Review Open Access
Copyright ©The Author(s) 2015. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Transl Med. Apr 12, 2015; 4(1): 25-37
Published online Apr 12, 2015. doi: 10.5528/wjtm.v4.i1.25
Therapeutic targets in gastrointestinal stromal tumors
Jia-Qing Zhu, Wen-Bin Ou, College of Life Sciences, Zhejiang Sci-Tech University, Hangzhou 310018, Zhejiang Province, China
Wen-Bin Ou, Zhejiang Provincial Key Laboratory of Applied Enzymology, Yangtze Delta Region Institute of Tsinghua University, Jiaxing 314006, Zhejiang Province, China
Author contributions: Zhu JQ preformed research, analyzed data, consulted literatures and wrote the paper; Ou WB consulted literatures, provided ideas, designed research, contributed new reagents, analyzed data, and wrote the paper.
Supported by The Special Project of Zhejiang Province, No. 2012C03007-4; Zhejiang Public Technology Research Program, No. 2014C33234; Zhejiang Provincial Top Key Discipline of Biology, and Science Foundation of Zhejiang Sci-Tech University, No. 14042107-Y.
Conflict-of-interest: These authors declare no conflict of interest.
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Correspondence to: Dr. Wen-Bin Ou, College of Life Sciences, Zhejiang Sci-Tech University, 866 Yuhangtang Road, Hangzhou 310018, Zhejiang Province, China. ouwenbin@tsinghua.org.cn
Telephone: +86-573-82582765 Fax: +86-573-82582765
Received: July 26, 2014
Peer-review started: July 27, 2014
First decision: September 1, 2014
Revised: September 14, 2014
Accepted: November 27, 2014
Article in press: December 1, 2014
Published online: April 12, 2015

Abstract

Gastrointestinal stromal tumors (GISTs) are the most common type of mesenchymal tumor of the gastrointestinal tract. The tumorigenesis of GISTs is driven by gain-of-function mutations in KIT or platelet-derived growth factor receptor α (PDGFRA), resulting in constitutive activation of the tyrosine kinase and its downstream signaling pathways. Oncogenic KIT or PDGFRA mutations are compelling therapeutic targets for the treatment of GISTs, and the KIT/PDGFRA inhibitor imatinib is the standard of care for patients with metastatic GISTs. However, most GIST patients develop clinical resistance to imatinib and other tyrosine kinase inhibitors. Five mechanisms of resistance have been characterized: (1) acquisition of a secondary point mutation in KIT or PDGFRA; (2) genomic amplification of KIT; (3) activation of an alternative receptor tyrosine kinase; (4) loss of KIT oncoprotein expression; and (5) wild-type GIST. Currently, sunitinib is used as a second-line treatment for patients after imatinib failure, and regorafenib has been approved for patients whose disease is progressing on both imatinib and sunitinib. Phase II/III trials are currently in progress to evaluate novel inhibitors and immunotherapies targeting KIT, its downstream effectors such as phosphatidylinositol 3-kinase, protein kinase B and mammalian target of rapamycin, heat shock protein 90, and histone deacetylase inhibitor. Other candidate targets have been identified, including ETV1, AXL, insulin-like growth factor 1 receptor, KRAS, FAS receptor, protein kinase c theta, ANO1 (DOG1), CDC37, and aurora kinase A. These candidates warrant clinical evaluation as novel therapeutic targets in GIST.

Key Words: Gastrointestinal stromal tumors, Tyrosine kinase inhibitors, KIT, Platelet-derived growth factor receptor α, Targets

Core tip: Oncogenic KIT and platelet-derived growth factor receptor α (PDGFRA) mutations are compelling therapeutic targets in gastrointestinal stromal tumors (GISTs), and the KIT/PDGFRA kinase inhibitors imatinib, sunitinib, and regorafenib are the standards of care for patients with unresectable or metastatic GIST. However, most patients eventually develop resistance to these kinase inhibitors, resulting in an urgent need to identify biologically rational targets for novel therapies. Herein, we review advances in the research on GIST and the therapies that are used to treat it. Additionally, we discuss novel agents, targets, and strategies for the future treatment of GIST.



INTRODUCTION

Gastrointestinal stromal tumors (GISTs) were originally described as smooth muscle or neural tumors of the gastrointestinal (GI) tract; however, in 1983, Mazur et al[1] referred to GISTs as “stromal tumors”[2,3]. Subsequent studies identified the interstitial cells of Cajal as the origin of GISTs. In 1998, activating mutations of the KIT receptor tyrosine kinase (RTK) were found in GISTs[4]. In 2003, platelet-derived growth factor receptor α (PDGFRA) mutations, an alternative target, were identified in GISTs that lacked KIT mutations[5].

GISTs are the most common mesenchymal tumors of the GI tract and are frequently seen in the stomach (60%), small intestine (25%), colorectum (5%-10%) and occasionally in the esophagus and appendix[5]. Histologically, GISTs may be composed of spindle cells (70%), epithelioid cells (20%), or a mixture of these types (10%)[6]. Morphologically, GISTs may be mistaken for smooth muscle neoplasms, such as leiomyoma and leiomyosarcoma (Figure 1)[6]. Consensus guidelines for GIST prognosis, accentuate risk stratification based on the tumor volume and mitotic index of the primary tumors (Table 1)[2].

Table 1 Risk stratification of primary gastrointestinal stromal tumor by mitotic index, size and anatomic location[2].
Prognosis of primary GIST
Risk
Size (cm)
Mitotic count (per 50 HPF)
Very low risk< 2< 5
Low risk2-5< 5
Intermediate risk< 56-10
5-10< 5
High risk> 5> 5
> 10> Any mitotic rate
Any tumor> 10
Figure 1
Figure 1 Morphologic similarities of low-risk gastrointestinal stromal tumor and leiomyoma and of a high-risk gastrointestinal stromal tumor and leiomyosarcoma. GIST cells can divided into 3 types: spindle cell (70% of cases), epithelioid cell (20% of cases), and mixed cell (containing a mixture of spindle and epithelioid cells). GIST: Gastrointestinal stromal tumor.

The majority of GISTs contain oncogenic mutations of KIT (approximately 85%) or PDGFRA (approximately 5%-10%)[2,4-6]. The resulting mutant oncoproteins are crucial for GIST oncogenesis, proliferation, and survival, as demonstrated by the clinical successes of small molecule therapeutics targeting KIT and PDGFRA[7-9]. Imatinib, sunitinib, and regorafenib are the standard first-, second- and third-line therapies, respectively, in patients with inoperable GISTs[10-12], and adjuvant imatinib is used in patients with localized GISTs with a high risk of recurrence[13].

Except from imatinib, sunitinib, and regorafenib, which target the activated oncoproteins KIT and PDGFRA in inoperable or metastatic GIST, the increasing novel drugs are currently in clinical trials, and additional potential therapeutic targets have been identified. Herein, we summarize these agents, targets, and strategies for the future treatment of GIST.

KIT AND PDGFRA ARE MAJOR THERAPEUTIC TARGETS IN GISTS

Oncogenic mutant KIT and PDGFRA play a critical function in the initiation of the transformation event that leads to GIST. Mutations in KIT are usually found in the regulatory and dimerization domains, which are located in the extracellular region encoded by exon 9 (approximately 13% of GISTs), the juxtamembrane region encoded by exon 11 (approximately 66% of GISTs), or the tyrosine kinase (TK)[I] [adenosine triphosphate (ATP) binding pocket]; and TK[II] (activation loop) domains encoded by exon 13 (approximately 1% of GISTs) and exon 17 (approximately 0.6% of GISTs), respectively[2,14,15]. Five percent to ten percent of GISTs contain mutations in PDGFRA exon 12 (juxtamembrane region) (1.5%) or exon 18 (activation loop) (5.6%). The remainder (10%-12%) are wild-type for both KIT and PDGFRA[2,6]. The percentage of population of KIT and PDGFRA mutations is shown in Figure 2[2].

Figure 2
Figure 2 Schematic structure of KIT and platelet-derived growth factor receptor α receptor tyrosine kinases and distribution of KIT mutations in gastrointestinal stromal tumor. EC: Extracellular; JM: Juxtamembrane; TK[I]: Tyrosine kinase domain I; EX: Exon.

GISTs harboring insertions, deletions, and missense mutations in KIT exon 11 can be found throughout the GI tract[16]. A enhanced metastasis and proliferation has been associated with loss of heterozygosity at the KIT locus[17,18]. The vast majority of GIST cases with alterations of KIT in exon 9 involve an insertion of six base pairs, resulting in the duplication of Ala and Tyr residues. These mutations often occur in high-risk primary GISTs of the small intestine[17,19,20], advanced or relapsed GISTs[18,21]. A recent study demonstrated that GISTs harboring KIT exon 17 and exon 13 mutations show slightly overrun population among a subset of GISTs. Most of single base pair substitution KIT mutations in exon 13 and 17 in small intestinal GISTs, have no marked effects on the clinicopathologic characteristics when compared to the “average” small intestinal GIST[22].

The majority of PDGFRA exon 14 and 18 alterations are missense mutations. GISTs harboring PDGFRA mutations are confined to the stomach and omentum. These tumors are shortage of KIT expression, they typically present an epithelioid morphology, and they are commonly associated with a benign prognosis[23,24]. GISTs harboring a D842V PDGFRA exon 18 mutation are resistant to imatinib and other RTK inhibitors[25-28].

Inhibition of KIT or PDGFRA kinase activity by imatinib results in an objective response in approximately 80% of metastatic GIST patients (approximately 50% partial response, approximately 30% stable) with a 3-year survival rate of 69%-74%[8]. However, the median survival of metastatic GIST patients was 19 mo in the pre-imatinib period[10,15]. Constitutive activation of KIT or PDGFRA results in the activation of downstream signaling intermediates necessary for proliferation, survival, adhesion, and blockage of differentiation, including the phosphatidylinositol 3-kinase (PI3K)/protein kinase B (AKT)/mammalian target of rapamycin (mTOR) and RAF/mitogen-activated protein kinase (MAPK) pathways. Targeting KIT/PDGFRA and its downstream intermediates has proven to be an effective strategy in the treatment of GISTs[29-32].

MECHANISMS OF IMATINIB RESISTANCE

Imatinib, an ATP-competitive inhibitor of KIT and PDGFRA, is the first-line therapy for patients with advanced GIST or primary GIST with a significant risk of recurrence after surgery[28,33-35]. Among patients with advanced GIST, 75% to 90% will show a response to imatinib[15]. Analysis of the crystal structure of the KIT-imatinib complex reveals that the drug fills a hydrophobic region of the ATP binding pocket, effectively blocking ATP binding and inactivating KIT and its downstream signaling[36,37].

Despite the dramatic clinical success of imatinib, most inoperable GIST patients eventually develop resistant to imatinib. Imatinib resistance in GIST is classified as either primary or secondary imatinib resistance. Approximately 10% of GISTs demonstrate primary imatinib resistance of clinical progression within 3 to 6 mo of the start of treatment[28,38]. Primary imatinib resistance is usually observed in tumors that lack KIT or PDGFRA mutations (wild-type GISTs), but it is also common in tumors harboring KIT exon 9 mutations[28,38]. Approximately 40% to 50% of GIST patients experience secondary imatinib resistance of clinical progression after 12-36 mo of response or disease stabilization. Molecular studies showed that activated KIT expression in imatinib-resistant tumors was similar to or greater than those typically found in untreated GISTs[15]. Secondary KIT mutations were rare in GISTs with primary resistance but often found in GISTs with secondary resistance (10% vs 67%; P = 0.002). Polyclonal secondary kinase mutation was detected in 18.8% patients. The secondary kinase mutations were nonrandomly distributed and were associated with attenuated imatinib sensitivity compared with KIT exon 9 and exon 11[15]. Mechanisms of acquired resistance include secondary mutations in KIT or PDGFRA, genomic amplification of KIT, or activation of an alternative RTK[6,14,39-46]. An even more challenging resistance mechanism, seen in approximately 5%-10% of clinically progressing KIT-mutant GISTs involves a transition from dependence on oncogenic KIT to a new imatinib-insensitive oncogenic driver, accompanied by the loss of former KIT expression[39,40].

NOVEL INHIBITORS IN PRE-CLINICAL MODELS AND CLINICAL TRIALS

Tumorigenesis is a complex, multi-step process, and oncogenic RTK proteins frequently play key roles[47] (Table 2). Oncogenic RTK mutations can lead to constitutive kinase activation and thereby enhance growth and survival in cancer cells[48,49]. Tyrosine kinases can be divided into two categories: receptor tyrosine kinases and non-receptor tyrosine kinases. At present, approximately 90 types of TK members have been identified, including 58 RTKs, such as PDGFR, epidermal growth factor receptor (EGFR), fibroblast growth factor receptor, and 32 non-RTKs[50]. Oncogenic RTK mutants are useful therapeutic targets, as shown by the clinical benefit of small molecular inhibitor therapies in chronic myeloid leukemia (BCR-ABL)[51], metastatic breast cancer [human epidermal growth factor receptor 2 (HER2)][52], GIST (KIT/PDGFRA)[8], and non-small-cell lung cancer [EGFR, hepatocyte growth factor receptor (MET), anaplasticlymphoma kinase, HER2][47,53-62].

Table 2 Novel agents are being developed for gastrointestinal stromal tumor therapy[10,13,14,21,30,64,67-84].
AgentMolecular targetPhase
Kinase inhibitors
NilotinibKIT, PDGFRs, BCR-ABLI
SorafenibRaf, KIT, PDGFRB, VEGFR, FLT3, RET71%
DasatinibSrc, ABL, KIT, PDGFRsPhase II ongoing in advanced sarcomas and accepting patients
Cediranib (AZD2171)VEGFR, KIT, PDGFRsPhase II ongoing
OSI-930VEGFR, KITPhase II ongoing, not recruiting
Linsitinib (OSI-906 )IGF1RPhase III
Vatalanib (PTK787)VEGFR, KIT, PDGFRs67%
Motesanib (AMG706)VEGFR, KIT, PDGFRs, RET24%-27%
XL820KIT, PDGFRB, VEGFRPhase II ongoing, not recruiting
mTOR and AKT inhibitors
PerifosineAKTPhase II ongoing in combination with imatinib
EverolimusmTOR26%
TemsirolimusmTORPhase II ongoing, closed recruitment
Hsp90 inhibitors
17-AAGHsp90Phase II/III
Ganetespib (STA-9090)Hsp90Phase II
AUY922Hsp90Phase II
AT13387Hsp90Phase II ongoing in combination with imatinib
IPI-504Hsp9078%, phase III ended due to safety concerns
Others
FlavopiridolTranscription inhibitorPhase I ongoing in combination with doxorubicin
Clinical benefit is defined as complete or partial response or stable disease

Sunitinib is an oral multi-target tyrosine kinase inhibitor (TKI) with activity against KIT, PDGFRA, FMS-Like Tyrosine Kinase 3, Vascular endothelial growth factor receptor (VEGFR), and orphan receptor tyrosine kinase[63]. Sunitinib is approved for use as a second-line therapy for patients with imatinib-resistant GIST[9,64,65]. A clinical benefit of sunitinib was seen in common primary GIST with KIT exon 9 (58%), KIT exon 11 (34%), and wild-type KIT/PDGFRA (56%)[9]. Progression-free survival (PFS) was greater improvement for patients with a wild-type genotype (P = 0.0356) or with primary KIT exon 9 mutations (P = 0.0005) than for those with KIT exon 11 mutations. Overall survival (OS) showed the similar pattern. The PFS and OS were greater improvement for patients with secondary KIT exon 13 or 14 mutations than for those with exon 17 or 18 mutations[9]. The safety and efficacy of regorafenib in metastatic or unresectable GIST patients after failure of imatinib and sunitinib were evaluated in phase III, and the results showed that regorafenib can markedly improve PFS compared with control in metastatic GIST patients with progression after standard treatments[12,66]. Currently, regorafenib has been approved for patients whose tumors are progressing on both imatinib and sunitinib. A large number of therapies are in various stages of pre-clinical and clinical trial development and are summarized in Table 2[10,13,14,21,30,64,67-84]. These therapies can be divided into four groups: TKIs, PI3K/mTOR inhibitors, heat shock protein 90 (HSP90) inhibitors, and others.

Multiple TKIs, including nilotinib, sorafenib, dasatinib, vatalanib, and motesanib, are being investigated as potential therapies for GIST. Nilotinib, an inhibitor of KIT, PDGFRA and BCR-ABL, has been shown to be active in a small series of imatinib-resistant and sunitinib-resistant GIST patients in a phase I study[67,71,74,85]. Sorafenib, an inhibitor of RAF kinase, VEGFR, PDGFR, and KIT, inhibited KIT activity in some KIT primary and secondary mutations in a phase II trial in imatinib- and sunitinib-resistant GIST[69,80,86,87]. Dasatinib, a dual SRC/ABL kinase inhibitor, binds and inactivates wild-type and mutant KIT regardless of the conformation of the KIT activation loop[42,43]. Linsitinib (OSI-906) is a selective inhibitor of insulin-like growth factor receptor (IGFR)/insulin receptor. The combination of imatinib and linsitinib has been shown to be effective in wild-type GIST with insulin-like growth factor 1 receptor (IGF1R) overexpression or amplification[88]. Vatalanib (PTK787) and motesanib (AMG706), multi-kinase inhibitors, have been evaluated in phase II trials for patients who are resistant to both imatinib and sunitinib[89,90]. Vatalanib has shown activity in patients with imatinib-resistant or both imatinib- and sunitinib-resistant GIST[89,90]. Motesanib treatment was shown to have acceptable toxicity, and it resulted in disease stabilization in GIST patients[82].

The PI3K/AKT/mTOR pathway is crucial for proliferation and survival in GIST[29,30,68,91-93]. Preclinical experiments have confirmed that targeting the PI3K/AKT/mTOR pathway is a rational therapeutic strategy. Early studies with mTOR inhibitors have shown limited success, possibly due to feedback activation of AKT after mTORC1 inhibition. Simultaneous targeting of multiple nodes in the PI3K/AKT/mTOR pathway prevents feedback activation and may translate into more complete pathway inhibition. A few therapies targeting this pathway are currently being evaluated in phase I and II clinical trials[94]. A number of drugs currently in development include inhibitors of pan-Class I PI3K (BKM120 and GDC0941), PI3K/mTOR (BEZ235, SF1126 and GDC0980), AKT (Perifosin), and mTOR (Everolimus/RAD001 and Temsirolimus). Additionally, combined inhibition of KIT and PI3K/AKT/mTOR results in a greater response compared to either intervention alone[73,94-97].

Heat shock proteins control the proper folding, function, and stabilization of various client proteins. HSP90 optimizes and maintains the folding and localization of many activated tyrosine kinases and also prevents proteasomal degradation[98]. HSP90 is abundant in eukaryotic cells, comprising up to 1%-2% of total cellular protein, and it plays key roles in regulating cell proliferation, differentiation, and apoptosis[99,100]. The HSP90 inhibitor 17-allylamino-17-demethoxygeldanamycin (17-AAG), a geldanamycin derivative[101], binds a ATP-interaction pocket in the HSP90 NH2-terminal domain[102] and shows anti-proliferative effects in various human cancers, where it can degrade HSP90-client oncoproteins with high selectivity[103,104]. Whereas the clinical application of 17-AAG has been hampered by its low water solubility, IPI-504, a 17-AAG derivative, exhibits improved aqueous solubility while maintaining the biological HSP90-inhibitory properties of 17-AAG[105]. Furthermore, clinical trials with new-generation synthetic HSP90 inhibitors are ongoing in various cancer types. HSP90 is an attractive target in GIST as it is a key chaperone for KIT and PDGFRA[79,106]. Targeting HSP90 results in pro-apoptotic and anti-proliferative effects in GIST and is associated with the inhibition of KIT and PDGFRA signaling[72,79,107,108]. Other HSP90 inhibitors are in development (NVP-AUY922, AT-13387, KW-2478, and SNX-5422) and show promise for GIST treatment, particularly in combination with TKI[109].

Other drugs are in various stages of development for the treatment of GIST. Flavopiridol, a transcription inhibitor, has been evaluated in an ongoing phase I trial in combination with doxorubicin[110]. Histone deacetylase inhibitors (HDACIs) alone or in combination with imatinib have shown pro-apoptotic and anti-proliferative effects in GIST and are associated with inhibition of KIT and a reduction in the expression and activities of downstream pathways[111].

NOVEL CANDIDATE THERAPEUTIC TARGETS

Other therapeutic targets have been identified for the treatment of GIST, including Ets Variant 1 (ETV1), AXL, FAS, IGF1R, protein kinase c theta (PKCθ), RAS, CDC37, cyclin D1, Dog1, and aurora kinase A. Inhibitors targeting these candidates are being developed, and some are being evaluated in clinical trials.

The E26 transformation-specific family member ETV1 is overexpressed in the GIST and is required in the development of both imatinib-sensitive and imatinib-resistant GIST[112-114]. ETV1 enhancer binding is a master regulator of an ICC-GIST-specific transcription network. Activated KIT cooperates with ETV1 to induce development of GIST, regulating the ETV1 transcriptional program by prolonging ETV1 protein stability through MAPK signaling[112,114]. Inhibition of ETV1 reduces the expression of KIT, reduces mutagenesis, and stabilizes the GIST genome, thereby inhibiting GIST growth and progression and inducing apoptosis.

AXL (UFO/ARK/Tyro), an RTK stimulated by its ligand growth arrest-specific 6, shows potent oncogenic and transforming activity in normal and cancer cells[115-117]. AXL also plays a role in tumor cell invasion, metastasis, and survival[41,118,119]. AXL is active in GIST metastases that lose KIT expression at the time of clinical progression on imatinib[41,120]. In KIT-independent GISTs, AXL knockdown results in upregulation of p21, p27 and p53 protein expression and shows anti-proliferative effects[120]. MP470, a KIT/AXL inhibitor, shows a synergistic cytotoxic effect in GIST cells when combined with docetaxel (taxotere)[41].

Fas and its ligand FasL belong to the tumor necrosis factor family of death receptors. Activation of Fas by FasL induces cell apoptosis through caspase 8 signaling. Down-regulation of Fas is associated with tumorgenesis[121,122]. Fas and FasL expression were positively correlated in primary GISTs, but there was no association KIT mutation status[123]. MegaFasL, a hexameric form of soluble FasL, is an active apoptosis-inducing agent and potentiated the apoptotic effects of imatinib in GIST cell lines[123].

The IGF/IGF1R signaling system has been implicated as a relevant therapeutic target in a variety of cancers. When IGF1 binds with IGF1R, it activates downstream signaling cascades, such as the PI3K/AKT/mTOR and RAF/MEK/MAPK pathways, to trigger protein synthesis, and it also activates anti-apoptotic and proliferative pathways[124-126]. Recent reports have shown that IGF1R is amplified in a subset of GISTs[127] and over-expressed in wild-type and pediatric GIST[88,128,129]. Recent studies have shown that the IGF/IGF1R pathway may be a promising therapeutic target for GIST[127,130-135].

PKCθ, a member of the protein kinase C family commonly expressed in T cells and myogenic cells[136,137], is expressed at high levels and activated in GIST irrespective of the KIT or PDGFRA status. Therefore, PKCθ serves as a diagnostic marker of GIST[138-141]. PKCθ knockdown is accompanied by inactivation of KIT in KIT+/PKCθ+ GIST cell lines. PKCθ knockdown resulted in inhibition of PI3K/AKT signaling, upregulation of pro-apoptotic proteins p21 and p27, cell cycle arrest, and apoptosis, recapitulating the effect of direct KIT targeting[142]. PKCθ is a compelling therapeutic target in GISTs, including those with mutations that confer resistance to KIT/PDGFRA inhibitors.

Wild-type GISTs often demonstrate primary imatinib resistance. In some cases, these tumors are succinate dehydrogenase (SDH)-deficient GISTs with mutations in SDHA, SDHB, or SDHC[143,144], while others have no known genetic mutations. A recent report suggested that KRAS mutations might confer imatinib resistance in GIST, and although rare, KRAS gain-of-function mutations contribute to clinical imatinib resistance[145,146]. Serrano et al[145] used a Sequenom panel to screen for RAS, BRAF, and PI3KCA mutations in 27 wild-type GIST patients. Only one of these 27 GISTs contained a mutation in this pathway, harboring concomitant HRAS G12V and PIK3CA H1047R mutations[145]. KRAS and HRAS can contribute to GIST oncogenesis and indicate the importance of the PI3K/AKT and RAS/RAF pathways in GIST tumorigenesis.

As discussed previously, HSP90 inhibitors strongly inactive KIT kinase activity, but clinical applications in GIST patients have been prevented due to the toxicity resulting from inactivation of HSP90 client proteins beyond KIT and PDGFRA. Genome-scale short-hairpin RNA (shRNA) screening identified CDC37, an HSP90 cofactor, as an essential GIST-specific gene[147]. Validation studies in treatment-naive and imatinib-resistant GIST cell lines demonstrated that CDC37 is a viable therapeutic target in GIST, recapitulating the effect of HSP90 inhibition while remaining selective for KIT/PDGFRA and a limited number of other HSP90 clients[147]. CDC37 inhibition represents a potential HSP90 targeting strategy that limits toxicity for GIST patients.

The strongly expressed DOG1 (ANO1/TMEM16A) has been used as a diagnostic marker to differentiate GIST from other sarcomas[148-151]. Loss of DOG1 expression occurs together with loss of KIT expression in a subset of GISTs that are resistant to imatinib. Although DOG1 inhibition do not inhibit cell growth in vitro, DOG1 knockdown delays the growth of xenograft models of GIST and is associated with the upregulation of insulin-like growth factor binding protein 5, a potent antiangiogenic factor implicated in tumor suppression[152]. These findings suggest that DOG1 is a potential target in GIST through its role in IGFR signaling.

A recent analysis of the prognostic significance of aurora kinase A (AURKA) in imatinib-treated patients with advanced GIST suggested that the expression of AURKA may predict recurrence in patients with primary, surgically resected GISTs[153,154]. AURKA overexpression is a prognostic factor of poor PFS and OS. Inhibition of AURKA suppresses the growth of both imatinib-sensitive and imatinib-resistant GIST cells in a concentration-dependent manner, and it results in a synergistic cytotoxicity with imatinib[154].

CONCLUSION

Oncogenic KIT or PDGFRA receptor tyrosine kinase mutations are compelling therapeutic targets in GISTs, and the KIT/PDGFRA kinase inhibitors imatinib, sunitinib, and regorafenib are standards of care for patients with unresectable or metastatic GIST. However, most patients eventually develop resistance to KIT/PDGFRA kinase inhibitors, indicating that there is an urgent need to identify novel therapeutic strategies. A number of novel drugs are undergoing clinical trials, and several novel therapeutic targets have been identified, showing promise for the future treatment of GIST.

ACKNOWLEDGMENTS

We thank Mr. Grant Eilers for proof-reading the manuscript and providing crucial suggestions.

Footnotes

P- Reviewer: Sadoghi P S- Editor: Tian YL L- Editor: A E- Editor: Liu SQ

References
1.  Mazur MT, Clark HB. Gastric stromal tumors. Reappraisal of histogenesis. Am J Surg Pathol. 1983;7:507-519.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 581]  [Cited by in F6Publishing: 548]  [Article Influence: 13.4]  [Reference Citation Analysis (0)]
2.  Liegl-Atzwanger B, Fletcher JA, Fletcher CD. Gastrointestinal stromal tumors. Virchows Arch. 2010;456:111-127.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 158]  [Cited by in F6Publishing: 152]  [Article Influence: 10.9]  [Reference Citation Analysis (0)]
3.  Fletcher CD, Berman JJ, Corless C, Gorstein F, Lasota J, Longley BJ, Miettinen M, O’Leary TJ, Remotti H, Rubin BP. Diagnosis of gastrointestinal stromal tumors: A consensus approach. Hum Pathol. 2002;33:459-465.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2231]  [Cited by in F6Publishing: 2087]  [Article Influence: 94.9]  [Reference Citation Analysis (1)]
4.  Hirota S, Isozaki K, Moriyama Y, Hashimoto K, Nishida T, Ishiguro S, Kawano K, Hanada M, Kurata A, Takeda M. Gain-of-function mutations of c-kit in human gastrointestinal stromal tumors. Science. 1998;279:577-580.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3215]  [Cited by in F6Publishing: 3008]  [Article Influence: 115.7]  [Reference Citation Analysis (0)]
5.  Heinrich MC, Corless CL, Duensing A, McGreevey L, Chen CJ, Joseph N, Singer S, Griffith DJ, Haley A, Town A. PDGFRA activating mutations in gastrointestinal stromal tumors. Science. 2003;299:708-710.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1712]  [Cited by in F6Publishing: 1660]  [Article Influence: 79.0]  [Reference Citation Analysis (0)]
6.  Corless CL, Fletcher JA, Heinrich MC. Biology of gastrointestinal stromal tumors. J Clin Oncol. 2004;22:3813-3825.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 847]  [Cited by in F6Publishing: 809]  [Article Influence: 40.5]  [Reference Citation Analysis (0)]
7.  Tuveson DA, Willis NA, Jacks T, Griffin JD, Singer S, Fletcher CD, Fletcher JA, Demetri GD. STI571 inactivation of the gastrointestinal stromal tumor c-KIT oncoprotein: biological and clinical implications. Oncogene. 2001;20:5054-5058.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 522]  [Cited by in F6Publishing: 535]  [Article Influence: 23.3]  [Reference Citation Analysis (0)]
8.  Demetri GD, von Mehren M, Blanke CD, Van den Abbeele AD, Eisenberg B, Roberts PJ, Heinrich MC, Tuveson DA, Singer S, Janicek M. Efficacy and safety of imatinib mesylate in advanced gastrointestinal stromal tumors. N Engl J Med. 2002;347:472-480.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3203]  [Cited by in F6Publishing: 3013]  [Article Influence: 137.0]  [Reference Citation Analysis (0)]
9.  Heinrich MC, Maki RG, Corless CL, Antonescu CR, Harlow A, Griffith D, Town A, McKinley A, Ou WB, Fletcher JA. Primary and secondary kinase genotypes correlate with the biological and clinical activity of sunitinib in imatinib-resistant gastrointestinal stromal tumor. J Clin Oncol. 2008;26:5352-5359.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 533]  [Cited by in F6Publishing: 544]  [Article Influence: 34.0]  [Reference Citation Analysis (0)]
10.  Blay JY, Bonvalot S, Casali P, Choi H, Debiec-Richter M, Dei Tos AP, Emile JF, Gronchi A, Hogendoorn PC, Joensuu H. Consensus meeting for the management of gastrointestinal stromal tumors. Report of the GIST Consensus Conference of 20-21 March 2004, under the auspices of ESMO. Ann Oncol. 2005;16:566-578.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 500]  [Cited by in F6Publishing: 564]  [Article Influence: 29.7]  [Reference Citation Analysis (0)]
11.  Demetri GD, Heinrich MC, Fletcher JA, Fletcher CD, Van den Abbeele AD, Corless CL, Antonescu CR, George S, Morgan JA, Chen MH. Molecular target modulation, imaging, and clinical evaluation of gastrointestinal stromal tumor patients treated with sunitinib malate after imatinib failure. Clin Cancer Res. 2009;15:5902-5909.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 101]  [Cited by in F6Publishing: 111]  [Article Influence: 7.4]  [Reference Citation Analysis (0)]
12.  Demetri GD, Reichardt P, Kang YK, Blay JY, Rutkowski P, Gelderblom H, Hohenberger P, Leahy M, von Mehren M, Joensuu H. Efficacy and safety of regorafenib for advanced gastrointestinal stromal tumours after failure of imatinib and sunitinib (GRID): an international, multicentre, randomised, placebo-controlled, phase 3 trial. Lancet. 2013;381:295-302.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 942]  [Cited by in F6Publishing: 938]  [Article Influence: 85.3]  [Reference Citation Analysis (0)]
13.  Dematteo RP, Ballman KV, Antonescu CR, Maki RG, Pisters PW, Demetri GD, Blackstein ME, Blanke CD, von Mehren M, Brennan MF. Adjuvant imatinib mesylate after resection of localised, primary gastrointestinal stromal tumour: a randomised, double-blind, placebo-controlled trial. Lancet. 2009;373:1097-1104.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 986]  [Cited by in F6Publishing: 913]  [Article Influence: 60.9]  [Reference Citation Analysis (0)]
14.  Liegl B, Kepten I, Le C, Zhu M, Demetri GD, Heinrich MC, Fletcher CD, Corless CL, Fletcher JA. Heterogeneity of kinase inhibitor resistance mechanisms in GIST. J Pathol. 2008;216:64-74.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 313]  [Cited by in F6Publishing: 317]  [Article Influence: 19.8]  [Reference Citation Analysis (0)]
15.  Heinrich MC, Corless CL, Blanke CD, Demetri GD, Joensuu H, Roberts PJ, Eisenberg BL, von Mehren M, Fletcher CD, Sandau K. Molecular correlates of imatinib resistance in gastrointestinal stromal tumors. J Clin Oncol. 2006;24:4764-4774.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 596]  [Cited by in F6Publishing: 596]  [Article Influence: 33.1]  [Reference Citation Analysis (0)]
16.  Rubin BP, Singer S, Tsao C, Duensing A, Lux ML, Ruiz R, Hibbard MK, Chen CJ, Xiao S, Tuveson DA. KIT activation is a ubiquitous feature of gastrointestinal stromal tumors. Cancer Res. 2001;61:8118-8121.  [PubMed]  [DOI]  [Cited in This Article: ]
17.  Lasota J, Dansonka-Mieszkowska A, Stachura T, Schneider-Stock R, Kallajoki M, Steigen SE, Sarlomo-Rikala M, Boltze C, Kordek R, Roessner A. Gastrointestinal stromal tumors with internal tandem duplications in 3’ end of KIT juxtamembrane domain occur predominantly in stomach and generally seem to have a favorable course. Mod Pathol. 2003;16:1257-1264.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 87]  [Cited by in F6Publishing: 71]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
18.  Lasota J, Miettinen M. Clinical significance of oncogenic KIT and PDGFRA mutations in gastrointestinal stromal tumours. Histopathology. 2008;53:245-266.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 305]  [Cited by in F6Publishing: 331]  [Article Influence: 20.7]  [Reference Citation Analysis (1)]
19.  Antonescu CR, Sommer G, Sarran L, Tschernyavsky SJ, Riedel E, Woodruff JM, Robson M, Maki R, Brennan MF, Ladanyi M. Association of KIT exon 9 mutations with nongastric primary site and aggressive behavior: KIT mutation analysis and clinical correlates of 120 gastrointestinal stromal tumors. Clin Cancer Res. 2003;9:3329-3337.  [PubMed]  [DOI]  [Cited in This Article: ]
20.  Miettinen M, Makhlouf H, Sobin LH, Lasota J. Gastrointestinal stromal tumors of the jejunum and ileum: a clinicopathologic, immunohistochemical, and molecular genetic study of 906 cases before imatinib with long-term follow-up. Am J Surg Pathol. 2006;30:477-489.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 465]  [Cited by in F6Publishing: 513]  [Article Influence: 28.5]  [Reference Citation Analysis (0)]
21.  Heinrich MC, Owzar K, Corless CL, Hollis D, Borden EC, Fletcher CD, Ryan CW, von Mehren M, Blanke CD, Rankin C. Correlation of kinase genotype and clinical outcome in the North American Intergroup Phase III Trial of imatinib mesylate for treatment of advanced gastrointestinal stromal tumor: CALGB 150105 Study by Cancer and Leukemia Group B and Southwest Oncology Group. J Clin Oncol. 2008;26:5360-5367.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 441]  [Cited by in F6Publishing: 439]  [Article Influence: 27.4]  [Reference Citation Analysis (0)]
22.  Lasota J, Corless CL, Heinrich MC, Debiec-Rychter M, Sciot R, Wardelmann E, Merkelbach-Bruse S, Schildhaus HU, Steigen SE, Stachura J. Clinicopathologic profile of gastrointestinal stromal tumors (GISTs) with primary KIT exon 13 or exon 17 mutations: a multicenter study on 54 cases. Mod Pathol. 2008;21:476-484.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 120]  [Cited by in F6Publishing: 130]  [Article Influence: 8.1]  [Reference Citation Analysis (0)]
23.  Lasota J, Stachura J, Miettinen M. GISTs with PDGFRA exon 14 mutations represent subset of clinically favorable gastric tumors with epithelioid morphology. Lab Invest. 2006;86:94-100.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 103]  [Cited by in F6Publishing: 103]  [Article Influence: 5.7]  [Reference Citation Analysis (0)]
24.  Lasota J, Dansonka-Mieszkowska A, Sobin LH, Miettinen M. A great majority of GISTs with PDGFRA mutations represent gastric tumors of low or no malignant potential. Lab Invest. 2004;84:874-883.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 238]  [Cited by in F6Publishing: 217]  [Article Influence: 10.9]  [Reference Citation Analysis (0)]
25.  Debiec-Rychter M, Sciot R, Le Cesne A, Schlemmer M, Hohenberger P, van Oosterom AT, Blay JY, Leyvraz S, Stul M, Casali PG. KIT mutations and dose selection for imatinib in patients with advanced gastrointestinal stromal tumours. Eur J Cancer. 2006;42:1093-1103.  [PubMed]  [DOI]  [Cited in This Article: ]
26.  Corless CL, Schroeder A, Griffith D, Town A, McGreevey L, Harrell P, Shiraga S, Bainbridge T, Morich J, Heinrich MC. PDGFRA mutations in gastrointestinal stromal tumors: frequency, spectrum and in vitro sensitivity to imatinib. J Clin Oncol. 2005;23:5357-5364.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 579]  [Cited by in F6Publishing: 568]  [Article Influence: 29.9]  [Reference Citation Analysis (0)]
27.  Hirota S, Ohashi A, Nishida T, Isozaki K, Kinoshita K, Shinomura Y, Kitamura Y. Gain-of-function mutations of platelet-derived growth factor receptor alpha gene in gastrointestinal stromal tumors. Gastroenterology. 2003;125:660-667.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 492]  [Cited by in F6Publishing: 470]  [Article Influence: 22.4]  [Reference Citation Analysis (0)]
28.  Heinrich MC, Corless CL, Demetri GD, Blanke CD, von Mehren M, Joensuu H, McGreevey LS, Chen CJ, Van den Abbeele AD, Druker BJ. Kinase mutations and imatinib response in patients with metastatic gastrointestinal stromal tumor. J Clin Oncol. 2003;21:4342-4349.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1782]  [Cited by in F6Publishing: 1602]  [Article Influence: 76.3]  [Reference Citation Analysis (0)]
29.  Duensing A, Medeiros F, McConarty B, Joseph NE, Panigrahy D, Singer S, Fletcher CD, Demetri GD, Fletcher JA. Mechanisms of oncogenic KIT signal transduction in primary gastrointestinal stromal tumors (GISTs). Oncogene. 2004;23:3999-4006.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 245]  [Cited by in F6Publishing: 238]  [Article Influence: 11.9]  [Reference Citation Analysis (0)]
30.  Bauer S, Duensing A, Demetri GD, Fletcher JA. KIT oncogenic signaling mechanisms in imatinib-resistant gastrointestinal stromal tumor: PI3-kinase/AKT is a crucial survival pathway. Oncogene. 2007;26:7560-7568.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 193]  [Cited by in F6Publishing: 195]  [Article Influence: 11.5]  [Reference Citation Analysis (0)]
31.  Bachet JB, Tabone-Eglinger S, Dessaux S, Besse A, Brahimi-Adouane S, Emile JF, Blay JY, Alberti L. Gene expression patterns of hemizygous and heterozygous KIT mutations suggest distinct oncogenic pathways: a study in NIH3T3 cell lines and GIST samples. PLoS One. 2013;8:e61103.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 10]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
32.  Beadling C, Patterson J, Justusson E, Nelson D, Pantaleo MA, Hornick JL, Chacón M, Corless CL, Heinrich MC. Gene expression of the IGF pathway family distinguishes subsets of gastrointestinal stromal tumors wild type for KIT and PDGFRA. Cancer Med. 2013;2:21-31.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 26]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
33.  Le Cesne A, Blay JY, Reichardt P, Joensuu H. Optimizing tyrosine kinase inhibitor therapy in gastrointestinal stromal tumors: exploring the benefits of continuous kinase suppression. Oncologist. 2013;18:1192-1199.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 13]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
34.  Buchdunger E, Cioffi CL, Law N, Stover D, Ohno-Jones S, Druker BJ, Lydon NB. Abl protein-tyrosine kinase inhibitor STI571 inhibits in vitro signal transduction mediated by c-kit and platelet-derived growth factor receptors. J Pharmacol Exp Ther. 2000;295:139-145.  [PubMed]  [DOI]  [Cited in This Article: ]
35.  Heinrich MC, Griffith DJ, Druker BJ, Wait CL, Ott KA, Zigler AJ. Inhibition of c-kit receptor tyrosine kinase activity by STI 571, a selective tyrosine kinase inhibitor. Blood. 2000;96:925-932.  [PubMed]  [DOI]  [Cited in This Article: ]
36.  Mol CD, Dougan DR, Schneider TR, Skene RJ, Kraus ML, Scheibe DN, Snell GP, Zou H, Sang BC, Wilson KP. Structural basis for the autoinhibition and STI-571 inhibition of c-Kit tyrosine kinase. J Biol Chem. 2004;279:31655-31663.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 445]  [Cited by in F6Publishing: 484]  [Article Influence: 24.2]  [Reference Citation Analysis (0)]
37.  Yuzawa S, Opatowsky Y, Zhang Z, Mandiyan V, Lax I, Schlessinger J. Structural basis for activation of the receptor tyrosine kinase KIT by stem cell factor. Cell. 2007;130:323-334.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 236]  [Cited by in F6Publishing: 248]  [Article Influence: 14.6]  [Reference Citation Analysis (0)]
38.  Gramza AW, Corless CL, Heinrich MC. Resistance to Tyrosine Kinase Inhibitors in Gastrointestinal Stromal Tumors. Clin Cancer Res. 2009;15:7510-7518.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 156]  [Cited by in F6Publishing: 167]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
39.  Fletcher JA, Corless CL, Dimitrijevic S, von Mehren M, Eisenberg B, Joensuu H, Fletcher CDM, Blanke CD, Demetri GD, Heinrich MC. Mechanisms of resistance to imatinib mesylate (IM) in advanced gastrointestinal stromal tumor (GIST). Proc Am Soc Clin Oncol. 2003;22:815.  [PubMed]  [DOI]  [Cited in This Article: ]
40.  Debiec-Rychter M, Cools J, Dumez H, Sciot R, Stul M, Mentens N, Vranckx H, Wasag B, Prenen H, Roesel J. Mechanisms of resistance to imatinib mesylate in gastrointestinal stromal tumors and activity of the PKC412 inhibitor against imatinib-resistant mutants. Gastroenterology. 2005;128:270-279.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 375]  [Cited by in F6Publishing: 366]  [Article Influence: 19.3]  [Reference Citation Analysis (0)]
41.  Mahadevan D, Cooke L, Riley C, Swart R, Simons B, Della Croce K, Wisner L, Iorio M, Shakalya K, Garewal H. A novel tyrosine kinase switch is a mechanism of imatinib resistance in gastrointestinal stromal tumors. Oncogene. 2007;26:3909-3919.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 212]  [Cited by in F6Publishing: 219]  [Article Influence: 12.9]  [Reference Citation Analysis (0)]
42.  Schittenhelm MM, Shiraga S, Schroeder A, Corbin AS, Griffith D, Lee FY, Bokemeyer C, Deininger MW, Druker BJ, Heinrich MC. Dasatinib (BMS-354825), a dual SRC/ABL kinase inhibitor, inhibits the kinase activity of wild-type, juxtamembrane, and activation loop mutant KIT isoforms associated with human malignancies. Cancer Res. 2006;66:473-481.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 351]  [Cited by in F6Publishing: 353]  [Article Influence: 19.6]  [Reference Citation Analysis (0)]
43.  Rossi F, Yozgat Y, de Stanchina E, Veach D, Clarkson B, Manova K, Giancotti FG, Antonescu CR, Besmer P. Imatinib upregulates compensatory integrin signaling in a mouse model of gastrointestinal stromal tumor and is more effective when combined with dasatinib. Mol Cancer Res. 2010;8:1271-1283.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 29]  [Cited by in F6Publishing: 31]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
44.  Balachandran VP, Cavnar MJ, Zeng S, Bamboat ZM, Ocuin LM, Obaid H, Sorenson EC, Popow R, Ariyan C, Rossi F. Imatinib potentiates antitumor T cell responses in gastrointestinal stromal tumor through the inhibition of Ido. Nat Med. 2011;17:1094-1100.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 396]  [Cited by in F6Publishing: 408]  [Article Influence: 31.4]  [Reference Citation Analysis (0)]
45.  Li J, Gao J, Hong J, Shen L. Efficacy and safety of sunitinib in Chinese patients with imatinib-resistant or -intolerant gastrointestinal stromal tumors. Future Oncol. 2012;8:617-624.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 21]  [Cited by in F6Publishing: 23]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
46.  Li J, Gong JF, Li J, Gao J, Sun NP, Shen L. Efficacy of imatinib dose escalation in Chinese gastrointestinal stromal tumor patients. World J Gastroenterol. 2012;18:698-703.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 19]  [Cited by in F6Publishing: 21]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
47.  Gschwind A, Fischer OM, Ullrich A. The discovery of receptor tyrosine kinases: targets for cancer therapy. Nat Rev Cancer. 2004;4:361-370.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 874]  [Cited by in F6Publishing: 851]  [Article Influence: 42.6]  [Reference Citation Analysis (0)]
48.  Drevs J, Medinger M, Schmidt-Gersbach C, Weber R, Unger C. Receptor tyrosine kinases: the main targets for new anticancer therapy. Curr Drug Targets. 2003;4:113-121.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 61]  [Cited by in F6Publishing: 62]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
49.  Medinger M, Drevs J. Receptor tyrosine kinases and anticancer therapy. Curr Pharm Des. 2005;11:1139-1149.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 27]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
50.  Blume-Jensen P, Hunter T. Oncogenic kinase signalling. Nature. 2001;411:355-365.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2716]  [Cited by in F6Publishing: 2623]  [Article Influence: 114.0]  [Reference Citation Analysis (0)]
51.  Druker BJ, Talpaz M, Resta DJ, Peng B, Buchdunger E, Ford JM, Lydon NB, Kantarjian H, Capdeville R, Ohno-Jones S. Efficacy and safety of a specific inhibitor of the BCR-ABL tyrosine kinase in chronic myeloid leukemia. N Engl J Med. 2001;344:1031-1037.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3787]  [Cited by in F6Publishing: 3543]  [Article Influence: 154.0]  [Reference Citation Analysis (0)]
52.  Pegram MD, Lipton A, Hayes DF, Weber BL, Baselga JM, Tripathy D, Baly D, Baughman SA, Twaddell T, Glaspy JA. Phase II study of receptor-enhanced chemosensitivity using recombinant humanized anti-p185HER2/neu monoclonal antibody plus cisplatin in patients with HER2/neu-overexpressing metastatic breast cancer refractory to chemotherapy treatment. J Clin Oncol. 1998;16:2659-2671.  [PubMed]  [DOI]  [Cited in This Article: ]
53.  Yonesaka K, Zejnullahu K, Okamoto I, Satoh T, Cappuzzo F, Souglakos J, Ercan D, Rogers A, Roncalli M, Takeda M. Activation of ERBB2 signaling causes resistance to the EGFR-directed therapeutic antibody cetuximab. Sci Transl Med. 2011;3:99ra86.  [PubMed]  [DOI]  [Cited in This Article: ]
54.  Xu L, Kikuchi E, Xu C, Ebi H, Ercan D, Cheng KA, Padera R, Engelman JA, Jänne PA, Shapiro GI. Combined EGFR/MET or EGFR/HSP90 inhibition is effective in the treatment of lung cancers codriven by mutant EGFR containing T790M and MET. Cancer Res. 2012;72:3302-3311.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 75]  [Cited by in F6Publishing: 85]  [Article Influence: 7.1]  [Reference Citation Analysis (0)]
55.  Shaw AT, Kim DW, Nakagawa K, Seto T, Crinó L, Ahn MJ, De Pas T, Besse B, Solomon BJ, Blackhall F. Crizotinib versus chemotherapy in advanced ALK-positive lung cancer. N Engl J Med. 2013;368:2385-2394.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2534]  [Cited by in F6Publishing: 2584]  [Article Influence: 234.9]  [Reference Citation Analysis (0)]
56.  Reckamp KL, Giaccone G, Camidge DR, Gadgeel SM, Khuri FR, Engelman JA, Koczywas M, Rajan A, Campbell AK, Gernhardt D. A phase 2 trial of dacomitinib (PF-00299804), an oral, irreversible pan-HER (human epidermal growth factor receptor) inhibitor, in patients with advanced non-small cell lung cancer after failure of prior chemotherapy and erlotinib. Cancer. 2014;120:1145-1154.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 100]  [Cited by in F6Publishing: 108]  [Article Influence: 10.8]  [Reference Citation Analysis (0)]
57.  Ou SH, Jänne PA, Bartlett CH, Tang Y, Kim DW, Otterson GA, Crinò L, Selaru P, Cohen DP, Clark JW. Clinical benefit of continuing ALK inhibition with crizotinib beyond initial disease progression in patients with advanced ALK-positive NSCLC. Ann Oncol. 2014;25:415-422.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 147]  [Cited by in F6Publishing: 178]  [Article Influence: 19.8]  [Reference Citation Analysis (0)]
58.  Engelman JA, Jänne PA. Mechanisms of acquired resistance to epidermal growth factor receptor tyrosine kinase inhibitors in non-small cell lung cancer. Clin Cancer Res. 2008;14:2895-2899.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 513]  [Cited by in F6Publishing: 548]  [Article Influence: 34.3]  [Reference Citation Analysis (0)]
59.  Engelman JA, Zejnullahu K, Mitsudomi T, Song Y, Hyland C, Park JO, Lindeman N, Gale CM, Zhao X, Christensen J. MET amplification leads to gefitinib resistance in lung cancer by activating ERBB3 signaling. Science. 2007;316:1039-1043.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3442]  [Cited by in F6Publishing: 3549]  [Article Influence: 208.8]  [Reference Citation Analysis (0)]
60.  de Figueiredo-Pontes LL, Wong DW, Tin VP, Chung LP, Yasuda H, Yamaguchi N, Nakayama S, Jänne PA, Wong MP, Kobayashi SS. Identification and characterization of ALK kinase splicing isoforms in non-small-cell lung cancer. J Thorac Oncol. 2014;9:248-253.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 11]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
61.  Cappuzzo F, Jänne PA, Skokan M, Finocchiaro G, Rossi E, Ligorio C, Zucali PA, Terracciano L, Toschi L, Roncalli M. MET increased gene copy number and primary resistance to gefitinib therapy in non-small-cell lung cancer patients. Ann Oncol. 2009;20:298-304.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 219]  [Cited by in F6Publishing: 246]  [Article Influence: 15.4]  [Reference Citation Analysis (0)]
62.  Paez JG, Jänne PA, Lee JC, Tracy S, Greulich H, Gabriel S, Herman P, Kaye FJ, Lindeman N, Boggon TJ. EGFR mutations in lung cancer: correlation with clinical response to gefitinib therapy. Science. 2004;304:1497-1500.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 7278]  [Cited by in F6Publishing: 7326]  [Article Influence: 366.3]  [Reference Citation Analysis (0)]
63.  Faivre S, Delbaldo C, Vera K, Robert C, Lozahic S, Lassau N, Bello C, Deprimo S, Brega N, Massimini G. Safety, pharmacokinetic, and antitumor activity of SU11248, a novel oral multitarget tyrosine kinase inhibitor, in patients with cancer. J Clin Oncol. 2006;24:25-35.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 909]  [Cited by in F6Publishing: 887]  [Article Influence: 46.7]  [Reference Citation Analysis (0)]
64.  Demetri GD, George S, Heinrich MC, Fletcher JA, Fletcher CDM, Desai J, Cohen DP, Scigalla P, Cherrington JM, Van den Abbeele AD. Clinical activity and tolerability of the multi-targeted tyrosine kinase inhibitor SU11248 in patients (pts) with metastatic gastrointestinal stromal tumor (GIST) refractory to imatinib mesylate. Proc Am Soc Clin Oncol. 2003;22:814.  [PubMed]  [DOI]  [Cited in This Article: ]
65.  Demetri GD, van Oosterom AT, Garrett CR, Blackstein ME, Shah MH, Verweij J, McArthur G, Judson IR, Heinrich MC, Morgan JA. Efficacy and safety of sunitinib in patients with advanced gastrointestinal stromal tumour after failure of imatinib: a randomised controlled trial. Lancet. 2006;368:1329-1338.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1942]  [Cited by in F6Publishing: 1830]  [Article Influence: 101.7]  [Reference Citation Analysis (0)]
66.  George S, Wang Q, Heinrich MC, Corless CL, Zhu M, Butrynski JE, Morgan JA, Wagner AJ, Choy E, Tap WD. Efficacy and safety of regorafenib in patients with metastatic and/or unresectable GI stromal tumor after failure of imatinib and sunitinib: a multicenter phase II trial. J Clin Oncol. 2012;30:2401-2407.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 201]  [Cited by in F6Publishing: 203]  [Article Influence: 16.9]  [Reference Citation Analysis (0)]
67.  Demetri GD, Casali PG, Blay JY, von Mehren M, Morgan JA, Bertulli R, Ray-Coquard I, Cassier P, Davey M, Borghaei H. A phase I study of single-agent nilotinib or in combination with imatinib in patients with imatinib-resistant gastrointestinal stromal tumors. Clin Cancer Res. 2009;15:5910-5916.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 83]  [Cited by in F6Publishing: 93]  [Article Influence: 6.2]  [Reference Citation Analysis (0)]
68.  Bayraktar UD, Bayraktar S, Rocha-Lima CM. Molecular basis and management of gastrointestinal stromal tumors. World J Gastroenterol. 2010;16:2726-2734.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 17]  [Cited by in F6Publishing: 21]  [Article Influence: 1.5]  [Reference Citation Analysis (1)]
69.  Montemurro M, Gelderblom H, Bitz U, Schütte J, Blay JY, Joensuu H, Trent J, Bauer S, Rutkowski P, Duffaud F. Sorafenib as third- or fourth-line treatment of advanced gastrointestinal stromal tumour and pretreatment including both imatinib and sunitinib, and nilotinib: A retrospective analysis. Eur J Cancer. 2013;49:1027-1031.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 49]  [Cited by in F6Publishing: 35]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
70.  Montemurro M, Bauer S. Treatment of gastrointestinal stromal tumor after imatinib and sunitinib. Curr Opin Oncol. 2011;23:367-372.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 15]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
71.  Montemurro M, Schöffski P, Reichardt P, Gelderblom H, Schütte J, Hartmann JT, von Moos R, Seddon B, Joensuu H, Wendtner CM. Nilotinib in the treatment of advanced gastrointestinal stromal tumours resistant to both imatinib and sunitinib. Eur J Cancer. 2009;45:2293-2297.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 109]  [Cited by in F6Publishing: 125]  [Article Influence: 8.3]  [Reference Citation Analysis (0)]
72.  Wagner AJ, Chugh R, Rosen LS, Morgan JA, George S, Gordon M, Dunbar J, Normant E, Grayzel D, Demetri GD. A phase I study of the HSP90 inhibitor retaspimycin hydrochloride (IPI-504) in patients with gastrointestinal stromal tumors or soft-tissue sarcomas. Clin Cancer Res. 2013;19:6020-6029.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 57]  [Cited by in F6Publishing: 61]  [Article Influence: 5.5]  [Reference Citation Analysis (0)]
73.  Schöffski P, Reichardt P, Blay JY, Dumez H, Morgan JA, Ray-Coquard I, Hollaender N, Jappe A, Demetri GD. A phase I-II study of everolimus (RAD001) in combination with imatinib in patients with imatinib-resistant gastrointestinal stromal tumors. Ann Oncol. 2010;21:1990-1998.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 93]  [Cited by in F6Publishing: 102]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
74.  Reichardt P, Blay JY, Gelderblom H, Schlemmer M, Demetri GD, Bui-Nguyen B, McArthur GA, Yazji S, Hsu Y, Galetic I. Phase III study of nilotinib versus best supportive care with or without a TKI in patients with gastrointestinal stromal tumors resistant to or intolerant of imatinib and sunitinib. Ann Oncol. 2012;23:1680-1687.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 98]  [Cited by in F6Publishing: 110]  [Article Influence: 9.2]  [Reference Citation Analysis (0)]
75.  Quek R, Wang Q, Morgan JA, Shapiro GI, Butrynski JE, Ramaiya N, Huftalen T, Jederlinic N, Manola J, Wagner AJ. Combination mTOR and IGF-1R inhibition: phase I trial of everolimus and figitumumab in patients with advanced sarcomas and other solid tumors. Clin Cancer Res. 2011;17:871-879.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 117]  [Cited by in F6Publishing: 137]  [Article Influence: 9.8]  [Reference Citation Analysis (0)]
76.  Norden-Zfoni A, Desai J, Manola J, Beaudry P, Force J, Maki R, Folkman J, Bello C, Baum C, DePrimo SE. Blood-based biomarkers of SU11248 activity and clinical outcome in patients with metastatic imatinib-resistant gastrointestinal stromal tumor. Clin Cancer Res. 2007;13:2643-2650.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 171]  [Cited by in F6Publishing: 177]  [Article Influence: 10.4]  [Reference Citation Analysis (0)]
77.  Poole CD, Connolly MP, Chang J, Currie CJ. Health utility of patients with advanced gastrointestinal stromal tumors (GIST) after failure of imatinib and sunitinib: findings from GRID, a randomized, double-blind, placebo-controlled phase III study of regorafenib versus placebo. Gastric Cancer. 2014;Epub ahead of print.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 12]  [Cited by in F6Publishing: 14]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
78.  Demetri GD. mTOR inhibitors in sarcoma. Clin Adv Hematol Oncol. 2011;9:145-147.  [PubMed]  [DOI]  [Cited in This Article: ]
79.  Bauer S, Yu LK, Demetri GD, Fletcher JA. Heat shock protein 90 inhibition in imatinib-resistant gastrointestinal stromal tumor. Cancer Res. 2006;66:9153-9161.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 204]  [Cited by in F6Publishing: 209]  [Article Influence: 11.6]  [Reference Citation Analysis (0)]
80.  Guo T, Agaram NP, Wong GC, Hom G, D’Adamo D, Maki RG, Schwartz GK, Veach D, Clarkson BD, Singer S. Sorafenib inhibits the imatinib-resistant KITT670I gatekeeper mutation in gastrointestinal stromal tumor. Clin Cancer Res. 2007;13:4874-4881.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 115]  [Cited by in F6Publishing: 124]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
81.  Dickson MA, Okuno SH, Keohan ML, Maki RG, D’Adamo DR, Akhurst TJ, Antonescu CR, Schwartz GK. Phase II study of the HSP90-inhibitor BIIB021 in gastrointestinal stromal tumors. Ann Oncol. 2013;24:252-257.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 79]  [Cited by in F6Publishing: 77]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
82.  Benjamin RS, Schöffski P, Hartmann JT, Van Oosterom A, Bui BN, Duyster J, Schuetze S, Blay JY, Reichardt P, Rosen LS. Efficacy and safety of motesanib, an oral inhibitor of VEGF, PDGF, and Kit receptors, in patients with imatinib-resistant gastrointestinal stromal tumors. Cancer Chemother Pharmacol. 2011;68:69-77.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 41]  [Cited by in F6Publishing: 42]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
83.  von Mehren M. Beyond imatinib: second generation c-KIT inhibitors for the management of gastrointestinal stromal tumors. Clin Colorectal Cancer. 2006;6 Suppl 1:S30-S34.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 29]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
84.  Overton LC, Heinrich MC. Regorafenib for treatment of advanced gastrointestinal stromal tumors. Expert Opin Pharmacother. 2014;15:549-558.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 12]  [Cited by in F6Publishing: 14]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
85.  Blay JY, von Mehren M. Nilotinib: a novel, selective tyrosine kinase inhibitor. Semin Oncol. 2011;38 Suppl 1:S3-S9.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 87]  [Cited by in F6Publishing: 89]  [Article Influence: 6.8]  [Reference Citation Analysis (0)]
86.  von Mehren M, Rankin C, Goldblum JR, Demetri GD, Bramwell V, Ryan CW, Borden E. Phase 2 Southwest Oncology Group-directed intergroup trial (S0505) of sorafenib in advanced soft tissue sarcomas. Cancer. 2012;118:770-776.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 84]  [Cited by in F6Publishing: 87]  [Article Influence: 6.7]  [Reference Citation Analysis (0)]
87.  Heinrich MC, Marino-Enriquez A, Presnell A, Donsky RS, Griffith DJ, McKinley A, Patterson J, Taguchi T, Liang CW, Fletcher JA. Sorafenib inhibits many kinase mutations associated with drug-resistant gastrointestinal stromal tumors. Mol Cancer Ther. 2012;11:1770-1780.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 55]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
88.  Adachi Y, Yamamoto H, Ohashi H, Endo T, Carbone DP, Imai K, Shinomura Y. A candidate targeting molecule of insulin-like growth factor-I receptor for gastrointestinal cancers. World J Gastroenterol. 2010;16:5779-5789.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 23]  [Cited by in F6Publishing: 27]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
89.  Joensuu H, De Braud F, Grignagni G, De Pas T, Spitalieri G, Coco P, Spreafico C, Boselli S, Toffalorio F, Bono P. Vatalanib for metastatic gastrointestinal stromal tumour (GIST) resistant to imatinib: final results of a phase II study. Br J Cancer. 2011;104:1686-1690.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 54]  [Cited by in F6Publishing: 56]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
90.  Joensuu H, De Braud F, Coco P, De Pas T, Putzu C, Spreafico C, Bono P, Bosselli S, Jalava T, Laurent D. Phase II, open-label study of PTK787/ZK222584 for the treatment of metastatic gastrointestinal stromal tumors resistant to imatinib mesylate. Ann Oncol. 2008;19:173-177.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 54]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
91.  Kim EJ, Zalupski MM. Systemic therapy for advanced gastrointestinal stromal tumors: beyond imatinib. J Surg Oncol. 2011;104:901-906.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 21]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
92.  Tarn C, Skorobogatko YV, Taguchi T, Eisenberg B, von Mehren M, Godwin AK. Therapeutic effect of imatinib in gastrointestinal stromal tumors: AKT signaling dependent and independent mechanisms. Cancer Res. 2006;66:5477-5486.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 58]  [Cited by in F6Publishing: 60]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
93.  Pantaleo MA, Nicoletti G, Nanni C, Gnocchi C, Landuzzi L, Quarta C, Boschi S, Nannini M, Di Battista M, Castellucci P. Preclinical evaluation of KIT/PDGFRA and mTOR inhibitors in gastrointestinal stromal tumors using small animal FDG PET. J Exp Clin Cancer Res. 2010;29:173.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 23]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
94.  Patel S. Exploring novel therapeutic targets in GIST: focus on the PI3K/Akt/mTOR pathway. Curr Oncol Rep. 2013;15:386-395.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 45]  [Cited by in F6Publishing: 48]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
95.  McCubrey JA, Steelman LS, Chappell WH, Abrams SL, Franklin RA, Montalto G, Cervello M, Libra M, Candido S, Malaponte G. Ras/Raf/MEK/ERK and PI3K/PTEN/Akt/mTOR cascade inhibitors: how mutations can result in therapy resistance and how to overcome resistance. Oncotarget. 2012;3:1068-1111.  [PubMed]  [DOI]  [Cited in This Article: ]
96.  Demetri GD, Chawla SP, Ray-Coquard I, Le Cesne A, Staddon AP, Milhem MM, Penel N, Riedel RF, Bui-Nguyen B, Cranmer LD. Results of an international randomized phase III trial of the mammalian target of rapamycin inhibitor ridaforolimus versus placebo to control metastatic sarcomas in patients after benefit from prior chemotherapy. J Clin Oncol. 2013;31:2485-2492.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 175]  [Cited by in F6Publishing: 179]  [Article Influence: 16.3]  [Reference Citation Analysis (0)]
97.  Floris G, Wozniak A, Sciot R, Li H, Friedman L, Van Looy T, Wellens J, Vermaelen P, Deroose CM, Fletcher JA. A potent combination of the novel PI3K Inhibitor, GDC-0941, with imatinib in gastrointestinal stromal tumor xenografts: long-lasting responses after treatment withdrawal. Clin Cancer Res. 2013;19:620-630.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 55]  [Cited by in F6Publishing: 59]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
98.  Workman P. Combinatorial attack on multistep oncogenesis by inhibiting the Hsp90 molecular chaperone. Cancer Lett. 2004;206:149-157.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 203]  [Cited by in F6Publishing: 193]  [Article Influence: 9.7]  [Reference Citation Analysis (0)]
99.  Sreedhar AS, Soti C, Csermely P. Inhibition of Hsp90: a new strategy for inhibiting protein kinases. Biochim Biophys Acta. 2004;1697:233-242.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 109]  [Cited by in F6Publishing: 95]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
100.  Sreedhar AS, Kalmár E, Csermely P, Shen YF. Hsp90 isoforms: functions, expression and clinical importance. FEBS Lett. 2004;562:11-15.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 438]  [Cited by in F6Publishing: 420]  [Article Influence: 21.0]  [Reference Citation Analysis (0)]
101.  Sharp S, Workman P. Inhibitors of the HSP90 molecular chaperone: current status. Adv Cancer Res. 2006;95:323-348.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 256]  [Cited by in F6Publishing: 264]  [Article Influence: 14.7]  [Reference Citation Analysis (0)]
102.  Isaacs JS, Xu W, Neckers L. Heat shock protein 90 as a molecular target for cancer therapeutics. Cancer Cell. 2003;3:213-217.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 464]  [Cited by in F6Publishing: 477]  [Article Influence: 22.7]  [Reference Citation Analysis (0)]
103.  Waza M, Adachi H, Katsuno M, Minamiyama M, Sang C, Tanaka F, Inukai A, Doyu M, Sobue G. 17-AAG, an Hsp90 inhibitor, ameliorates polyglutamine-mediated motor neuron degeneration. Nat Med. 2005;11:1088-1095.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 284]  [Cited by in F6Publishing: 302]  [Article Influence: 15.9]  [Reference Citation Analysis (0)]
104.  Csermely P, Schnaider T, Soti C, Prohászka Z, Nardai G. The 90-kDa molecular chaperone family: structure, function, and clinical applications. A comprehensive review. Pharmacol Ther. 1998;79:129-168.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 784]  [Cited by in F6Publishing: 735]  [Article Influence: 28.3]  [Reference Citation Analysis (0)]
105.  Sydor JR, Normant E, Pien CS, Porter JR, Ge J, Grenier L, Pak RH, Ali JA, Dembski MS, Hudak J. Development of 17-allylamino-17-demethoxygeldanamycin hydroquinone hydrochloride (IPI-504), an anti-cancer agent directed against Hsp90. Proc Natl Acad Sci USA. 2006;103:17408-17413.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 175]  [Cited by in F6Publishing: 159]  [Article Influence: 8.8]  [Reference Citation Analysis (0)]
106.  Fumo G, Akin C, Metcalfe DD, Neckers L. 17-Allylamino-17-demethoxygeldanamycin (17-AAG) is effective in down-regulating mutated, constitutively activated KIT protein in human mast cells. Blood. 2004;103:1078-1084.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 116]  [Cited by in F6Publishing: 123]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
107.  Smyth T, Van Looy T, Curry JE, Rodriguez-Lopez AM, Wozniak A, Zhu M, Donsky R, Morgan JG, Mayeda M, Fletcher JA. The HSP90 inhibitor, AT13387, is effective against imatinib-sensitive and -resistant gastrointestinal stromal tumor models. Mol Cancer Ther. 2012;11:1799-1808.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 46]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
108.  Floris G, Debiec-Rychter M, Wozniak A, Stefan C, Normant E, Faa G, Machiels K, Vanleeuw U, Sciot R, Schöffski P. The heat shock protein 90 inhibitor IPI-504 induces KIT degradation, tumor shrinkage, and cell proliferation arrest in xenograft models of gastrointestinal stromal tumors. Mol Cancer Ther. 2011;10:1897-1908.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 34]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
109.  Dewaele B, Wasag B, Cools J, Sciot R, Prenen H, Vandenberghe P, Wozniak A, Schöffski P, Marynen P, Debiec-Rychter M. Activity of dasatinib, a dual SRC/ABL kinase inhibitor, and IPI-504, a heat shock protein 90 inhibitor, against gastrointestinal stromal tumor-associated PDGFRAD842V mutation. Clin Cancer Res. 2008;14:5749-5758.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 97]  [Cited by in F6Publishing: 105]  [Article Influence: 6.6]  [Reference Citation Analysis (0)]
110.  Reichardt P. Novel approaches to imatinib- and sunitinib-resistant GIST. Curr Oncol Rep. 2008;10:344-349.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 24]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
111.  Mühlenberg T, Zhang Y, Wagner AJ, Grabellus F, Bradner J, Taeger G, Lang H, Taguchi T, Schuler M, Fletcher JA. Inhibitors of deacetylases suppress oncogenic KIT signaling, acetylate HSP90, and induce apoptosis in gastrointestinal stromal tumors. Cancer Res. 2009;69:6941-6950.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 67]  [Cited by in F6Publishing: 76]  [Article Influence: 5.1]  [Reference Citation Analysis (0)]
112.  Chi P, Chen Y, Zhang L, Guo X, Wongvipat J, Shamu T, Fletcher JA, Dewell S, Maki RG, Zheng D. ETV1 is a lineage survival factor that cooperates with KIT in gastrointestinal stromal tumours. Nature. 2010;467:849-853.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 218]  [Cited by in F6Publishing: 241]  [Article Influence: 17.2]  [Reference Citation Analysis (0)]
113.  Kubota D, Yoshida A, Tsuda H, Suehara Y, Okubo T, Saito T, Orita H, Sato K, Taguchi T, Yao T. Gene expression network analysis of ETV1 reveals KCTD10 as a novel prognostic biomarker in gastrointestinal stromal tumor. PLoS One. 2013;8:e73896.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 21]  [Cited by in F6Publishing: 22]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
114.  Birner P, Beer A, Vinatzer U, Stary S, Höftberger R, Nirtl N, Wrba F, Streubel B, Schoppmann SF. MAPKAP kinase 2 overexpression influences prognosis in gastrointestinal stromal tumors and associates with copy number variations on chromosome 1 and expression of p38 MAP kinase and ETV1. Clin Cancer Res. 2012;18:1879-1887.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 29]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
115.  Wimmel A, Glitz D, Kraus A, Roeder J, Schuermann M. Axl receptor tyrosine kinase expression in human lung cancer cell lines correlates with cellular adhesion. Eur J Cancer. 2001;37:2264-2274.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 91]  [Cited by in F6Publishing: 89]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
116.  Varnum BC, Young C, Elliott G, Garcia A, Bartley TD, Fridell YW, Hunt RW, Trail G, Clogston C, Toso RJ. Axl receptor tyrosine kinase stimulated by the vitamin K-dependent protein encoded by growth-arrest-specific gene 6. Nature. 1995;373:623-626.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 341]  [Cited by in F6Publishing: 348]  [Article Influence: 12.0]  [Reference Citation Analysis (0)]
117.  Stitt TN, Conn G, Gore M, Lai C, Bruno J, Radziejewski C, Mattsson K, Fisher J, Gies DR, Jones PF. The anticoagulation factor protein S and its relative, Gas6, are ligands for the Tyro 3/Axl family of receptor tyrosine kinases. Cell. 1995;80:661-670.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 532]  [Cited by in F6Publishing: 550]  [Article Influence: 19.0]  [Reference Citation Analysis (0)]
118.  Jacob AN, Kalapurakal J, Davidson WR, Kandpal G, Dunson N, Prashar Y, Kandpal RP. A receptor tyrosine kinase, UFO/Axl, and other genes isolated by a modified differential display PCR are overexpressed in metastatic prostatic carcinoma cell line DU145. Cancer Detect Prev. 1999;23:325-332.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 37]  [Cited by in F6Publishing: 38]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
119.  Holland SJ, Powell MJ, Franci C, Chan EW, Friera AM, Atchison RE, McLaughlin J, Swift SE, Pali ES, Yam G. Multiple roles for the receptor tyrosine kinase axl in tumor formation. Cancer Res. 2005;65:9294-9303.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 146]  [Cited by in F6Publishing: 153]  [Article Influence: 8.1]  [Reference Citation Analysis (0)]
120.  Ou WB, Fletcher CDM, Demetri GD, Fletcher JA.  Activated Tyrosine Kinases in Gastrointestinal Stromal Tumor with Loss of KIT Oncoprotein Expression: 100th AACR Annual Meeting; 2009, April 18-22. Denver, CO: AACR Meeting Proceedings 2009; 324.  [PubMed]  [DOI]  [Cited in This Article: ]
121.  Liu F, Bardhan K, Yang D, Thangaraju M, Ganapathy V, Waller JL, Liles GB, Lee JR, Liu K. NF-κB directly regulates Fas transcription to modulate Fas-mediated apoptosis and tumor suppression. J Biol Chem. 2012;287:25530-25540.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 99]  [Cited by in F6Publishing: 102]  [Article Influence: 8.5]  [Reference Citation Analysis (0)]
122.  Li JH, Rosen D, Sondel P, Berke G. Immune privilege and FasL: two ways to inactivate effector cytotoxic T lymphocytes by FasL-expressing cells. Immunology. 2002;105:267-277.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 32]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
123.  Rikhof B, van der Graaf WT, Meijer C, Le PT, Meersma GJ, de Jong S, Fletcher JA, Suurmeijer AJ. Abundant Fas expression by gastrointestinal stromal tumours may serve as a therapeutic target for MegaFasL. Br J Cancer. 2008;99:1600-1606.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 10]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
124.  Xu Q, Wu Z. The insulin-like growth factor-phosphatidylinositol 3-kinase-Akt signaling pathway regulates myogenin expression in normal myogenic cells but not in rhabdomyosarcoma-derived RD cells. J Biol Chem. 2000;275:36750-36757.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 139]  [Cited by in F6Publishing: 145]  [Article Influence: 6.0]  [Reference Citation Analysis (0)]
125.  Sekyi-Otu A, Bell RS, Ohashi C, Pollak M, Andrulis IL. Insulin-like growth factor 1 (IGF-1) receptors, IGF-1, and IGF-2 are expressed in primary human sarcomas. Cancer Res. 1995;55:129-134.  [PubMed]  [DOI]  [Cited in This Article: ]
126.  Trent JC, Ramdas L, Dupart J, Hunt K, Macapinlac H, Taylor E, Hu L, Salvado A, Abbruzzese JL, Pollock R. Early effects of imatinib mesylate on the expression of insulin-like growth factor binding protein-3 and positron emission tomography in patients with gastrointestinal stromal tumor. Cancer. 2006;107:1898-1908.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 30]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
127.  Tarn C, Rink L, Merkel E, Flieder D, Pathak H, Koumbi D, Testa JR, Eisenberg B, von Mehren M, Godwin AK. Insulin-like growth factor 1 receptor is a potential therapeutic target for gastrointestinal stromal tumors. Proc Natl Acad Sci USA. 2008;105:8387-8392.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 190]  [Cited by in F6Publishing: 204]  [Article Influence: 12.8]  [Reference Citation Analysis (0)]
128.  Pantaleo MA, Astolfi A, Di Battista M, Heinrich MC, Paterini P, Scotlandi K, Santini D, Catena F, Manara MC, Nannini M. Insulin-like growth factor 1 receptor expression in wild-type GISTs: a potential novel therapeutic target. Int J Cancer. 2009;125:2991-2994.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 57]  [Cited by in F6Publishing: 65]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
129.  Janeway KA, Zhu MJ, Barretina J, Perez-Atayde A, Demetri GD, Fletcher JA. Strong expression of IGF1R in pediatric gastrointestinal stromal tumors without IGF1R genomic amplification. Int J Cancer. 2010;127:2718-2722.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 54]  [Cited by in F6Publishing: 58]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
130.  Rikhof B, van der Graaf WT, Suurmeijer AJ, van Doorn J, Meersma GJ, Groenen PJ, Schuuring EM, Meijer C, de Jong S. ‘Big’-insulin-like growth factor-II signaling is an autocrine survival pathway in gastrointestinal stromal tumors. Am J Pathol. 2012;181:303-312.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 14]  [Cited by in F6Publishing: 14]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
131.  Rikhof B, van Doorn J, Suurmeijer AJ, Rautenberg MW, Groenen PJ, Verdijk MA, Jager PL, de Jong S, Gietema JA, van der Graaf WT. Insulin-like growth factors and insulin-like growth factor-binding proteins in relation to disease status and incidence of hypoglycaemia in patients with a gastrointestinal stromal tumour. Ann Oncol. 2009;20:1582-1588.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 18]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
132.  Braconi C, Bracci R, Cellerino R. Molecular targets in Gastrointestinal Stromal Tumors (GIST) therapy. Curr Cancer Drug Targets. 2008;8:359-366.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 24]  [Cited by in F6Publishing: 28]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
133.  Braconi C, Bracci R, Bearzi I, Bianchi F, Sabato S, Mandolesi A, Belvederesi L, Cascinu S, Valeri N, Cellerino R. Insulin-like growth factor (IGF) 1 and 2 help to predict disease outcome in GIST patients. Ann Oncol. 2008;19:1293-1298.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 63]  [Cited by in F6Publishing: 69]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
134.  Belinsky MG, Rink L, Flieder DB, Jahromi MS, Schiffman JD, Godwin AK, Mehren Mv. Overexpression of insulin-like growth factor 1 receptor and frequent mutational inactivation of SDHA in wild-type SDHB-negative gastrointestinal stromal tumors. Genes Chromosomes Cancer. 2013;52:214-224.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 54]  [Cited by in F6Publishing: 60]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
135.  Belinsky MG, Rink L, Cai KQ, Ochs MF, Eisenberg B, Huang M, von Mehren M, Godwin AK. The insulin-like growth factor system as a potential therapeutic target in gastrointestinal stromal tumors. Cell Cycle. 2008;7:2949-2955.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 30]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
136.  Baier G, Telford D, Giampa L, Coggeshall KM, Baier-Bitterlich G, Isakov N, Altman A. Molecular cloning and characterization of PKC theta, a novel member of the protein kinase C (PKC) gene family expressed predominantly in hematopoietic cells. J Biol Chem. 1993;268:4997-5004.  [PubMed]  [DOI]  [Cited in This Article: ]
137.  Chang JD, Xu Y, Raychowdhury MK, Ware JA. Molecular cloning and expression of a cDNA encoding a novel isoenzyme of protein kinase C (nPKC). A new member of the nPKC family expressed in skeletal muscle, megakaryoblastic cells, and platelets. J Biol Chem. 1993;268:14208-14214.  [PubMed]  [DOI]  [Cited in This Article: ]
138.  Allander SV, Nupponen NN, Ringnér M, Hostetter G, Maher GW, Goldberger N, Chen Y, Carpten J, Elkahloun AG, Meltzer PS. Gastrointestinal stromal tumors with KIT mutations exhibit a remarkably homogeneous gene expression profile. Cancer Res. 2001;61:8624-8628.  [PubMed]  [DOI]  [Cited in This Article: ]
139.  Medeiros F, Corless CL, Duensing A, Hornick JL, Oliveira AM, Heinrich MC, Fletcher JA, Fletcher CD. KIT-negative gastrointestinal stromal tumors: proof of concept and therapeutic implications. Am J Surg Pathol. 2004;28:889-894.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 368]  [Cited by in F6Publishing: 321]  [Article Influence: 16.1]  [Reference Citation Analysis (0)]
140.  Nielsen TO, West RB, Linn SC, Alter O, Knowling MA, O’Connell JX, Zhu S, Fero M, Sherlock G, Pollack JR. Molecular characterisation of soft tissue tumours: a gene expression study. Lancet. 2002;359:1301-1307.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 443]  [Cited by in F6Publishing: 400]  [Article Influence: 18.2]  [Reference Citation Analysis (0)]
141.  Duensing A, Joseph NE, Medeiros F, Smith F, Hornick JL, Heinrich MC, Corless CL, Demetri GD, Fletcher CD, Fletcher JA. Protein Kinase C theta (PKCtheta) expression and constitutive activation in gastrointestinal stromal tumors (GISTs). Cancer Res. 2004;64:5127-5131.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 94]  [Cited by in F6Publishing: 100]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
142.  Ou WB, Zhu MJ, Demetri GD, Fletcher CD, Fletcher JA. Protein kinase C-theta regulates KIT expression and proliferation in gastrointestinal stromal tumors. Oncogene. 2008;27:5624-5634.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 59]  [Article Influence: 3.7]  [Reference Citation Analysis (0)]
143.  Belinsky MG, Rink L, von Mehren M. Succinate dehydrogenase deficiency in pediatric and adult gastrointestinal stromal tumors. Front Oncol. 2013;3:117.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 40]  [Cited by in F6Publishing: 40]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
144.  Janeway KA, Kim SY, Lodish M, Nosé V, Rustin P, Gaal J, Dahia PL, Liegl B, Ball ER, Raygada M. Defects in succinate dehydrogenase in gastrointestinal stromal tumors lacking KIT and PDGFRA mutations. Proc Natl Acad Sci USA. 2011;108:314-318.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 436]  [Cited by in F6Publishing: 433]  [Article Influence: 30.9]  [Reference Citation Analysis (0)]
145.  Serrano C, Wang Y, Mariño-Enríquez A, Lee JC, Ravegnini G, Morgan JA, Bertagnolli MM, Beadling C, Demetri GD, Corless CL. KRAS and KIT Gatekeeper Mutations Confer Polyclonal Primary Imatinib Resistance in GI Stromal Tumors: Relevance of Concomitant Phosphatidylinositol 3-Kinase/AKT Dysregulation. J Clin Oncol. 2014;Epub ahead of print.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 38]  [Cited by in F6Publishing: 44]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
146.  Miranda C, Nucifora M, Molinari F, Conca E, Anania MC, Bordoni A, Saletti P, Mazzucchelli L, Pilotti S, Pierotti MA. KRAS and BRAF mutations predict primary resistance to imatinib in gastrointestinal stromal tumors. Clin Cancer Res. 2012;18:1769-1776.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 124]  [Cited by in F6Publishing: 133]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
147.  Mariño-Enríquez A, Ou WB, Cowley G, Luo B, Jonker AH, Mayeda M, Okamoto M, Eilers G, Czaplinski JT, Sicinska E. Genome-wide functional screening identifies CDC37 as a crucial HSP90-cofactor for KIT oncogenic expression in gastrointestinal stromal tumors. Oncogene. 2014;33:1872-1876.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 23]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
148.  Wada T, Tanabe S, Ishido K, Higuchi K, Sasaki T, Katada C, Azuma M, Naruke A, Kim M, Koizumi W. DOG1 is useful for diagnosis of KIT-negative gastrointestinal stromal tumor of stomach. World J Gastroenterol. 2013;19:9133-9136.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 10]  [Cited by in F6Publishing: 12]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
149.  West RB, Corless CL, Chen X, Rubin BP, Subramanian S, Montgomery K, Zhu S, Ball CA, Nielsen TO, Patel R. The novel marker, DOG1, is expressed ubiquitously in gastrointestinal stromal tumors irrespective of KIT or PDGFRA mutation status. Am J Pathol. 2004;165:107-113.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 462]  [Cited by in F6Publishing: 440]  [Article Influence: 22.0]  [Reference Citation Analysis (0)]
150.  Liegl B, Hornick JL, Corless CL, Fletcher CD. Monoclonal antibody DOG1.1 shows higher sensitivity than KIT in the diagnosis of gastrointestinal stromal tumors, including unusual subtypes. Am J Surg Pathol. 2009;33:437-446.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 203]  [Cited by in F6Publishing: 184]  [Article Influence: 12.3]  [Reference Citation Analysis (0)]
151.  Espinosa I, Lee CH, Kim MK, Rouse BT, Subramanian S, Montgomery K, Varma S, Corless CL, Heinrich MC, Smith KS. A novel monoclonal antibody against DOG1 is a sensitive and specific marker for gastrointestinal stromal tumors. Am J Surg Pathol. 2008;32:210-218.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 319]  [Cited by in F6Publishing: 350]  [Article Influence: 21.9]  [Reference Citation Analysis (0)]
152.  Simon S, Grabellus F, Ferrera L, Galietta L, Schwindenhammer B, Mühlenberg T, Taeger G, Eilers G, Treckmann J, Breitenbuecher F. DOG1 regulates growth and IGFBP5 in gastrointestinal stromal tumors. Cancer Res. 2013;73:3661-3670.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 55]  [Cited by in F6Publishing: 58]  [Article Influence: 5.3]  [Reference Citation Analysis (0)]
153.  Yen CC, Yeh CN, Cheng CT, Jung SM, Huang SC, Chang TW, Jan YY, Tzeng CH, Chao TC, Chen YY. Integrating bioinformatics and clinicopathological research of gastrointestinal stromal tumors: identification of aurora kinase A as a poor risk marker. Ann Surg Oncol. 2012;19:3491-3499.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 21]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
154.  Yeh CN, Yen CC, Chen YY, Cheng CT, Huang SC, Chang TW, Yao FY, Lin YC, Wen YS, Chiang KC. Identification of aurora kinase A as an unfavorable prognostic factor and potential treatment target for metastatic gastrointestinal stromal tumors. Oncotarget. 2014;5:4071-4086.  [PubMed]  [DOI]  [Cited in This Article: ]