is a nongeldanamycin triazolone-containing Hsp90 inhibitor. In preclinical studies, ganetespib demonstrated activity against KRAS mutant, EGFR mutant (including EGFRdel 19 and EGFRL858R/T790M), ERBB2 mutant and c-MET amplification in NSCLC in animal models [Acquaviva et al. 2012; Shimamura et al. 2012; Ying et al. 2012]. In ALK-positive NSCLC cell lines, ganetespib induced loss of EML4–ALK expression and depletion of oncogenic signalling proteins [Sang et al. 2013]. Ganetespib overcame multiple forms of crizotinib resistance, including some secondary ALK mutations, and ganetespib in combination with novel ALK inhibitors other than crizotinib also led to increased activity [Sang et al. 2013]. In addition to ALK-positive NSCLC cell lines, cancer cells driven by oncogenic rearrangement of ROS1 and RET were sensitive to ganetespib [Sang et al. 2013].

In a phase II study of ganetespib monotherapy for genotypically defined advanced NSCLC, 99 patients were enrolled and ORR was reported to be 4%. All of these responders (n=4) were patients with crizotinib-naïve ALK-positive NSCLC. The treatment duration of the responders ranged from 7.4 to 21 months. Three out of the remaining four patients with ALK-positive NSCLC achieved SD as the best response. The median PFS of these 8 patients was 8.1 months, which was longer than the patients without ALK rearrangement (HR, 0.223; 95% CI 0.085–0.582) [Socinski et al. 2013]. The most common AEs were diarrhoea (81.8%), fatigue (57.6%), nausea (41.4%), decreased appetite (37.4%) and constipation (26.3%). Two patients experienced treatment-related death (one cardiac arrest and one renal failure) [Socinski et al. 2013].

A phase II study of ganetespib for ALK-positive NSCLC [ClinicalTrials.gov identifier: NCT01562015] and a phase I study of ganetespib plus crizotinib for crizotinib-naïve advanced ALK-positive NSCLC [ClinicalTrials.gov identifier: NCT01579994] are ongoing. Ganetespib is under clinical development in combination with docetaxel in patients with advanced NSCLC, which is not restricted to ALK-positive NSCLC (GALAXY-2) [ClinicalTrials.gov identifier: NCT01798485] [Proia et al. 2012; Ramalingam et al. 2013, 2014].

AUY922

AUY922 is a potent isoxazole-based nongeldanamycin Hsp90 inhibitor, which acts via cytostasis, apoptosis, invasion and angiogenesis to inhibit tumour growth and metastasis [Eccles et al. 2008]. A phase II study [ClinicalTrials.gov identifier: NCT01124864] of AUY922 for advanced NSCLC enrolled 121 patients. A total of 6 out of 21 patients (29%) with ALK-positive NSCLC achieved PR, and 4 of the 6 responders were crizotinib-naïve. The estimated PFS rate at 18 weeks was 42%. The most common AEs were eye disorder (77%), diarrhoea (74%) and nausea (46%) [Felip et al. 2012]. A phase II study of AUY922 for advanced NSCLC pretreated with crizotinib [ClinicalTrials.gov identifier: NCT01752400] and a phase Ib study of combination therapy of ceritinib plus AUY922 for crizotinib-pretreated ALK-positive NSCLC [ClinicalTrials.gov identifier: NCT01772797] are ongoing. In addition to ALK-positive population, investigators also tested the efficacy of AUY922 in HER2-mutant/amplification, EGFR mutant (including EGFRT790M resistance mutation, exon 20 mutation and other uncommon mutations), BRAF mutant and ROS1 or RET rearrangement in NSCLC [ClinicalTrials.gov identifier: NCT01922583, NCT01854034, NCT01646125] [Garon et al. 2013; Nogova et al. 2014].

AT13387

AT13387 is a high-affinity Hsp90 inhibitor [Woodhead et al. 2010]. Preclinical studies demonstrated its activity against EGFR mutant and c-MET-amplified NSCLC [Graham et al. 2012]. A phase I/II study of AT13387 alone or in combination with crizotinib for ALK-positive and crizotinib-pretreated patients [ClinicalTrials.gov identifier: NCT01712217] is ongoing.

Summary of Hsp90 inhibitors

Hsp90 inhibitors had shown activity against ALK-positive NSCLC in early phase studies and even overcame crizotinib-resistant mutations [Katayama et al. 2012; Sang et al. 2013]. However, Hsp90 inhibitors had limited activity against CNS metastatic tumours and their clinical benefits were restricted to patients without CNS metastases. However, the AEs of Hsp90 inhibitor therapy were higher than with second-generation ALK inhibitors. While there are many second-generation ALK inhibitors available in clinical practice or clinical trial settings, the development of Hsp90 inhibitors should be influenced. Novel approaches such as combination therapy with crizotinib or second-generation ALK inhibitors in either crizotinib-naïve or crizotinib-pretreated patients are under investigation. We encourage patients to participate in clinical trials to address the best combination or treatment strategy of Hsp90 inhibitors.

Conclusion

In patients with advanced ALK-positive NSCLC, crizotinib therapy was deemed to be indispensable. After disease progression, second-generation ALK inhibitors, ceritinib and alectinib, provided opportunities to overcome acquired resistance and achieve tumour control. Second-generation ALK inhibitors are not widely available around the world and cytotoxic chemotherapy is still the standard of care. We hope clinical trials are able to develop the next generation of ALK inhibitors to overcome resistance, be effective in treating CNS metastases, and extend survival in patients. Hsp90 inhibitors are currently not available in daily practice. We still rely on clinical trials to identify the best way to incorporate these drugs into clinical practice.

Acknowledgments

The authors would like to thank Enago (www.enago.tw) for the English language editing.

Footnotes

Funding: No sources of funding were used to assist with the preparation of this review. This study was supported in part by grant from the Ministry of Science and Technology, Taiwan (MOST 103-2325-B-002-034).

Conflict of interest statement: J.C.-H.Y. is a consultant and received honoraria from AstraZeneca, Roche/Genentech, Boehringer Ingelheim, MSD, Merck Serono, Novartis, Pfizer, Clovis Oncology, Eli Lilly, Bayer, Celgene, Astellas, Innopharma, Ono Pharmaceutical, Chugai pharmaceutical. J.-Y.S. is a compensated advisor for Boehringer Ingelheim, Roche and AstraZeneca, and received honoraria as a speaker from AstraZeneca, Eli Lilly, Pfizer, Boehringer Ingelheim and Roche. C.-C.L. and B.-C.L. report no conflicts of interest in preparing this aticle.

Contributor Information

Bin-Chi Liao, Department of Oncology, National Taiwan University Hospital, Taipei, Taiwan.

Chia-Chi Lin, Department of Oncology, National Taiwan University Hospital, 7 Chung-Shan South Road, Taipei, Taiwan.

Jin-Yuan Shih, Department of Internal Medicine, National Taiwan University Hospital, Taiwan Department of Internal Medicine, College of Medicine, National Taiwan University, Taipei, Taiwan.

James Chih-Hsin Yang, Department of Oncology, National Taiwan University Hospital, Taiwan Graduate Institute of Oncology and Cancer Research Center, College of Medicine, National Taiwan University, Taipei, Taiwan.

References

  1. Acquaviva J., Smith D., Sang J., Friedland J., He S., Sequeira M., et al. (2012) Targeting KRAS-mutant non–small cell lung cancer with the Hsp90 inhibitor ganetespib. Mol Cancer Ther 11: 2633–2643. [DOI] [PubMed] [Google Scholar]
  2. Ardini E., Menichincheri M., Banfi P., Saccardo M., Rusconi L., Avanzi N., et al. (2011) Abstract A232: In vitro and in vivo activity of NMS-E628 against ALK mutations resistant to xalkori. Mol Cancer Ther 10: A232. [Google Scholar]
  3. Ardini E., Menichincheri M., De Ponti C., Amboldi N., Saccardo M., Texido G., et al. (2009) Characterization of NMS-E628, a small molecule inhibitor of anaplastic lymphoma kinase with antitumor efficacy in ALK-dependent lymphoma and non-small cell lung cancer models. Mol Cancer Ther 8: A243. [Google Scholar]
  4. Bergethon K., Shaw A., Ignatius Ou S., Katayama R., Lovly C., McDonald N., et al. (2012) ROS1 rearrangements define a unique molecular class of lung cancers. J Clin Oncol 30: 863–870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Camidge D., Bang Y., Kwak E., Iafrate A., Varella-Garcia M., Fox S., et al. (2012) Activity and safety of crizotinib in Patients with ALK-positive non-small-cell lung cancer: updated results from a phase 1 study. Lancet Oncol 13: 1011–1019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Camidge D., Bazhenova L., Salgia R., Weiss G., Langer C., Shaw A., et al. (2013) First-in-human dose-finding study of the ALK/EGFR inhibitor AP26113 in patients with advanced malignancies: updated results. ASCO Meeting Abstracts 31: abstract 8031. [Google Scholar]
  7. Camidge D., Kono S., Lu X., Okuyama S., Barón A., Oton A., et al. (2011) Anaplastic lymphoma kinase gene rearrangements in non-small cell lung cancer are associated with prolonged progression-free survival on pemetrexed. J Thorac Oncol 6: 774–780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Cheng M., Quail M., Gingrich D., Ott G., Lu L., Wan W., et al. (2012) CEP-28122, a highly potent and selective orally active inhibitor of anaplastic lymphoma kinase with antitumor activity in experimental models of human cancers. Mol Cancer Ther 11: 670–679. [DOI] [PubMed] [Google Scholar]
  9. Choi Y., Soda M., Yamashita Y., Ueno T., Takashima J., Nakajima T., et al. (2010) EML4-ALK mutations in lung cancer that confer resistance to ALK inhibitors. N Engl J Med 363: 1734–1739. [DOI] [PubMed] [Google Scholar]
  10. Conklin C., Craddock K., Have C., Laskin J., Couture C., Ionescu D. (2013) Immunohistochemistry is a reliable screening tool for identification of ALK rearrangement in non–small-cell lung carcinoma and is antibody dependent. J Thorac Oncol 8: 45–51. [DOI] [PubMed] [Google Scholar]
  11. Costa D., Kobayashi S. (2012) Acquired resistance to the ALK Inhibitor crizotinib in the absence of an ALK mutation. J Thorac Oncol 7: 623–625. [DOI] [PubMed] [Google Scholar]
  12. Costa D., Shaw A., Ou S., Solomon B., et al. (2015) Clinical experience with crizotinib in patients with advanced ALK-rearranged non–small-cell lung cancer and brain metastases. J Clin Oncol. 26 January 2015. [Epub ahead of print] [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Cui J., Tran-Dubé M., Shen H., Nambu M., Kung P., Pairish M., et al. (2011) Structure based drug design of crizotinib (PF-02341066), a potent and selective dual inhibitor of mesenchymal–epithelial transition factor (c-MET) kinase and anaplastic lymphoma kinase (ALK). J Med Chem 54: 6342–6363. [DOI] [PubMed] [Google Scholar]
  14. De Braud F., Pilla L., Niger M., Damian S., Bardazza B., Martinetti A., et al. (2014a) Phase 1 open label, dose escalation study of RXDX101, an oral Pan-TRK, ROS1, and ALK Inhibitor, in patients with advanced solid tumors with relevant molecular alterations. ASCO Meeting Abstracts 32: abstract 2502. [Google Scholar]
  15. De Braud F., Pilla L., Niger M., Damian S., Bardazza B., Martinetti A., et al. (2014b) 448PDRXDX-101, an oral pan-TRK, ROS1, and ALK inhibitor, in patients with advanced solid tumors with relevant molecular alterations. Ann Oncol 25: iv148–iv149. [Google Scholar]
  16. Doebele R., Pilling A., Aisner D., Kutateladze T., Le A., Weickhardt A., et al. (2012) Mechanisms of resistance to crizotinib in patients with ALK gene rearranged non–small cell lung cancer. Clin Cancer Res 18: 1472–1482. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Eccles S., Massey A., Raynaud F., Sharp S., Box G., Valenti M., et al. (2008) NVP-AUY922: a novel heat shock protein 90 inhibitor active against xenograft tumor growth, angiogenesis, and metastasis. Cancer Res 68: 2850–2860. [DOI] [PubMed] [Google Scholar]
  18. Felip E., Carcereny E., Barlesi F., Gandhi L., Sequist L., Kim S., et al. (2012) Phase II activity of the Hsp90 Inhibitor AUY922 in patients with ALK-rearranged (ALK+) or EGFR-mutated advanced non-small cell lung cancer (NSCLC). Ann Oncol 23: ix152–ix174. [Google Scholar]
  19. Friboulet L., Li N., Katayama R., Lee C., Gainor J., Crystal A., et al. (2014) The ALK inhibitor ceritinib overcomes crizotinib resistance in non–small cell lung cancer. Cancer Discov 4: 662–673. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Gadgeel S., Gandhi L., Riely G., Chiappori A., West H., Azada M., et al. (2014) Safety and activity of alectinib against systemic disease and brain metastases in patients with crizotinib-resistant ALK-rearranged non-small-cell lung cancer (AF-002JG): results from the dose-finding portion of a phase 1/2 study. Lancet Oncol 15: 1119–1128. [DOI] [PubMed] [Google Scholar]
  21. Gainor J., Ou S., Logan J., Borges L., Shaw A. (2013a) The central nervous system as a sanctuary site in ALK-positive non–small-cell lung cancer. J Thorac Oncol 8: 1570–1573. [DOI] [PubMed] [Google Scholar]
  22. Gainor J., Shaw A. (2013) Emerging paradigms in the development of resistance to tyrosine kinase inhibitors in lung cancer. J Clin Oncol 31: 3987–3996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Gainor J., Varghese A., Ou S., Kabraji S., Awad M., Katayama R., et al. (2013b) ALK rearrangements are mutually exclusive with mutations in EGFR or KRAS: an analysis of 1,683 patients with non–small cell lung cancer. Clin Cancer Res 19: 4273–4281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Galkin A., Melnick J., Kim S., Hood T., Li N., Li L., et al. (2007) Identification of NVP-TAE684, a potent, selective, and efficacious inhibitor of NPM-ALK. Proc Natl Acad Sci U S A 104: 270–275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Gandhi L., Drappatz J., Ramaiya N., Otterson G. (2013) High-dose pemetrexed in combination with high-dose crizotinib for the treatment of refractory CNS metastases in ALK-rearranged non–small-cell lung cancer. J Thorac Oncol 8: e3–e5. [DOI] [PubMed] [Google Scholar]
  26. Garon E., Finn R., Hamidi H., Dering J., Pitts S., Kamranpour N., et al. (2013) The Hsp90 inhibitor NVP-AUY922 potently inhibits non–small cell lung cancer growth. Mol Cancer Ther 12: 890–900. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Gettinger S., Bazhenova L., Salgia R., Langer C., Gold K., Rosell R., et al. (2014a) 1292PALK inhibitor AP26113 in patients with advanced malignancies, including ALK+ non-small cell lung cancer (NSCLC): updated efficacy and safety data. Ann Oncol 25: iv455. [Google Scholar]
  28. Gettinger S., Bazhenova L., Salgia R., Langer C., Gold K., Rosell R., et al. (2014b) Updated efficacy and safety of the ALK inhibitor AP26113 in patients (pts) with advanced malignancies, including ALK+ non-small cell lung cancer (NSCLC). ASCO Meeting Abstracts 32: abstract 8047. [Google Scholar]
  29. Graham B., Curry J., Smyth T., Fazal L., Feltell R., Harada I., et al. (2012) The heat shock protein 90 inhibitor, AT13387, displays a long duration of action in vitro and in vivo in non-small cell lung cancer. Cancer Sci 103: 522–527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Hanna N., Shepherd F., Fossella F., Pereira J., De Marinis F., Von Pawel J., et al. (2004) Randomised phase III trial of pemetrexed versus docetaxel in patients with non–small-cell lung cancer previously treated with chemotherapy. J Clin Oncol 22: 1589–1597. [DOI] [PubMed] [Google Scholar]
  31. Horn L., Infante J., Blumenschein G., Wakelee H., Arkenau H., Dukart G., et al. (2014) A Phase I trial of X-396, a novel ALK inhibitor, in patients with advanced solid tumors. ASCO Meeting Abstracts 32: abstract 8030. [Google Scholar]
  32. Huang D., Kim D., Kotsakis A., Deng S., Lira P., Ho S., et al. (2013) Multiplexed deep sequencing analysis of ALK kinase domain identifies resistance mutations in relapsed patients following crizotinib treatment. Genomics 102: 157–162. [DOI] [PubMed] [Google Scholar]
  33. Inamura K., Takeuchi K., Togashi Y., Hatano S., Ninomiya H., Motoi N., et al. (2009) EML4-ALK lung cancers are characterized by rare other mutations, a TTF-1 cell lineage, an acinar histology, and young onset. Mod Pathol 22: 508–515. [DOI] [PubMed] [Google Scholar]
  34. Inamura K., Takeuchi K., Togashi Y., Nomura K., Ninomiya H., Okui M., et al. (2008) EML4-ALK fusion is linked to histological characteristics in a subset of lung cancers. J Thorac Oncol 3: 13–17. [DOI] [PubMed] [Google Scholar]
  35. Johnson T., Bailey S., Burke B., Collins M., Cui J., Deal J., et al. (2013) Abstract PR10: Is CNS availability for oncology a no-brainer? Discovery of PF-06463922, a novel small molecule inhibitor of ALK/ROS1 with preclinical brain availability and broad spectrum potency against ALK-resistant mutations. Mol Cancer Ther 12: PR10. [Google Scholar]
  36. Johnson T., Richardson P., Bailey S., Brooun A., Burke B., Collins M., et al. (2014) Discovery of (10R)-7-amino-12-fluoro-2,10,16-trimethyl-15-oxo-10,15,16,17-tetrahydro-2H-8,4-(metheno)pyrazolo[4,3-h][2,5,11]-benzoxadiazacyclotetradecine-3-carbonitrile (PF-06463922), a macrocyclic inhibitor of anaplastic lymphoma kinase (ALK) and c-ros oncogene 1 (ROS1) with preclinical brain exposure and broad-spectrum potency against ALK-resistant mutations. J Med Chem 57: 4720–4744. [DOI] [PubMed] [Google Scholar]
  37. Jokoji R., Yamasaki T., Minami S., Komuta K., Sakamaki Y., Takeuchi K., et al. (2010) Combination of morphological feature analysis and immunohistochemistry is useful for screening of EML4-ALK-positive lung adenocarcinoma. J Clin Pathol 63: 1066–1070. [DOI] [PubMed] [Google Scholar]
  38. Katayama R., Friboulet L., Koike S., Lockerman E., Khan T., Gainor J., et al. (2014) Two novel ALK mutations mediate acquired resistance to the next-generation ALK inhibitor alectinib. Clin Cancer Res 20: 5686–5696. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Katayama R., Shaw A., Khan T., Mino-Kenudson M., Solomon B., Halmos B., et al. (2012) Mechanisms of acquired crizotinib resistance in ALK-rearranged lung cancers. Sci Transl Med 4: 120ra117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Kim D., Ahn M., Shi Y., De Pas T., Yang P., Riely G., et al. (2012) Results of a global phase II study with crizotinib in advanced ALK-positive non-small cell lung cancer (NSCLC). ASCO Meeting Abstracts 30: abstract 7533. [Google Scholar]
  41. Kim D., Mehra R., Tan D., Felip E., Chow L., Camidge D., et al. (2014) Ceritinib in advanced anaplastic lymphoma kinase (ALK)-rearranged (ALK+) non-small cell lung cancer (NSCLC): results of the ASCEND-1 trial. ASCO Meeting Abstracts 32: abstract 8003. [Google Scholar]
  42. Kim H., Kim W., Choi Y., Choi C., Rho J., Lee J. (2013a) Epithelial-mesenchymal transition leads to crizotinib resistance in H2228 lung cancer cells with EML4-ALK translocation. Mol Oncol 7: 1093–1102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Kim S., Kim T., Kim D., Go H., Keam B., Lee S., et al. (2013b) Heterogeneity of genetic changes associated with acquired crizotinib resistance in ALK-rearranged lung cancer. J Thorac Oncol 8: 415–422. [DOI] [PubMed] [Google Scholar]
  44. Kim Y., Ozasa H., Nagai H., Sakamori Y., Yoshida H., Yagi Y., et al. (2013c) High-dose crizotinib for brain metastases refractory to standard-dose crizotinib. J Thorac Oncol 8: e85–e86. [DOI] [PubMed] [Google Scholar]
  45. Kinoshita K., Asoh K., Furuichi N., Ito T., Kawada H., Hara S., et al. (2012) Design and synthesis of a highly selective, orally active and potent anaplastic lymphoma kinase inhibitor (CH5424802). Bioorg Med Chem 20: 1271–1280. [DOI] [PubMed] [Google Scholar]
  46. Kobayashi Y., Sakao Y., Ito S., Park J., Kuroda H., Sakakura N., et al. (2013) Transformation to sarcomatoid carcinoma in ALK-rearranged adenocarcinoma, which developed acquired resistance to crizotinib and received subsequent chemotherapies. J Thorac Oncol 8: e75–e78. [DOI] [PubMed] [Google Scholar]
  47. Kodama T., Hasegawa M., Takanashi K., Sakurai Y., Kondoh O., Sakamoto H. (2014a) Antitumor activity of the selective ALK inhibitor alectinib in models of intracranial metastases. Cancer Chemother Pharmacol 74: 1023–1028. [DOI] [PubMed] [Google Scholar]
  48. Kodama T., Tsukaguchi T., Yoshida M., Kondoh O., Sakamoto H. (2014b) Selective ALK inhibitor alectinib with potent antitumor activity in models of crizotinib resistance. Cancer Lett 351: 215–221. [DOI] [PubMed] [Google Scholar]
  49. Koivunen J., Mermel C., Zejnullahu K., Murphy C., Lifshits E., Holmes A., et al. (2008) EML4-ALK fusion gene and efficacy of an ALK kinase inhibitor in lung cancer. Clin Cancer Res 14: 4275–4283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Kwak E., Bang Y., Camidge D., Shaw A., Solomon B., Maki R., et al. (2010) Anaplastic lymphoma kinase inhibition in non–small-cell lung cancer. N Engl J Med 363: 1693–1703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Lee J., Kim T., Lee S., Kim D., Kim S., Jeon Y., et al. (2011) Anaplastic lymphoma kinase translocation: a predictive biomarker of pemetrexed in patients with non-small cell lung cancer. J Thorac Oncol 6: 1474–1480. [DOI] [PubMed] [Google Scholar]
  52. Lin Y., Wang Y., Yang J., Yu C., Wu S., Shih J., et al. (2014) Development of renal cysts after crizotinib treatment in advanced ALK-positive non–small-cell lung cancer. J Thorac Oncol 9: 1720–1725. [DOI] [PubMed] [Google Scholar]
  53. Lovly C., Heuckmann J., De Stanchina E., Chen H., Thomas R., Liang C., et al. (2011) Insights into ALK-driven cancers revealed through development of novel ALK tyrosine kinase inhibitors. Cancer Res 71: 4920–4931. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Lovly C., McDonald N., Chen H., Ortiz-Cuaran S., Heukamp L., Yan Y., et al. (2014) Rationale for co-targeting IGF-1R and ALK in ALK fusion-positive lung cancer. Nat Med 20: 1027–1034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Lovly C., Pao W. (2012) Escaping ALK inhibition: mechanisms of and strategies to overcome resistance. Sci Transl Med 4: 120ps2. [DOI] [PubMed] [Google Scholar]
  56. Maemondo M., Inoue A., Kobayashi K., Sugawara S., Oizumi S., Isobe H., et al. (2010) Gefitinib or chemotherapy for non–small-cell lung cancer with mutated EGFR. N Engl J Med 362: 2380–2388. [DOI] [PubMed] [Google Scholar]
  57. Maillet D., Martel-Lafay I., Arpin D., Pérol M. (2013) Ineffectiveness of crizotinib on brain metastases in two cases of lung adenocarcinoma with EML4-ALK rearrangement. J Thorac Oncol 8: e30–e31. [DOI] [PubMed] [Google Scholar]
  58. Maitland M., Ou S., Tolcher A., Lorusso P., Bahceci E., Ball H., et al. (2014) Safety, activity, and pharmacokinetics of an oral anaplastic lymphoma kinase (ALK) inhibitor, ASP3026, observed in a “fast follower” phase 1 trial design. ASCO Meeting Abstracts 32: abstract 2624. [Google Scholar]
  59. Marsilje T., Pei W., Chen B., Lu W., Uno T., Jin Y., et al. (2013) Synthesis, structure–activity relationships, and in vivo efficacy of the novel potent and selective anaplastic lymphoma kinase (ALK) inhibitor 5-chloro-N2-(2-isopropoxy-5-methyl-4-(piperidin-4-yl)phenyl)-N4-(2-(isopropylsulfonyl)phenyl)pyrimidine-2,4-diamine (LDK378) currently in phase 1 and phase 2 clinical trials. J Med Chem 56: 5675–5690. [DOI] [PubMed] [Google Scholar]
  60. Mok T., Wu Y., Thongprasert S., Yang C., Chu D., Saijo N., et al. (2009) Gefitinib or carboplatin–paclitaxel in pulmonary adenocarcinoma. N Engl J Med 361: 947–957. [DOI] [PubMed] [Google Scholar]
  61. Mori M., Ueno Y., Konagai S., Fushiki H., Shimada I., Kondoh Y., et al. (2014) The selective anaplastic lymphoma receptor tyrosine kinase inhibitor ASP3026 induces tumor regression and prolongs survival in non–small cell lung cancer model mice. Mol Cancer Ther 13: 329–340. [DOI] [PubMed] [Google Scholar]
  62. Nakagawa K., Hida T., Seto T., Satouchi M., Nishio M., Hotta K., et al. (2014) Antitumor activity of alectinib (CH5424802/RO5424802) for ALK-rearranged NSCLC with or without prior crizotinib treatment in bioequivalence study. ASCO Meeting Abstracts 32: abstract 8103. [Google Scholar]
  63. Neckers L., Workman P. (2012) Hsp90 molecular chaperone inhibitors: are we there yet? Clin Cancer Res 18: 64–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Nogova L., Mattonet C., Scheffler M., Michels S., Bos M., Heukamp L., et al. (2014) 1238PTRY - a phase II study to evaluate safety and efficacy of combined trastuzumab and the Hsp90 inhibitor AUY922 in advanced non-small-cell lung cancer (NSCLC) with HER2 overexpression or amplification or mutation. Ann Oncol 25(Suppl. 4): iv433. [Google Scholar]
  65. Normant E., Paez G., West K., Lim A., Slocum K., Tunkey C., et al. (2011) The Hsp90 inhibitor IPI-504 rapidly lowers EML4-ALK levels and induces tumor regression in ALK-driven NSCLC models. Oncogene 30: 2581–2586. [DOI] [PubMed] [Google Scholar]
  66. Ou S., Azada M., Dy J., Stiber J. (2011a) Asymptomatic profound sinus bradycardia (heart rate ⩽45) in non-small cell lung cancer patients treated with crizotinib. J Thorac Oncol 6: 2135–2137. [DOI] [PubMed] [Google Scholar]
  67. Ou S., Jänne P., Bartlett C., Tang Y., Kim D., Otterson G., et al. (2014a) Clinical benefit of continuing ALK inhibition with crizotinib beyond initial disease progression in patients with advanced ALK-positive NSCLC. Ann Oncol 25: 415–422. [DOI] [PubMed] [Google Scholar]
  68. Ou S., Azada M., Hsiang D., Herman J., Kain T., Siwak-Tapp C., et al. (2014b) Nextgeneration sequencing reveals a Novel NSCLC ALK F1174V mutation and confirms ALK G1202R mutation confers high-level resistance to alectinib (CH5424802/RO5424802) in ALK-rearranged NSCLC patients who progressed on crizotinib. J Thorac Oncol 9: 549–553. [DOI] [PubMed] [Google Scholar]
  69. Ou S., Kwak E., Siwak-Tapp C., Dy J., Bergethon K., Clark J., et al. (2011b) Activity of crizotinib (PF02341066), a dual mesenchymal-epithelial transition (MET) and anaplastic lymphoma kinase (ALK) inhibitor, in a non-small cell lung cancer patient with de novo MET amplification. J Thorac Oncol 6: 942–946. [DOI] [PubMed] [Google Scholar]
  70. Ou S., Tong W., Azada M., Siwak-Tapp C., Dy J., Stiber J. (2013) Heart rate decrease during crizotinib treatment and potential correlation to clinical response. Cancer 119: 1969–1975. [DOI] [PubMed] [Google Scholar]
  71. Park H., Lee J., Kim D., Kulig K., Kim T., Lee S., et al. (2012) Immunohistochemical screening for anaplastic lymphoma kinase (ALK) rearrangement in advanced non-small cell lung cancer patients. Lung Cancer 77: 288–292. [DOI] [PubMed] [Google Scholar]
  72. Pillai R., Ramalingam S. (2014) Heat shock protein 90 inhibitors in non-small-cell lung cancer. Curr Opin Oncol 26: 159–164. [DOI] [PubMed] [Google Scholar]
  73. Politi K., Gettinger S. (2014) Perfect alkemy: optimizing the use of ALK-directed therapies in lung cancer. Clin Cancer Res 20: 5576–5578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Proia D., Sang J., He S., Smith D., Sequeira M., Zhang C., et al. (2012) Synergistic activity of the Hsp90 inhibitor ganetespib with taxanes in non-small cell lung cancer models. Invest New Drugs 30: 2201–2209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Ramalingam S., Goss G., Andric Z., Bondarenko I., Zaric B., Ceric T., et al. (2013) A randomised study of ganetespib, a heat shock protein 90 inhibitor, in combination with docetaxel versus docetaxel alone for second-line therapy of lung adenocarcinoma (GALAXY-1). ASCO Meeting Abstracts 31: abstract CRA8007. [DOI] [PubMed] [Google Scholar]
  76. Ramalingam S., Zaric B., Ceric T., Ciuleanu T., Spicer J., Bondarenko I., et al. (2014) Galaxy-2 trial (NCT01798485): a randomised phase 3 study of ganetespib in combination with docetaxel versus docetaxel alone in patients with advanced lung adenocarcinoma. ASCO Meeting Abstracts 32: abstract TPS8118. [Google Scholar]
  77. Rivera V., Wang F., Anjum R., Zhang S., Squillace R., Keats J., et al. (2012) AP26113 is a dual ALK/EGFR inhibitor: characterization against EGFR T790M in cell and mouse models of NSCLC [AACR abstract 1794]. Cancer Res 72: 1794. [Google Scholar]
  78. Rodig S., Mino-Kenudson M., Dacic S., Yeap B., Shaw A., Barletta J., et al. (2009) Unique clinicopathologic features characterize ALK-rearranged lung adenocarcinoma in the western population. Clin Cancer Res 15: 5216–5223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Rothenstein J., Letarte N. (2014) Managing treatment–related adverse events associated with ALK inhibitors. Curr Oncol 21: 19–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Sakamoto H., Tsukaguchi T., Hiroshima S., Kodama T., Kobayashi T., Fukami T., et al. (2011) CH5424802, a selective ALK inhibitor capable of blocking the resistant gatekeeper mutant. Cancer Cell 19: 679–690. [DOI] [PubMed] [Google Scholar]
  81. Sang J., Acquaviva J., Friedland J., Smith D., Sequeira M., Zhang C., et al. (2013) Targeted inhibition of the molecular chaperone Hsp90 overcomes ALK inhibitor resistance in non–small cell lung cancer. Cancer Discov 3: 430–443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Sasaki T., Koivunen J., Ogino A., Yanagita M., Nikiforow S., Zheng W., et al. (2011) A novel ALK secondary mutation and EGFR signaling cause resistance to ALK kinase inhibitors. Cancer Res 71: 6051–6060. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Sasaki T., Okuda K., Zheng W., Butrynski J., Capelletti M., Wang L., et al. (2010) The neuroblastoma-associated F1174l ALK mutation causes resistance to an ALK kinase inhibitor in ALK-translocated cancers. Cancer Res 70: 10038–10043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Sequist L., Gettinger S., Senzer N., Martins R., Jänne P., Lilenbaum R., et al. (2010) Activity of IPI-504, a novel heat-shock protein 90 inhibitor, in patients with molecularly defined non–small-cell lung cancer. J Clin Oncol 28: 4953–4960. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Sequist L., Yang J., Yamamoto N., O’Byrne K., Hirsh V., Mok T., et al. (2013) Phase III study of afatinib or cisplatin plus pemetrexed in patients with metastatic lung adenocarcinoma with EGFR mutations. J Clin Oncol 31: 3327–3334. [DOI] [PubMed] [Google Scholar]
  86. Seto T., Kiura K., Nishio M., Nakagawa K., Maemondo M., Inoue A., et al. (2013) CH5424802 (RO5424802) for patients with ALK-rearranged advanced non-small-cell lung cancer (AF-001JP study): a single-arm, open-label, phase 1-2 study. Lancet Oncol 14: 590–598. [DOI] [PubMed] [Google Scholar]
  87. Shanthi E., Krishna M., Arunesh G., Venkateswara Reddy K., Sooriya Kumar J., Viswanadhan V. (2014) Focal adhesion kinase inhibitors in the treatment of metastatic cancer: a patent review. Expert Opin Ther Pat 24: 1077–1100. [DOI] [PubMed] [Google Scholar]
  88. Shaw A., Kim D., Mehra R., Tan D., Felip E., Chow L., et al. (2014) Ceritinib in ALK-rearranged non–small-cell lung cancer. N Engl J Med 370: 1189–1197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Shaw A., Kim D., Nakagawa K., Seto T., Crinó L., Ahn M., et al. (2013) Crizotinib versus chemotherapy in advanced ALK-positive lung cancer. N Engl J Med 368: 2385–2394. [DOI] [PubMed] [Google Scholar]
  90. Shaw A., Yeap B., Mino-Kenudson M., Digumarthy S., Costa D., Heist R., et al. (2009) Clinical features and outcome of patients with non–small-cell lung cancer who harbor EML4-ALK. J Clin Oncol 27: 4247–4253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Shimamura T., Perera S., Foley K., Sang J., Rodig S., Inoue T., et al. (2012) Ganetespib (STA-9090), a nongeldanamycin Hsp90 inhibitor, has potent antitumor activity in in vitro and in vivo models of non–small cell lung cancer. Clin Cancer Res 18: 4973–4985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Socinski M., Goldman J., El-Hariry I., Koczywas M., Vukovic V., Horn L., et al. (2013) A multicenter phase II study of ganetespib monotherapy in patients with genotypically defined advanced non–small cell lung cancer. Clin Cancer Res 19: 3068–3077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Soda M., Choi Y., Enomoto M., Takada S., Yamashita Y., Ishikawa S., et al. (2007) Identification of the transforming EML4-ALK fusion gene in non-small-cell lung cancer. Nature 448: 561–566. [DOI] [PubMed] [Google Scholar]
  94. Soda M., Takada S., Takeuchi K., Choi Y., Enomoto M., Ueno T., et al. (2008) A mouse model for EML4-ALK-positive lung cancer. Proc Natl Acad Sci U S A 105: 19893–19897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Solomon B., Mok T., Kim D., Wu Y., Nakagawa K., Mekhail T., et al. (2014) First-line crizotinib versus chemotherapy in ALK-positive lung cancer. N Engl J Med 371: 2167–2177. [DOI] [PubMed] [Google Scholar]
  96. Squillace R., Anjum R., Miller D., Vodala S., Moran L., Wang F., et al. (2013) AP26113 possesses pan-inhibitory activity versus crizotinib-resistant ALK mutants and oncogenic ROS1 fusions [AACR abstract 5655]. Cancer Res 73. [Google Scholar]
  97. Takeda M., Okamoto I., Nakagawa K. (2013) Clinical impact of continued crizotinib administration after isolated central nervous system progression in patients with lung cancer positive for ALK rearrangement. J Thorac Oncol 8: 654–657. [DOI] [PubMed] [Google Scholar]
  98. Takeuchi K., Choi Y., Togashi Y., Soda M., Hatano S., Inamura K., et al. (2009) KIF5B-ALK, a novel fusion oncokinase identified by an immunohistochemistry-based diagnostic system for ALK-positive lung cancer. Clin Cancer Res 15: 3143–3149. [DOI] [PubMed] [Google Scholar]
  99. Tamiya A., Okamoto I., Miyazaki M., Shimizu S., Kitaichi M., Nakagawa K. (2013) Severe acute interstitial lung disease after crizotinib therapy in a patient with EML4-ALK positive non-small cell lung cancer. J Clin Oncol 31: e15–e17. [DOI] [PubMed] [Google Scholar]
  100. Tanizaki J., Okamoto I., Okabe T., Sakai K., Tanaka K., Hayashi H., et al. (2012) Activation of her family signaling as a mechanism of acquired resistance to ALK inhibitors in EML4-ALK–positive non–small cell lung cancer. Clin Cancer Res 18: 6219–6226. [DOI] [PubMed] [Google Scholar]
  101. Togashi Y., Soda M., Sakata S., Sugawara E., Hatano S., Asaka R., et al. (2012) KLC1-ALK: a novel fusion in lung cancer identified using a formalin-fixed paraffin-embedded tissue only. PLoS ONE 7: e31323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Trepel J., Mollapour M., Giaccone G., Neckers L. (2010) Targeting the dynamic Hsp90 complex in cancer. Nat Rev Cancer 10: 537–549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Vaishnavi A., Capelletti M., Le A., Kako S., Butaney M., Ercan D., et al. (2013) Oncogenic and drug-sensitive NTRK1 rearrangements in lung cancer. Nat Med 19: 1469–1472. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Wang W., Shiao H., Lee C., Fung K., Hsieh H. (2014) Anaplastic lymphoma kinase (ALK) inhibitors: a review of design and discovery. Med Chem Comm 5: 1266–1279. [Google Scholar]
  105. Weickhardt A., Doebele R., Purcell W., Bunn P., Oton A., Rothman M., et al. (2013) Symptomatic reduction in free testosterone levels secondary to crizotinib use in male cancer patients. Cancer 119: 2383–2390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Weickhardt A., Rothman M., Salian-Mehta S., Kiseljak-Vassiliades K., Oton A., Doebele R., et al. (2012) Rapid-onset hypogonadism secondary to crizotinib use in men with metastatic nonsmall cell lung cancer. Cancer 118: 5302–5309. [DOI] [PubMed] [Google Scholar]
  107. Weiss G., Sachdev J., Infante J., Mita M., Natale R., Arkenau H., et al. (2014) Phase (Ph) 1/2 study of TSR-011, a potent inhibitor of ALK and TRK, including crizotinib-resistant ALK mutations. ASCO Meeting Abstracts 32: abstract e19005. [Google Scholar]
  108. Whitesell L., Lindquist S. (2005) Hsp90 and the chaperoning of cancer. Nat Rev Cancer 5: 761–772. [DOI] [PubMed] [Google Scholar]
  109. Wilcoxen K., Brake R., Saffran D., Teffera Y., Choquette D., Whittington D., et al. (2012) Characterization of a novel series of potent, selective inhibitors of wild type and mutant/fusion anaplastic lymphoma kinase [AACR abstract 1795]. Cancer Res 72: 1795. [Google Scholar]
  110. Wong D., Leung E., So K., Tam I., Sihoe A., Cheng L., et al. (2009) The EML4-ALK fusion gene is involved in various histologic types of lung cancers from nonsmokers with wild-type EGFR and KRAS. Cancer 115: 1723–1733. [DOI] [PubMed] [Google Scholar]
  111. Woodhead A., Angove H., Carr M., Chessari G., Congreve M., Coyle J., et al. (2010) Discovery of (2,4-dihydroxy-5-isopropylphenyl)-[5-(4-methylpiperazin-1-ylmethyl)-1,3-dihydroisoindol-2-yl]methanone (AT13387), a novel inhibitor of the molecular chaperone Hsp90 by fragment based drug design. J Med Chem 53: 5956–5969. [DOI] [PubMed] [Google Scholar]
  112. Wu S., Kuo Y., Chang Y., Shih J., Chen Y., Tsai M., et al. (2012) EML4-ALK translocation predicts better outcome in lung adenocarcinoma patients with wild-type EGFR. J Thorac Oncol 7: 98–104. [DOI] [PubMed] [Google Scholar]
  113. Yamada T., Takeuchi S., Nakade J., Kita K., Nakagawa T., Nanjo S., et al. (2012) Paracrine receptor activation by microenvironment triggers bypass survival signals and ALK inhibitor resistance in EML4-ALK lung cancer cells. Clin Cancer Res 18: 3592–3602. [DOI] [PubMed] [Google Scholar]
  114. Yamaguchi N., Lucena-Araujo A., Nakayama S., De Figueiredo-Pontes L., Gonzalez D., Yasuda H., et al. (2014) Dual ALK and EGFR inhibition targets a mechanism of acquired resistance to the tyrosine kinase inhibitor crizotinib in ALK rearranged lung cancer. Lung Cancer 83: 37–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Yang J. (2013) A selective ALK inhibitor in ALK-rearranged patients. Lancet Oncol 14: 564–565. [DOI] [PubMed] [Google Scholar]
  116. Yang J., Hirsh V., Schuler M., Yamamoto N., O’Byrne K., Mok T., et al. (2013) Symptom control and quality of life in LUX-Lung 3: a phase III study of afatinib or cisplatin/pemetrexed in patients with advanced lung adenocarcinoma with EGFR mutations. J Clin Oncol 31: 3342–3350. [DOI] [PubMed] [Google Scholar]
  117. Yasuda H., De Figueiredo-Pontes L., Kobayashi S., Costa D. (2012) Preclinical rationale for use of the clinically available multitargeted tyrosine kinase inhibitor crizotinib in ROS1-translocated lung cancer. J Thorac Oncol 7: 1086–1090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Yi E., Boland J., Maleszewski J., Roden A., Oliveira A., Aubry M., et al. (2011) Correlation of IHC and fish for ALK gene rearrangement in non-small cell lung carcinoma: IHC ccore algorithm for fish. J Thorac Oncol 6: 459–465. [DOI] [PubMed] [Google Scholar]
  119. Ying W., Du Z., Sun L., Foley K., Proia D., Blackman R., et al. (2012) Ganetespib, a unique triazolone-containing Hsp90 inhibitor, exhibits potent antitumor activity and a superior safety profile for cancer therapy. Mol Cancer Ther 11: 475–484. [DOI] [PubMed] [Google Scholar]
  120. Zhang S., Wang F., Keats J., Ning Y., Wardwell S., Moran L., et al. (2010) AP26113, a potent ALK inhibitor, overcomes mutations in EML4-ALK that confer resistance to PF-02341066 [AACR abstract 3623]. Cancer Res 70: LB-298. [Google Scholar]
  121. Zhang S., Wang F., Keats J., Zhu X., Ning Y., Wardwell S., et al. (2011) Crizotinib-resistant mutants of EML4-ALK identified through an accelerated mutagenesis screen. Chem Biol Drug Des 78: 999–1005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Zou H., Engstrom L., Li Q., West Lu M., Tang R., Wang H., et al. (2013a) Abstract A277: PF-06463922, a novel ROS1/ALK inhibitor, demonstrates sub-nanomolar potency against oncogenic ROS1 fusions and capable of blocking the resistant ROS1G2032R mutant in preclinical tumor models. Mol Cancer Ther 12: A277. [Google Scholar]
  123. Zou H., Engstrom L., Li Q., West Lu M., Tang R., Wang H., et al. (2013b) Abstract C253: PF-06463922, a novel brain-penetrating small molecule inhibitor of ALK/ROS1 with potent activity against a broad spectrum of ALK resistant mutations in preclinical models in vitro and in vivo. Mol Cancer Ther 12: C253. [Google Scholar]

Articles from Therapeutic Advances in Medical Oncology are provided here courtesy of SAGE Publications

RESOURCES