somatic alterations as the basis for resistance to targeted therapies

11
Journal of Pathology J Pathol 2014; 232: 244–254 Published online in Wiley Online Library (wileyonlinelibrary.com) DOI: 10.1002/path.4278 INVITED REVIEW Somatic alterations as the basis for resistance to targeted therapies Brian G Blair, 1 Alberto Bardelli 2 and Ben Ho Park 1 * 1 Sidney Kimmel Comprehensive Cancer Center at Johns Hopkins, Baltimore, MD, USA 2 Institute for Cancer Research and Treatment (IRCC), University of Torino Medical School, Candiolo, Torino, Italy *Correspondence to: BH Park, Sidney Kimmel Comprehensive Cancer Center at Johns Hopkins, 1650 Orleans Street, Room 151, Baltimore, MD 21287, USA. E-mail: [email protected] Abstract Recent advances in genetics and genomics have revealed new genes and pathways that are somatically altered in human malignancies. This wealth of knowledge has translated into molecularly defined targets for therapy over the past two decades, serving as key examples that translation of laboratory findings can have great impact on the ability to treat patients with cancer. However, given the genetic instability and heterogeneity that are characteristic of all human cancers, drug resistance to virtually all therapies has emerged, posing further and future challenges for clinical oncology. Here we review the history of targeted therapies, including examples of genetically defined cancer targets and their approved therapies. We also discuss resistance mechanisms that have been uncovered, with an emphasis on somatic genetic alterations that lead to these phenotypes. Copyright 2013 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd. Keywords: targeted therapies; drug resistance; somatic alterations; cancer Received 10 August 2013; Revised 23 August 2013; Accepted 21 September 2013 No conflicts of interest were declared. First-generation targeted therapies Multi-kinase inhibitors The first drug that revolutionized the concept of small molecule kinase inhibitors was imatinib (STI571, Gleevec, Glivec). Imatinib was originally developed in the late 1990s as a targeted therapy against the BCR–ABL protein, present in virtually all chronic myelogenous leukaemia (CML) patients (reviewed in [1]). CML develops from the reciprocal transloca- tion between chromosomes 9 and 22, leading to the ‘Philadelphia chromosome’ (Ph + ), which is a genetic hallmark of this disease [2,3]. This translocation creates the BCR–ABL fusion gene product, encoding a hyper- activated ABL tyrosine kinase. Thus, inhibition of ABL kinase activity would, in theory, specifically deter the proliferation of the malignant Ph + clonal cancer cells, yet non-CML cells would be unaffected. High- throughput screening against protein kinase C (PKC) activity teased out imatinib as having high potential for pharmacological inhibition [1]. Prior to the use of ima- tinib, only 30% of CML patients survived 5 years after diagnosis. After FDA approval of imatinib in 2001, this percentage has steadily risen to nearly 90% [4]. Imatinib occupies an area of the ATP-binding pocket of ABL, locking it in the inhibited or closed confirma- tion [5]. In this way, the normally constitutively active BCR–ABL kinase is turned off and it can no longer phosphorylate downstream targets. However, imatinib is also an inhibitor to other kinases, most notably c-Kit and PDGFR [4]. In hindsight, this promiscuity is in fact advantageous, as imatinib has proved to be an effective treatment for gastrointestinal stromal tumours (GISTs), which exhibit mutated c-KIT or activated PDGFR [6]. Although an undeniable success, at least 10% of patients will not demonstrate prolonged benefit from imatinib therapy [7]. This can be due to initial resistance or an acquired resistance after an initial response to the drug. Several mechanisms for innate and acquired resistance to imatinib have been pro- posed. Amplification or duplication of the BCR–ABL kinase, c-KIT or PDGFR may be one such mode of resistance, in which the increased tyrosine kinase production outcompetes the action of the drug [7,8]. Importantly, point mutations in the ABL gene have also been linked to imatinib resistance, most frequently a threonine-to-isoleucine substitution (T315I) in the kinase domain that prevents imatinib association with the ATP-binding pocket [7–9]. Not only do these mutations confer resistance to imatinib treatment but they also exhibit increases in cancerous phenotypes [10]. Many other mutations of the ABL gene have been identified as contributing to resistant phenotypes, including mutations within the P-loop, SH2 domain, A-loop, C-helix and substrate-binding domain [7]. Mutations in c-KIT and PGDFR are also associ- ated with imatinib resistance in GIST patients [8]. Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254 Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

Upload: ben-ho

Post on 01-Feb-2017

212 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Somatic alterations as the basis for resistance to targeted therapies

Journal of PathologyJ Pathol 2014; 232: 244–254Published online in Wiley Online Library(wileyonlinelibrary.com) DOI: 10.1002/path.4278

INVITED REVIEW

Somatic alterations as the basis for resistance to targetedtherapiesBrian G Blair,1 Alberto Bardelli2 and Ben Ho Park1*

1 Sidney Kimmel Comprehensive Cancer Center at Johns Hopkins, Baltimore, MD, USA2 Institute for Cancer Research and Treatment (IRCC), University of Torino Medical School, Candiolo, Torino, Italy

*Correspondence to: BH Park, Sidney Kimmel Comprehensive Cancer Center at Johns Hopkins, 1650 Orleans Street, Room 151, Baltimore, MD21287, USA. E-mail: [email protected]

AbstractRecent advances in genetics and genomics have revealed new genes and pathways that are somatically alteredin human malignancies. This wealth of knowledge has translated into molecularly defined targets for therapyover the past two decades, serving as key examples that translation of laboratory findings can have great impacton the ability to treat patients with cancer. However, given the genetic instability and heterogeneity that arecharacteristic of all human cancers, drug resistance to virtually all therapies has emerged, posing further andfuture challenges for clinical oncology. Here we review the history of targeted therapies, including examples ofgenetically defined cancer targets and their approved therapies. We also discuss resistance mechanisms that havebeen uncovered, with an emphasis on somatic genetic alterations that lead to these phenotypes.Copyright 2013 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

Keywords: targeted therapies; drug resistance; somatic alterations; cancer

Received 10 August 2013; Revised 23 August 2013; Accepted 21 September 2013

No conflicts of interest were declared.

First-generation targeted therapies

Multi-kinase inhibitorsThe first drug that revolutionized the concept ofsmall molecule kinase inhibitors was imatinib (STI571,Gleevec, Glivec). Imatinib was originally developedin the late 1990s as a targeted therapy against theBCR–ABL protein, present in virtually all chronicmyelogenous leukaemia (CML) patients (reviewed in[1]). CML develops from the reciprocal transloca-tion between chromosomes 9 and 22, leading to the‘Philadelphia chromosome’ (Ph+), which is a genetichallmark of this disease [2,3]. This translocation createsthe BCR–ABL fusion gene product, encoding a hyper-activated ABL tyrosine kinase. Thus, inhibition ofABL kinase activity would, in theory, specifically deterthe proliferation of the malignant Ph+ clonal cancercells, yet non-CML cells would be unaffected. High-throughput screening against protein kinase C (PKC)activity teased out imatinib as having high potential forpharmacological inhibition [1]. Prior to the use of ima-tinib, only 30% of CML patients survived 5 years afterdiagnosis. After FDA approval of imatinib in 2001, thispercentage has steadily risen to nearly 90% [4].

Imatinib occupies an area of the ATP-binding pocketof ABL, locking it in the inhibited or closed confirma-tion [5]. In this way, the normally constitutively activeBCR–ABL kinase is turned off and it can no longer

phosphorylate downstream targets. However, imatinibis also an inhibitor to other kinases, most notably c-Kitand PDGFR [4]. In hindsight, this promiscuity is in factadvantageous, as imatinib has proved to be an effectivetreatment for gastrointestinal stromal tumours (GISTs),which exhibit mutated c-KIT or activated PDGFR [6].

Although an undeniable success, at least 10%of patients will not demonstrate prolonged benefitfrom imatinib therapy [7]. This can be due to initialresistance or an acquired resistance after an initialresponse to the drug. Several mechanisms for innateand acquired resistance to imatinib have been pro-posed. Amplification or duplication of the BCR–ABLkinase, c-KIT or PDGFR may be one such modeof resistance, in which the increased tyrosine kinaseproduction outcompetes the action of the drug [7,8].Importantly, point mutations in the ABL gene havealso been linked to imatinib resistance, most frequentlya threonine-to-isoleucine substitution (T315I) in thekinase domain that prevents imatinib association withthe ATP-binding pocket [7–9]. Not only do thesemutations confer resistance to imatinib treatment butthey also exhibit increases in cancerous phenotypes[10]. Many other mutations of the ABL gene havebeen identified as contributing to resistant phenotypes,including mutations within the P-loop, SH2 domain,A-loop, C-helix and substrate-binding domain [7].Mutations in c-KIT and PGDFR are also associ-ated with imatinib resistance in GIST patients [8].

Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

Page 2: Somatic alterations as the basis for resistance to targeted therapies

Somatic alterations as the basis for resistance to targeted therapies 245

Alternative mechanisms of imatinib resistance havealso been described and include changes in druginflux/efflux, epigenetic modification and activation ofalternative growth pathways independent of tyrosinekinase receptors [7,8].

Following the success of imatinib in treating CMLand GIST, the development of kinase inhibitors to treatcancers began in earnest. Sorafenib and sunitinib aretwo compounds that are multi-kinase inhibitors. Thesedrugs have shown to inhibit various growth signallingpathways mediated by receptor tyrosine kinases [11].Sorafenib and sunitinib interact with the ATP-bindingpocket within the kinase domain of their respective tar-gets, competitively deactivating the growth signallingof the various kinases [12,13]. Originally designed asa RAF kinase inhibitor, sorafenib also targets VEGFR,PDGFR and the RAF kinases CRAF and BRAF,and has been utilized as a successful treatment forrenal and liver cancers [12,14]. Sunitinib is approvedfor use in renal cell carcinoma and GIST, and hasbeen shown to target VEGFR, PDGFR, KIT, RET,FLT3 and other receptor tyrosine kinases [11,15–19].Sunitinib inhibits c-KIT and PDGF in a manner dis-tinctly different than imatinib, and therefore it is clini-cally used for the treatment of imatinib-resistant GIST[20]. The multi-kinase inhibitory properties also makethese drugs less selective for cancer cells, resultingin numerous side-effects for most patients. Resistance

can arise to either or both of the anti-proliferative andanti-angiogenic properties of these kinase inhibitors.Activation of alternative signalling pathways (AKT,PI3K, etc.), drug sequestration, up-regulation of non-VEGF-dependent angiogenesis factors (FGF, IL8, etc.)and pro-angiogenic pericyte and monocyte recruitmentare some of the documented mechanisms of resis-tance to PDGFR and VEGFR inhibitors, includingsorafenib and sunitinib [8,21,22]. In addition, hepato-cyte growth factor/MET signalling has also been pro-posed as an alternative pro-angiogenic pathway con-tributing to VEGF inhibitor resistance [23]. Althoughthese drugs are effective for certain indications, theirlack of specificity and toxicity profiles dictated furtherrefinements in targeted therapies for cancer.

mTOR inhibitorsThe mammalian target of rapamycin (mTOR) isa regulator of cell signalling found downstreamof several growth pathways that are mutated incancers, including the PI3K–AKT signalling cas-cade (Figure 1). mTOR was discovered as a tar-get for the anti-fungal rapamycin in yeast cells[24–26]. Rapamycin interacts with the protein FKBP12(Rbp1), forming a complex that then binds the FRB(FKB) domain of mTOR, thereby inhibiting mTORkinase activity [27,28]. mTOR associates with two

ActivatingKRAS mutations

EGFR RTK

p

p

p

IRS-

1

p85

KRAS

p110α

PIP2 PIP3 PIP2

PTEN

AKT

Cell cycleprogression

Increased cellsurvival

Proteinsynthesis andcell growth

Apoptosis

AlternativeAKT pathway

activation

*

*

*

*

RAF

MEK

ERK

EGFR Antibodies

BRAF V600Emutation

Alternative ERKpathway activation

Understanding Disease

Figure 1. The MAP kinase and PI3 kinase pathways, targeted therapies and somatic alterations leading to resistance. Epidermal growthfactor receptor (EGFR), a receptor tyrosine kinase (RTK), is a target for antibody therapies. The PI3 kinase and MAP kinase pathways thatare activated by aberrant EGFR signalling in cancers are shown. *Mechanisms of resistance at key nodal points.

Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

Page 3: Somatic alterations as the basis for resistance to targeted therapies

246 BG Blair et al

proteins; RAPTOR and RICTOR, forming two com-plexes, mTORC1 and mTORC2, respectively [25,28].Rapamycin predominantly interacts with mTORC1-RAPTOR and inhibits the phosphorylation of the tar-gets S6 kinase 1 (S6K1, p70) and eukaryotic initiationfactor 4E binding protein (4EBP1) [29,30].

The pursuit of mTOR inhibitors as an anti-cancertherapy first gained traction with the discovery ofthe rapamycin analogues (rapalogues) temsirolimus,everolimus and ridaforolimus (deforolimus). Likerapamycin, the rapalogues target FKBP-12 rapamycin-binding domain (FRB, FKB) of the mTORC1 complex[29,30]. Unlike rapamycin, these rapalogues do nothave the same degree of immunosuppressive proper-ties. Temsirolimus and everolimus are prodrugs, suchthat their metabolites are the active rapamycin-likeforms. These compounds inhibit several downstreamtargets of mTOR, such as HIF1α, cyclin D and VEGF,thereby affecting angiogenesis, cell proliferation,cellular metabolism and survival [31]. Temsirolimus,the first rapalogue to be approved for clinical use,is primarily administered for the treatment of renalcancer. In the past few years, everolimus has beenapproved for the treatment of kidney, pancreatic andER positive, HER2-negative breast cancers.

Mutations in FKBP12 or the FKBP12-bindingdomain of mTORC1 can result in rapalogue resistance,due to a drop in binding affinity [32–34]. Additionally,up-regulation of PI3K, AKT, MAPK, PIM kinaseand PDK activity has been observed in response torapamycin and rapalogue treatment [35–37]. Thismechanism of resistance is common among targetedtherapies, that is, disruption of a signalling pathwayleads to activation of alternative pathways by negativefeedback loops. This mechanism of resistance isespecially vexing with rapalogues because mTORC2phosphorylates AKT at serine 473 in response to nega-tive feedback from mTORC1 inhibition [38]. Decreasesin 4E-BP1 protein levels can also confer resistance torapalogues, and increased expression of eIF4E can alsolead to resistance [39,40]. Increases in anti-apoptoticsignals, through up-regulation of Bcl-2 or the IAPs,has also been indicated to contribute to rapalogueresistance [41,42]. It has also been proposed that inhi-bition of HIF-1α by rapalogues can trigger decreases inangiogenesis and limit tumour growth, and alternativeangiogenesis signals may bypass this component ofthe anti-tumour effect of rapalogues [42,43].

Second-generation mTOR inhibitors (mTOR kinaseinhibitors) are being developed that competitivelyinhibit the serine/threonine kinase activity of mTOR[44]. The advantage of this mode of mTOR inhibi-tion is that it is independent of FKBP12 and tar-gets both mTORC1 and mTORC2. Another potentialmethod to overcome the common causes of rapalogueresistance is treatment with rapalogues together withcompounds that target alternative proliferative path-ways, such as the PI3 kinase and/or MAP kinase path-ways. Furthermore, single compounds that block both

PI3K and mTOR are being explored for improvedefficacy [44,45].

Current and future cancer-targeted therapies

CD20 and rituximabOne of the first and most successful antibodies incancer therapy is rituximab, a monoclonal antibody(mAb) against the B cell antigen CD20. CD20, initiallydescribed in 1980, is a member of the membrane-spanning 4-A (MS4A) family of proteins [46–49]. Theclinical significance of this protein was quickly realizedupon the observation that CD20 is highly expressed inB cell lymphomas, hairy cell leukaemias and B cellchronic lymphocytic leukaemias [47,48].

Rituximab is a chimeric mouse/human antibodyagainst CD20 [50]. It was approved to treat non-Hodgkin’s lymphoma and is now approved for useagainst other lymphomas and leukaemias, as wellas autoimmune diseases such as rheumatoid arthri-tis. Rituximab marks CD20-positive cells for destruc-tion by natural killer (NK) cells, neutrophils andmacrophages, thereby selectively targeting malignantB cells for antibody-dependent cellular cytotoxicity(ADCC) [51,52]. The CD20-bound rituximab recruitsthe innate effector cells responsible for ADCC throughan interaction of the Fc region of rituximab and theFcγ receptor (FcγR, CD16) on NK cells [53,54].

Rituximab, while highly effective in many patients,is subject to therapeutic resistance. It is estimated that30–60% of patients may show resistance to rituximab[55]. As there are several mechanisms of rituximabcytotoxicity, there are likewise many means by whichresistance may develop. For example, the destructionof rituximab-bound B cells by immune effector cells isdependent on two variables: the binding of rituximabto FcγR and the congregation of the bound CD20into lipid rafts [54–56]. Thus, decreases in the affinityof FcγR to rituximab or low numbers of NK cellswith FcγR can diminish the efficacy of the treatment[55,56]. Similarly, a lack of rituximab–CD20 polar-ization can inhibit ADCC [51]. Over-expression of theMAP kinase or PI3 kinase pathways, as well as NF-κBhyperphosphorylation, can all contribute to an attenu-ation of the pro-apoptotic nature of rituximab [56,57].While it would seem that the most likely mechanismfor resistance would be changes to CD20 expression oraffinity to rituximab, evidence for this is lacking [55].

BRCA and PARP inhibitorsCarriers of BRCA1 and BRCA2 mutations have upto an 80% lifetime risk of developing breast cancer[58]. These patients also have up to a 55% chance ofdeveloping ovarian cancer [58]. BRCA1 and BRCA2act as part of a complex of proteins, most notablyinteracting with the recombinase RAD1, to repairdouble-strand breaks (DSBs) through homologous

Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

Page 4: Somatic alterations as the basis for resistance to targeted therapies

Somatic alterations as the basis for resistance to targeted therapies 247

recombination (HR) and stabilization of stalled DNAsynthesis [59–62]. Cells defective in BRCA1 orBRCA2 demonstrate increased mutational rates due tothe breakdown of DNA repair, and thus are likely tobecome tumourigenic.

Poly(ADP-ribose) polymerases (PARPs), are afamily of proteins whose primary function is tolocate and repair single-strand DNA breaks (SSBs).PARPs, and in particular PARP1, bind to both single-and double-strand breaks and initiate DNA repair[63–65]. In cases where BRCA1 or BRCA2 functionis compromised, PARPs play an important role in thecellular response to this loss of function. In additionto repairing SSBs and restarting stalled forks, in theabsence of BRCA1/BRCA2 it is believed that PARPscan serve a further role in DSB repair [66,67]. There-fore, in patients with defective BRCA homologousrecombination-mediated DNA repair, PARPs can bekey proteins in regulating the amount of DNA damageaccumulated within cancer cells.

The interplay of BRCA1/2 and PARPs in DNArepair offers a unique target for clinical treatment.Cells lacking HR, due to defective BRCA1, BRCA2or other repair proteins such as PALB2, have provedto be excellent targets for PARP inhibition. In thesecells, loss of PARP activity will increase the numberof SSBs, eventually leading to DSBs when replicationforks become stalled [68]. The vast increase in dam-aged DNA in cells lacking both HR and PARP activityinitiates a synthetic lethal effect [66]. Additionally, itis thought that PARP inhibitors ‘trap’ PARP proteins atthe site of DNA damage and that this PARP ‘trapping’is also toxic to the cell [69]. Both modes of actionfor PARP inhibitors are selective only for cancerouscells that lack HR. Many PARP inhibitors have beendeveloped, including olaparib, veliparib, rucaparib andMK4827.

Mechanisms of PARP inhibitor resistance are justbeginning to be understood. It has been found thatisoforms of BRCA2 can arise in culture that restoresome HR function to cancer cells harbouring BRCA2mutations, allowing these cells to bypass the syn-thetic lethality of PARP inhibition [70,71]. Also, manyBRCA1-defective cancers have low or no PARP1expression, and resistant or refractory cancers oftenshow decreased PARP activity relative to cancersthat show high PARP inhibitor sensitivity [72]. PARPinhibitor resistance can also arise from differentialexpression of other redundant DNA repair pathways.For example, it has been shown that some BRCA1-deficient breast cancers lose 53BP1 expression, achange that can lead to restoration of HR repair andPARP inhibitor resistance [73,74].

HER2HER2 (ErbB2) expression is amplified and the pro-tein over-expressed in approximately 15–20% of breastcancers [75]. HER2 is a transmembrane receptor tyro-sine kinase belonging to the ErbB/HER family, which

includes EGFR (HER1/ErbB1) and HER 2, 3 and 4.While each member of the family shares common tyro-sine kinase domain motifs, they each contain uniqueligand-binding domains [76]. HER2 is activated byautophosphorylation upon homo- and heterodimeriza-tion, and can dimerize with other members of the ErbBfamily. Over-expression of HER2 alone is often ade-quate to trigger receptor phosphorylation and activa-tion, even in the absence of ligand. HER2 activatesseveral proliferative pathways upon phosphorylation,including the MAP kinase and PI3 kinase pathways.

The over-expression of HER2 in breast cancers(and now other cancers, such as stomach cancers)has made it a strong candidate for targeted therapies.In 1998, the FDA approved trastuzumab (Herceptin)as the first targeted treatment against HER2-positivebreast cancers. Trastuzumab is a monoclonal antibodyspecific to the juxtamembrane domain of HER2 andresults in a down-regulation of the protein [77–79].Trastuzumab also has other proposed anti-proliferativeeffects, such as inhibiting transcription and decreasingthe levels of a constitutively active truncated formof HER2 (p95-HER2) [78,80]. Also, similar to themethod of action for rituximab, trastuzumab can triggerantibody-dependent cellular toxicity (ADCC) throughan interaction between the antibody and Fc receptorson immune cells [54].

The successes of trastuzumab are balanced byoccurrences of innate or acquired resistance. Thesemechanisms of resistance include HER2-independentactivation of the PI3 kinase pathway, up-regulation ofthe other HER family member receptors and increasedproduction of p95-HER2 [81,82]. One of the primarysignalling pathways activated by ErbB receptors is thePI3 kinase pathway. PI3K is composed of two subunits,the p85 regulatory subunit and the p110 catalytic com-ponent. These subunits interact with activated HER2,thus beginning a signalling cascade, which ultimatelyresults in cancerous phenotypes. In many cancers, PI3Kor downstream targets, such as AKT and MAPK, canbecome constitutively activated independent of the sig-nalling dependence from HER2 activation [83–85].For example, mutations that lead to loss of PTENfunction, or activation of the PIK3CA gene (PIK3CAencodes for p110α) are frequent in breast cancer [83].In these cases, targeting HER2 over-expression withtrastuzumab potentially lessens its anti-proliferativeeffects, as these cancers can grow without HER2 acti-vation and have been associated with resistance inpreclinical models [86,87].

Redundancy in the downstream signalling pathwaysof receptor tyrosine kinases indicates that loss of sig-nal from one of these receptors may not be enough toimpede MAPK and PI3K activation [88]. As example,the HER2–HER3 heterodimer has been shown to bean essential interaction for HER2-dependent tumouri-genesis [89]. HER2-independent activation of HER3or EGFR may lead to HER2-positive cancers thatlack sensitivity to trastuzumab [89–91]. In addition toredundancy within the members of the ErbB family of

Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

Page 5: Somatic alterations as the basis for resistance to targeted therapies

248 BG Blair et al

membrane receptors, IGF-1R has been demonstratedto activate PI3K through crosstalk with HER2 [92]. Ithas been proposed that IGF-1R stimulation in certaincancer cells may be a source of trastuzumab resis-tance, and inhibition of IGF-1R can restore trastuzumabsensitivity [93].

The truncated p95-HER2 protein lacks the amino-terminus of HER2 and therefore lacks the binding sitefor trastuzumab. Moreover, p95-HER2 is constitutivelyactive, and downstream targets become phosphorylatedindependent of ligand interaction. While trastuzumabtreatment has been shown to decrease p95-HER2expression, 30% of HER2-positive tumours over-express p95-HER2 [78,94]. Thus it has been proposedthat cancers with increased levels of p95-HER2 arerelatively resistant to trastuzumab’s anti-neoplasticeffects.

Additional treatments against HER2 have also beendeveloped. Pertuzumab is another monoclonal antibodydirected against HER2, but interacts with the dimer-ization domain of HER2, blocking the homo- and het-erodimerization of HER2 [82]. This can inhibit not onlyHER2-stimulated growth but also the dimerization-dependent activation of the other members of theErbB family, specifically HER3 [95–97]. Also, unliketrastuzumab, pertuzumab may be able to inhibit p95-HER2-mediated growth [98]. Clinical data indicate thattrastuzumab and pertuzumab are complementary andcan be highly effective when used in combination [99].

There have also been studies to increase the effec-tiveness of trastuzumab, leading to the developmentof TDM-1. TDM-1 is a conjugate of trastuzumabwith DM1 (derivative of maytansine 1), a micro-tubule inhibitor. This antibody–drug conjugate (ADC)is designed to combine the anti-proliferative effectsof trastuzumab with targeted delivery of the cytotoxiccompound DM-1 directly to HER2 over-expressingcells. TDM-1 shows promise in attacking cells that maybe resistant to trastuzumab, and demonstrates reten-tion of trastuzumab’s other anti-neoplastic functions,such as PI3K signal inhibition, decreases in p95-HER2-mediated signalling and ADCC [100,101].

In addition to these antibody-based approaches fortargeting HER2, small molecules that inhibit the kinaseactivity of HER2 have been developed, most notablylapatinib. Lapatanib is a small molecule inhibitor ofEGFR and HER2, which competitively binds to theATP-binding domain, thus blocking autophosphoryla-tion. Due to the specificity for the ATP-binding pocket,lapatinib may have distinct mechanisms of resistancecompared to trastuzumab. Nonetheless, lapatinib resis-tance can still occur via several mechanisms. Acti-vating mutations of the PI3 kinase pathway may alsocontribute to lapatinib resistance, as can loss of PTENfunction [102]. A mutation in the ATP-binding pocketof HER2 (L755S) has also been reported to confer lapa-tinib resistance in vitro [103]. It has also been proposedthat activation of IGF-1R, ER or other receptor tyrosinekinases (such as AXL), may lead to tumours resistantto lapatinib treatment [104,105].

EGFR inhibitorsEGFR (HER1, ErbB1) is over-expressed and/ormutated in many cancers, including lung and coloncancers. EGFR is a receptor tyrosine kinase activatedby binding with epidermal growth factor (EGF), trans-forming growth factor-α (TGFα) and other ligands[88]. Activation of EGFR initiates several cell prolifer-ating signal transduction cascades, including the MAPkinase, PI3 kinase, JNK and STAT pathways [88].To date, several compounds have been developed fortreatment against EGFR-activated cancers, includingsmall molecule inhibitors such as gefitinib and erlotiniband antibody-based therapies such as cetuximab.

Gefitinib (Iressa) and erlotinib (Tarceva) are smallmolecule inhibitors of EGFR which were approvedfor the treatment of EGFR-positive lung cancer inthe past decade. Both compounds reversibly inhibitEGFR kinase activity by competitively binding theATP-binding site [106,107]. It was soon discoveredthat not only do these drugs block EGFR activation,but that the best predictors of response were specificEGFR-activating mutations [108]. These sensitizingmutations were found in the tyrosine kinase domain,either in deletions within exon 19 between codons 746and 759, or in exon 21 at L858R [108,109]. However,while these mutations can increase the sensitivityof tumour cells to gefitinib or erlotinib, subsequentanalyses have identified mutations that can conferdrug resistance. The most clinically relevant resistancemutation in EGFR is in exon 20 at T790M [110,111].This mutation is thought to block the binding ofinhibitors to the ATP-binding site, thus leading to theconstitutive activation of EGFR [112]. This mechanismof resistance may be overcome with the introductionof irreversible EGFR inhibitors [113]. Interestingly,another EGFR mutation, within exon 19 at D761Y,has been linked to gefitinib and erlotinib resistance intumours that contain the L858R mutation [114,115].As in the case with many kinase inhibitors, resistancecan also occur through activation of downstream path-ways, such as PI3K [116]. Likewise, activation of otherreceptor tyrosine kinases, such as MET, IGFR-1, PDGFand additional ErbB family members, has been linkedto acquired gefitinib and erlotinib resistance [97,116].

Cetuximab (Erbitux) and panitumumab (Vectibix)are monoclonal antibodies against EGFR that areclinically used for the treatment of colorectal andother cancers. These antibodies bind to the extra-cellular domain of EGFR and are thought to changethe conformation of the dimerization domain, therebysimultaneously blocking ligand binding and preventingreceptor dimerization [117]. The resultant inhibition ofEGFR phosphorylation combined with several othermechanisms of action contribute towards cetuximab’sand panitumumab’s efficacy against EGFR-positivetumours [118].

One well classified determinant of EGFR antibodytherapeutic sensitivity is the mutational status ofKRAS . Constitutive activation of the MAP kinase

Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

Page 6: Somatic alterations as the basis for resistance to targeted therapies

Somatic alterations as the basis for resistance to targeted therapies 249

signalling pathway downstream of EGFR bypassesthe effectiveness of cetuximab and panitumumab.In fact, patients with KRAS activating mutationshave a poorer predicted outcome than those withwild-type KRAS when treated with EGFR-directedtherapies [119–121]. As a result, KRAS mutationalstatus is screened for prior to the administrationof these therapeutic antibodies. More recently, twoindependent groups have demonstrated the appearanceof KRAS mutations within sites of progressive diseasein metastatic colon cancer patients treated withanti-EGFR antibody therapies [122,123]. Mutations inBRAF , a serine–threonine kinase in the MAP kinasepathway immediately downstream of KRAS , have alsobeen linked to EGFR antibody resistance. Specifically,the V600E activating mutation is strongly associatedto poor response to EGFR antibody therapy [124].Activating mutations of other downstream effectors ofEGFR have also been linked to antibody resistance.For example, constitutive activation of AKT or PI3K,or loss of PTEN, can diminish responses to EGFRantibody therapies [118,125]. Finally, MET amplifi-cation has also been recently shown to be associatedwith acquired resistance to EGFR antibody therapiesin colon cancers that are wild-type for KRAS [126].

BRAFThe cytoplasmic serine/threonine protein kinase BRAFis a critical member of the MAP kinase signal trans-duction pathway. Oncogenic mutations in BRAF havebeen thoroughly described, mutations reported in sev-eral cancer types including melanoma, colorectal, ovar-ian and papillary thyroid carcinomas [13]. Nearly 66%of melanomas harbour BRAF mutations and, of those,80% of the mutations are a single base pair substi-tution in the kinase domain, yielding V600E [127].Currently, there are FDA-approved therapies that targetthe V600E and other V600 mutations in clinical use:vemurafenib and dabrafenib. Vemurafenib, approvedfor clinical use in 2011, is a kinase inhibitor thatselectively targets cells harbouring V600E mutations,but has little effect on wild-type BRAF cells [128].When interacting with mutated BRAF, vemurafeniblocks onto the ATP-binding site and will down-regulateERK signalling, inducing cell cycle arrest and activat-ing apoptosis [128,129]. Treatment with vemurafenibcan be highly effective and many patients demonstratesignificant (up to 80%) clinical response [130]. Morerecently, dabrafenib in combination with the MEKinhibitor trametinib has also been shown as effectivetherapy for metastatic melanomas that harbour BRAFV600 mutations [131].

Up-regulation of PDGFRβ has been recognized as amode of acquired resistance in a subset of melanomas[132]. Alternatively, the up-regulation or mutation ofoncogenic NRAS can also contribute to loss of vemu-rafenib sensitivity [132], further substantiated by arecent clinical report demonstrating mutations in NRASand MEK being associated with vemurafenib-resistant

melanomas [133]. The tumour micro-environment canalso play a role in BRAF inhibitor resistance. Secre-tion of hepatocyte growth factor can stimulate MET,leading to activation of the PI3 kinase and MAP kinasepathways independent of RAF inhibition [134]. In addi-tion, early attempts using vemurafenib in BRAF V600Emutant colon cancers did not demonstrate any apprecia-ble response. Although mutant BRAF has been shownto predict for resistance against EGFR-mediated thera-pies, as mentioned above, recent work suggests an evenmore complex interplay. It appears that mutant BRAFsuppression leads to feedback activation of EGFR incolon cancer cells resulting in vemurafenib resistance,suggesting the need to simultaneously target both pro-teins [135]. Interestingly, it has been shown that vemu-rafenib treatment can paradoxically stimulate MAPkinase signalling and may promote tumourigenesis incells with wild-type RAF [129].

EML4–ALKNearly 5% of non-small cell lung cancer casesare estimated to harbour a somatic translocationbetween echinoderm microtubule-associated protein-like 4 (EML4) and anaplastic lymphoma kinase (ALK)[136]. The resultant EML4–ALK protein and othersproduced from ALK translocations are hyperactive ver-sions of the ALK receptor tyrosine kinase [137,138].Inspired by the success of BCR–ABL targeting withimatinib, a selective inhibitor of ALK was quicklydeveloped. Crizotinib is a competitive inhibitor of thekinase activity that interacts with the ATP-bindingdomain of ALK, ROS1 and MET [139]. Crizotinibinhibits cell proliferation and angiogenesis, initiatescell arrest and can induce apoptosis [140]. With aresponse rate of ∼60%, acquired and innate resistanceto crizotinib has also been documented [138].

Several mutations in ALK have been linked tocrizotinib resistance. The most well-documented ofthese is L1196M, which impedes inhibitor bindingto the active site [141]. Other resistance mutationshave been identified in the kinase domain, as well asmutations that increase ATP affinity, thus decreasingthe effectiveness of competitive inhibition [138].Copy number increases in ALK can also decrease theefficacy of competitive inhibition by crizotinib [142].Alternatively, activation of other proliferative mech-anisms can decrease tumour sensitivity to crizotinib.Specifically, up-regulation of EGFR, HER2 or HER3,constitutive activating mutations in EGFR (L858R andS7681), as well as KRAS G12C and G12V mutations,have been reported to be associated with a decrease inALK inhibitor sensitivity [138,143].

Other targeted therapies and conclusions

The past success and future potential of targeted ther-apies for cancer are evident, and many new drugsare currently in development. An obvious candidate

Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

Page 7: Somatic alterations as the basis for resistance to targeted therapies

250 BG Blair et al

Table 1. Summary of targeted therapies in cancer andmechanisms of resistance

Drug Target Resistance mechanism

Imatinib BCR–ABL BCR–ABL, c-KIT, PDGFRamplification/mutations

c-KIT Abl T315I mutationPDGFR

Sorafenib VEGFR Activation of alternative growthpathways

PDGFR SequestrationC-Raf/B-Raf

(V600E)Angiogenesis up-regulation

Sunitinib VEGFR Activation of alternative growthpathways

PDGFR SequestrationKIT Angiogenesis up-regulationRETFLT3

Rapamycin/rapalogues

mTOR Activation of alternative growthpathways

Mutations in FKBP12Changes in 4E-BP1 or eIF4EUp-regulation of Bcl-2 or IAPsAngiogenesis up-regulation

Rituximab CD20 Changes in mCRP expressionDecrease of C1qDecrease in FcγR affinity

PARPinhibitors

PARP PARP1 mutationsExpression of BCRA2 isoformsLoss of 53BP1

Trastuzumab HER2 Activation of PI3K pathwayHER family up-regulationIncreases in p95-HER2

Lapatanib EGFR Activation of PI3K pathwayHER2 Activation of mTORHER3 HER2 L755S mutation

Activation of IGF-1, ERGefitinib EGFR EGFR mutations: T790M, D761YErlotinib Activation of MET, IGFR-1, PDGF or

ErbB RTKsCetuximab EGFR KRAS mutationsPanitumumab BRAF V600E muation

Activation of alternative growthpathways

Angiogenesis up-regulationVemurafenib B-Raf

(V600E)PDGFR or N-RAS up-regulationHepatocyte growth factor secretionEGFR activation

Crizotinib EML4–ALK ALK mutations (L11196M)ALK copy number increasesEGFR, HER2, HER3 up-regulationEGFR mutations (L858R, S768I)KRAS mutation (G12C, G12V)

for targeted therapy is PI3K, or specifically the p110α

component encoded by the PIK3CA gene. PIK3CA isthe second most commonly mutated gene in the cancergenome, with a relatively high frequency found in lung,breast, brain, gastric, colon and other cancers [83,144].As such, there is currently much interest in devel-oping PI3K inhibitors for cancer therapy [145–147].AKT (protein kinase B) is a serine/threonine kinase inthe PI3K–AKT–mTOR pathway that is often mutatedand/or activated in human cancers [148]. Similar toPIK3CA, AKT is a heavily pursued target for targeted

therapies with numerous inhibitors at various stages ofclinical development. MEK is another kinase down-stream of RAS-RAF signalling within the MAP kinasepathway. Targeting of this kinase has resulted inthe development of several small molecule inhibitors,including trametinib, as mentioned above, and selume-tinib (AZD6244).

Over the past two decades, targeted therapies forcancer have made a significant impact on the morbidityand mortality inflicted by this group of diseases.These historic and recent examples demonstrate thatthe future of cancer medicine lies in the pursuit ofbetter patient and tumour-specific therapies, and thatgenetic knowledge of cancer genomes can help guidethese efforts. Despite the past successes, drug resistancecontinues to be a common problem among targetedtherapies, limiting their use (Table 1). Future researchwill therefore emphasize the use of multiple therapiesthat block cancer cell signalling across many pathways,as well as nodal points within each of these pathways.Clearly, great strides have been made in the assessmentand treatment of cancer, yet much work remainsin finding meaningful long-term therapies for thisdisease.

Acknowledgements

We apologize to our colleagues whose important workcould not be cited due to space constraints. BGBwas supported by a DOD BCRP Postdoctoral Fellow-ship (Award No. BC100972). BHP receives supportfrom the Flight Attendant Medical Research Insti-tute (FAMRI), the Breast Cancer Research Foun-dation, the Avon Foundation, the CommonwealthFoundation, JHU Singapore/Santa Fe Foundation, theSafeway Foundation and the Eddie and Sandy GarciaFoundation.

Author contributions

BGB, AB and BHP wrote and revised the manuscript.

References1. Capdeville R, Buchdunger E, Zimmermann J, et al. Glivec

(STI571, imatinib), a rationally developed, targeted anticancerdrug. Nat Rev Drug Discov 2002; 1(7): 493–502.

2. Nowell P, Hungerford D. Minute chromosome in human chronicgranulocytic leukemia. Science 1960; 1497.

3. Rowley JD. Letter: a new consistent chromosomal abnormality inchronic myelogenous leukaemia identified by quinacrine fluores-cence and Giemsa staining. Nature 1973; 243(5405): 290–293.

4. Waller CF. Imatinib mesylate. Recent Results Cancer Res 2010;184: 3–20.

5. Schindler T. Structural mechanism for STI-571 inhibition ofAbelson tyrosine kinase. Science 2000; 289(5486): 1938–1942.

6. Lopes LF, Bacchi CE. Imatinib treatment for gastrointestinalstromal tumour (GIST). J Cell Mol Med 2010; 14(1–2): 42–50.

Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

Page 8: Somatic alterations as the basis for resistance to targeted therapies

Somatic alterations as the basis for resistance to targeted therapies 251

7. Bixby D, Talpaz M. Seeking the causes and solutions to imatinib-resistance in chronic myeloid leukemia. Leukemia 2011; 25(1):7–22.

8. Engelman JA, Settleman J. Acquired resistance to tyrosine kinaseinhibitors during cancer therapy. Curr Opin Genet Dev 2008;18(1): 73–79.

9. Gorre ME, Mohammed M, Ellwood K, et al. Clinical resistanceto STI-571 cancer therapy caused by BCR-ABL gene mutation oramplification. Science (New York, NY) 2001; 293(5531): 876–880.

10. Griswold IJ, MacPartlin M, Bumm T, et al. Kinase domain mutantsof Bcr-Abl exhibit altered transformation potency, kinase activity,and substrate utilization, irrespective of sensitivity to imatinib. Mol

Cell Biol 2006; 26(16): 6082–6093.11. Faivre S, Demetri G, Sargent W, et al. Molecular basis for

sunitinib efficacy and future clinical development. Nat Rev Drug

Discov 2007; 6(9): 734–745.12. Wilhelm S, Carter C, Lynch M, et al. Discovery and development

of sorafenib: a multikinase inhibitor for treating cancer. Nat Rev

Drug Discov 2006; 5(10): 835–844.13. Wan PTC, Garnett MJ, Roe SM, et al. Mechanism of activation

of the RAF–ERK signaling pathway by oncogenic mutations ofB-RAF . Cell 2004; 116(6): 855–867.

14. Keating GM, Santoro A. Sorafenib: a review of its use in advancedhepatocellular carcinoma. Drugs 2009; 69(2): 223–240.

15. Abrams TJ, Murray LJ, Pesenti E, et al. Preclinical evaluationof the tyrosine kinase inhibitor SU11248 as a single agent andin combination with ‘standard of care’ therapeutic agents forthe treatment of breast cancer. Mol Cancer Therap 2003; 2(10):1011–1021.

16. Baratte S, Sarati S, Frigerio E, et al. Quantitation of SU1 1248,an oral multi-target tyrosine kinase inhibitor, and its metabolitein monkey tissues by liquid chromatograph with tandem massspectrometry following semi-automated liquid-liquid extraction. J

Chromatogr A 2004; 1024(1–2): 87–94.17. Le Tourneau C, Raymond E, Faivre S. Sunitinib: a novel tyrosine

kinase inhibitor. A brief review of its therapeutic potential in thetreatment of renal carcinoma and gastrointestinal stromal tumors(GIST). Therap Clin Risk Manage 2007; 3(2): 341–348.

18. Mendel DB, Laird AD, Xin X, et al. In vivo antitumor activityof SU11248, a novel tyrosine kinase inhibitor targeting vascu-lar endothelial growth factor and platelet-derived growth factorreceptors: determination of a pharmacokinetic/pharmacodynamicrelationship. Clin Cancer Res 2003; 9(1): 327–337.

19. O’Farrell AM, Foran JM, Fiedler W, et al. An innovative phaseI clinical study demonstrates inhibition of FLT3 phosphorylationby SU11248 in acute myeloid leukemia patients. Clin Cancer Res

2003; 9(15): 5465–5476.20. Demetri GD, van Oosterom AT, Garrett CR, et al. Efficacy

and safety of sunitinib in patients with advanced gastrointestinalstromal tumour after failure of imatinib: a randomised controlledtrial. Lancet 2006; 368(9544): 1329–1338.

21. Bergers G, Hanahan D. Modes of resistance to anti-angiogenictherapy. Nat Rev Cancer 2008; 8(8): 592–603.

22. Gotink KJ, Broxterman HJ, Labots M, et al. Lysosomal seques-tration of sunitinib: a novel mechanism of drug resistance. Clin

Cancer Res 2011; 17(23): 7337–7346.23. Shojaei F, Lee JH, Simmons BH, et al. HGF/c-Met acts as

an alternative angiogenic pathway in sunitinib-resistant tumors.Cancer Res 2010; 70(24): 10090–10100.

24. Heitman J, Movva NR, Hall MN. Targets for cell cycle arrest bythe immunosuppressant rapamycin in yeast. Science (New York,

NY) 1991; 253(5022): 905–909.25. Foster KG, Fingar DC. Mammalian target of rapamycin (mTOR):

conducting the cellular signaling symphony. J Biol Chem 2010;285(19): 14071–14077.

26. Brown EJ, Albers MW, Bum Shin T, et al. A mammalian proteintargeted by G1-arresting rapamycin–receptor complex. Nature

1994; 369(6483): 756–758.27. Huang S, Bjornsti M-A, Houghton PJ. Rapamycins: mechanism

of action and cellular resistance. Cancer Biol Ther. 2003; 2(3):222–232.

28. Dobashi Y, Watanabe Y, Miwa C, et al. Mammalian target ofrapamycin: a central node of complex signaling cascades. Int J

Clin Exp Pathol 2011; 4(5): 476–495.29. Bjornsti M-A, Houghton PJ. The TOR pathway: a target for cancer

therapy. Nat Rev Cancer 2004; 4(5): 335–348.30. Faivre S, Kroemer G, Raymond E. Current development of mTOR

inhibitors as anticancer agents. Nat Rev Drug Discov 2006; 5(8):671–688.

31. Yuan R, Kay A, Berg WJ, et al. Targeting tumorigenesis: devel-opment and use of mTOR inhibitors in cancer therapy. J Hematol

Oncol 2009; 2(1): 45.32. Dumont FJ, Staruch MJ, Grammer T, et al. Dominant mutations

confer resistance to the immunosuppressant, rapamycin, in variantsof a T cell lymphoma. Cell Immunol 1995; 163(1): 70–79.

33. Fruman DA, Wood MA, Gjertson CK, et al. FK506 binding protein12 mediates sensitivity to both FK506 and rapamycin in murinemast cells. Eur J Immunol 1995; 25(2): 563–571.

34. Lorenz MC, Heitman J. TOR mutations confer rapamycin resis-tance by preventing interaction with FKBP12–rapamycin. J Biol

Chem 1995; 270(46): 27531–27537.35. Carracedo A, Ma L, Teruya-Feldstein J, et al. Inhibition of

mTORC1 leads to MAPK pathway activation through a PI3K-dependent feedback loop in human cancer. J Clin Invest 2008;118(9): 3065–3074.

36. Shi Y, Yan H, Frost P, et al. Mammalian target of rapamycininhibitors activate the AKT kinase in multiple myeloma cellsby up-regulating the insulin-like growth factor receptor/insulinreceptor substrate-1/phosphatidylinositol 3-kinase cascade. Mol

Cancer Therap 2005; 4(10): 1533–1540.37. Tan J, Lee PL, Li Z, et al. B55β-associated PP2A complex controls

PDK1-directed myc signaling and modulates rapamycin sensitivityin colorectal cancer. Cancer Cell 2010; 18(5): 459–471.

38. Sarbassov DD, Guertin DA, Ali SM, et al. Phosphorylation andregulation of Akt/PKB by the rictor–mTOR complex. Science

(New York, NY) 2005; 307(5712): 1098–1101.39. Dilling MB, Germain GS, Dudkin L, et al. 4E-binding proteins, the

suppressors of eukaryotic initiation factor 4E, are down-regulatedin cells with acquired or intrinsic resistance to rapamycin. J Biol

Chem 2002; 277(16): 13907–13917.40. Graff JR, Konicek BW, Carter JH, et al. Targeting the eukaryotic

translation initiation factor 4E for cancer therapy. Cancer Res

2008; 68(3): 631–634.41. Mahalingam D, Medina EC, Esquivel JA, et al. Vorinostat

enhances the activity of temsirolimus in renal cell carcinomathrough suppression of survivin levels. Clin Cancer Res 2010;16(1): 141–153.

42. Majumder PK, Febbo PG, Bikoff R, et al. mTOR inhibitionreverses Akt-dependent prostate intraepithelial neoplasia throughregulation of apoptotic and HIF-1-dependent pathways. Nat Med

2004; 10(6): 594–601.43. Guba M, von Breitenbuch P, Steinbauer M, et al. Rapamycin

inhibits primary and metastatic tumor growth by antiangiogenesis:involvement of vascular endothelial growth factor. Nat Med 2002;8(2): 128–135.

44. Chapuis N, Tamburini J, Green AS, et al. Perspectives oninhibiting mTOR as a future treatment strategy for hematologicalmalignancies. Leukemia 2010; 24(10): 1686–1699.

45. Kharas MG, Janes MR, Scarfone VM, et al. Ablation ofPI3K blocks BCR–ABL leukemogenesis in mice, and a dual

Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

Page 9: Somatic alterations as the basis for resistance to targeted therapies

252 BG Blair et al

PI3K/mTOR inhibitor prevents expansion of human BCR–ABL+

leukemia cells. J Clin Invest 2008; 118(9): 3038–3050.46. Ishibashi K, Suzuki M, Sasaki S, et al. Identification of a new

multigene four-transmembrane family (MS4A) related to CD20,HTm4 and β subunit of the high-affinity IgE receptor. Gene 2001;264(1): 87–93.

47. Nadler LM, Ritz J, Hardy R, et al. A unique cell surface antigenidentifying lymphoid malignancies of B cell origin. J Clin Invest

1981; 67(1): 134–140.48. Tedder TF, Engel P. CD20: a regulator of cell-cycle progression

of B lymphocytes. Immunol Today 1994; 15(9): 450–454.49. Oettgen HC, Bayard PJ, Van Ewijk W, et al. Further biochemical

studies of the human B-cell differentiation antigens B1 and B2.Hybridoma 1983; 2(1): 17–28.

50. Grillo-Lopez AJ, White CA, Varns C, et al. Overview of theclinical development of rituximab: first monoclonal antibodyapproved for the treatment of lymphoma. Semin Oncol 1999; 26(5,suppl 14): 66–73.

51. Rudnicka D, Oszmiana A, Finch DK, et al. Rituximab causes apolarisation of B cells which augments its therapeutic function inNK cell-mediated antibody-dependent cellular cytotoxicity. Blood

2013; 121: 4694–4702.52. Glennie MJ, French RR, Cragg MS, et al. Mechanisms of killing

by anti-CD20 monoclonal antibodies. Mol Immunol 2007; 44(16):3823–3837.

53. Cartron G, Dacheux L, Salles G, et al. Therapeutic activityof humanized anti-CD20 monoclonal antibody and polymor-phism in IgG Fc receptor FcγRIIIa gene. Blood 2002; 99(3):754–758.

54. Clynes RA, Towers TL, Presta LG, et al. Inhibitory Fc receptorsmodulate in vivo cytotoxicity against tumor targets. Nat Med 2000;6(4): 443–446.

55. Rezvani AR, Maloney DG. Rituximab resistance. Best practiceand research. Clin Haematol 2011; 24(2): 203–216.

56. Weiner GJ. Rituximab: mechanism of action. Semin Hematol

2010; 47(2): 115–123.57. Jazirehi AR, Vega MI, Bonavida B. Development of rituximab-

resistant lymphoma clones with altered cell signaling and cross-resistance to chemotherapy. Cancer Res 2007; 67(3): 1270–1281.

58. Euhus DM, Robinson L. Genetic predisposition syndromes andtheir management. Surg Clin N Am 2013; 93(2): 341–362.

59. Patel K. Involvement of Brca2 in DNA repair. Mol Cell 1998;1(3): 347–357.

60. Moynahan ME, Pierce AJ, Jasin M. BRCA2 is required forhomology-directed repair of chromosomal breaks. Mol Cell 2001;7(2): 263–272.

61. Moynahan ME, Chiu JW, Koller BH, et al. Brca1 con-trols homology-directed DNA repair. Mol Cell 1999; 4(4):511–518.

62. Lomonosov M, Anand S, Sangrithi M, et al. Stabilization of stalledDNA replication forks by the BRCA2 breast cancer susceptibilityprotein. Genes Dev 2003; 17(24): 3017–3022.

63. Gradwohl G, Menissier de Murcia JM, Molinete M, et al.

The second zinc-finger domain of poly(ADP-ribose) polymerasedetermines specificity for single-stranded breaks in DNA. Proc

Natl Acad Sci USA 1990; 87(8): 2990–2994.64. McGlynn P, Lloyd RG. Recombinational repair and restart of

damaged replication forks. Nat Rev Mol Cell Biol 2002; 3(11):859–870.

65. Rouleau M, Patel A, Hendzel MJ, et al. PARP inhibition: PARP1

and beyond. Nat Rev Cancer 2010; 10(4): 293–301.66. Saleh-Gohari N, Bryant HE, Schultz N, et al. Spontaneous

homologous recombination is induced by collapsed replicationforks that are caused by endogenous DNA single-strand breaks.Mol Cell Biol 2005; 25(16): 7158–7169.

67. Thompson LH. Recognition, signaling, and repair of DNA double-strand breaks produced by ionizing radiation in mammalian cells:the molecular choreography. Mutat Res 751(2): 158–246.

68. Helleday T, Petermann E, Lundin C, et al. DNA repair pathways astargets for cancer therapy. Nat Rev Cancer 2008; 8(3): 193–204.

69. Murai J, Huang SN, Das BB, et al. Trapping of PARP1 andPARP2 by clinical PARP inhibitors. Cancer Res 2012; 72(21):5588–5599.

70. Ashworth A. Drug resistance caused by reversion mutation.Cancer Res 2008; 68(24): 10021–10023.

71. Edwards SL, Brough R, Lord CJ, et al. Resistance to therapycaused by intragenic deletion in BRCA2 . Nature 2008; 451(7182):1111–1115.

72. Domagala P, Huzarski T, Lubinski J, et al. PARP-1 expressionin breast cancer including BRCA1-associated, triple negative andbasal-like tumors: possible implications for PARP-1 inhibitortherapy. Breast Cancer Res Treat 2011; 127(3): 861–869.

73. Bouwman P, Aly A, Escandell JM, et al. 53BP1 loss rescuesBRCA1 deficiency and is associated with triple-negative andBRCA-mutated breast cancers. Nat Struct Mol Biol 2010; 17(6):688–695.

74. Bunting SF, Callen E, Wong N, et al. 53BP1 inhibits homologousrecombination in Brca1 -deficient cells by blocking resection ofDNA breaks. Cell 2010; 141(2): 243–254.

75. Sjogren S, Inganas M, Lindgren A, et al. Prognostic and predictivevalue of c-erbB-2 overexpression in primary breast cancer, aloneand in combination with other prognostic markers. J Clin Oncol

1998; 16(2): 462–469.76. Slamon DJ, Godolphin W, Jones LA, et al. Studies of the HER-

2/neu proto-oncogene in human breast and ovarian cancer. Science

(New York, NY) 1989; 244(4905): 707–712.77. Cuello M, Ettenberg SA, Clark AS, et al. Down-regulation

of the erbB-2 receptor by trastuzumab (herceptin) enhancestumor necrosis factor-related apoptosis-inducing ligand-mediatedapoptosis in breast and ovarian cancer cell lines that overexpresserbB-2. Cancer Res 2001; 61(12): 4892–4900.

78. Molina MA, Codony-Servat J, Albanell J, et al. Trastuzumab(herceptin), a humanized anti-Her2 receptor monoclonal antibody,inhibits basal and activated Her2 ectodomain cleavage in breastcancer cells. Cancer Res 2001; 61(12): 4744–4749.

79. Bange J, Zwick E, Ullrich A. Molecular targets for breast cancertherapy and prevention. Nat Med 2001; 7(5): 548–552.

80. Junttila TT, Akita RW, Parsons K, et al. Ligand-independentHER2/HER3/PI3K complex is disrupted by trastuzumab and iseffectively inhibited by the PI3K inhibitor GDC-0941. Cancer Cell

2009; 15(5): 429–440.81. Lim B, Cream L V, Harvey HA. Update on clinical trials:

genetic targets in breast cancer. Adv Exp Med Biol 2013; 779:35–54.

82. Stopeck AT, Brown-Glaberman U, Wong HY, et al. The role oftargeted therapy and biomarkers in breast cancer treatment. Clin

Exp Metast 2012; 29(7): 807–819.83. Bachman KE, Argani P, Samuels Y, et al. The PIK3CA gene is

mutated with high frequency in human breast cancers. Cancer Biol

Therap 2004; 3(8): 772–775.84. Park BH, Davidson NE. PI3 kinase activation and response to

trastuzumab therapy: what’s neu with herceptin resistance? Cancer

Cell 2007; 12(4): 297–299.85. Gustin JP, Cosgrove DP, Park BH. The PIK3CA gene as a mutated

target for cancer therapy. Curr Cancer Drug Targets 2008; 8(8):733–740.

86. Nagata Y, Lan KH, Zhou X, et al. PTEN activation contributesto tumor inhibition by trastuzumab, and loss of PTEN pre-dicts trastuzumab resistance in patients. Cancer Cell 2004; 6(2):117–127.

Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

Page 10: Somatic alterations as the basis for resistance to targeted therapies

Somatic alterations as the basis for resistance to targeted therapies 253

87. Berns K, Horlings HM, Hennessy BT, et al. A functional geneticapproach identifies the PI3K pathway as a major determinant oftrastuzumab resistance in breast cancer. Cancer Cell 2007; 12(4):395–402.

88. Hynes NE, Lane HA. ERBB receptors and cancer: the complexityof targeted inhibitors. Nat Rev Cancer 2005; 5(5): 341–354.

89. Lee-Hoeflich ST, Crocker L, Yao E, et al. A central role forHER3 in HER2-amplified breast cancer: implications for targetedtherapy. Cancer Res 2008; 68(14): 5878–5887.

90. Sergina N V, Rausch M, Wang D, et al. Escape from HER-familytyrosine kinase inhibitor therapy by the kinase-inactive HER3.Nature 2007; 445(7126): 437–441.

91. Ritter CA, Perez-Torres M, Rinehart C, et al. Human breast cancercells selected for resistance to trastuzumab in vivo overexpressepidermal growth factor receptor and ErbB ligands and remaindependent on the ErbB receptor network. Clin Cancer Res 2007;13(16): 4909–4919.

92. Lu Y, Zi X, Zhao Y, et al. Insulin-like growth factor-I receptorsignaling and resistance to trastuzumab (Herceptin). J Natl Cancer

Inst 2001; 93(24): 1852–1857.93. Nahta R, Yuan LXH, Zhang B, et al. Insulin-like growth factor-I

receptor/human epidermal growth factor receptor 2 heterodimer-ization contributes to trastuzumab resistance of breast cancer cells.Cancer Res 2005; 65(23): 11118–11128.

94. Saez R, Molina MA, Ramsey EE, et al. p95HER-2 predictsworse outcome in patients with HER-2-positive breast cancer. Clin

Cancer Res 2006; 12(2): 424–431.95. Franklin MC, Carey KD, Vajdos FF, et al. Insights into ErbB

signaling from the structure of the ErbB2–pertuzumab complex.Cancer Cell 2004; 5(4): 317–328.

96. Adams CW, Allison DE, Flagella K, et al. Humanization of arecombinant monoclonal antibody to produce a therapeutic HERdimerization inhibitor, pertuzumab. Cancer Immunol Immunother

2006; 55(6): 717–727.97. Engelman JA, Zejnullahu K, Mitsudomi T, et al. MET ampli-

fication leads to gefitinib resistance in lung cancer by activat-ing ERBB3 signaling. Science (New York, NY) 2007; 316(5827):1039–1043.

98. Recupero D, Daniele L, Marchio C, et al. Spontaneous andpronase-induced HER2 truncation increases the trastuzumab bind-ing capacity of breast cancer tissues and cell lines. J Pathol 2013;229(3): 390–399.

99. Scheuer W, Friess T, Burtscher H, et al. Strongly enhancedantitumor activity of trastuzumab and pertuzumab combinationtreatment on HER2-positive human xenograft tumor models.Cancer Res 2009; 69(24): 9330–9336.

100. Lewis Phillips GD, Li G, Dugger DL, et al. TargetingHER2-positive breast cancer with trastuzumab–DM1, anantibody–cytotoxic drug conjugate. Cancer Res 2008; 68(22):9280–9290.

101. Junttila TT, Li G, Parsons K, et al. Trastuzumab-DM1 (T-DM1)retains all the mechanisms of action of trastuzumab and efficientlyinhibits growth of lapatinib insensitive breast cancer. Breast

Cancer Res Treat 2011; 128(2): 347–356.102. Hurvitz SA, Hu Y, O’Brien N, et al. Current approaches and future

directions in the treatment of HER2-positive breast cancer. Cancer

Treat Rev 2013; 39(3): 219–229.103. Bose R, Kavuri SM, Searleman AC, et al. Activating HER2

mutations in HER2 gene amplification-negative breast cancer.Cancer Discov 2013; 3(2): 224–237.

104. Liu L, Greger J, Shi H, et al. Novel mechanism of lapatinibresistance in HER2-positive breast tumor cells: activation of AXL.Cancer Res 2009; 69(17): 6871–6878.

105. Xia W, Bacus S, Hegde P, et al. A model of acquired autoresis-tance to a potent ErbB2 tyrosine kinase inhibitor and a therapeutic

strategy to prevent its onset in breast cancer. Proc Natl Acad Sci

USA 2006; 103(20): 7795–7800.106. Barker AJ, Gibson KH, Grundy W, et al. Studies leading to the

identification of ZD1839 (IRESSA): an orally active, selectiveepidermal growth factor receptor tyrosine kinase inhibitor targetedto the treatment of cancer. Bioorg Med Chem Lett 2001; 11(14):1911–1914.

107. Wakeling AE, Guy SP, Woodburn JR, et al. ZD1839 (Iressa):an orally active inhibitor of epidermal growth factor signalingwith potential for cancer therapy. Cancer Res 2002; 62(20):5749–5754.

108. Sharma S V, Bell DW, Settleman J, et al. Epidermal growth factorreceptor mutations in lung cancer. Nat Rev Cancer 2007; 7(3):169–181.

109. Kosaka T, Yatabe Y, Endoh H, et al. Mutations of the epidermalgrowth factor receptor gene in lung cancer: biological and clinicalimplications. Cancer Res 2004; 64(24): 8919–8923.

110. Kobayashi S, Boggon TJ, Dayaram T, et al. EGFR mutation andresistance of non-small-cell lung cancer to gefitinib. N Engl J Med

2005; 352(8): 786–792.111. Pao W, Miller VA, Politi KA, et al. Acquired resistance of lung

adenocarcinomas to gefitinib or erlotinib is associated with asecond mutation in the EGFR kinase domain. PLoS Med 2005;2(3): e73.

112. Yun C-H, Mengwasser KE, Toms A V, et al. The T790M mutationin EGFR kinase causes drug resistance by increasing the affinityfor ATP. Proc Natl Acad Sci USA 2008; 105(6): 2070–2075.

113. Kwak EL, Sordella R, Bell DW, et al. Irreversible inhibitors ofthe EGF receptor may circumvent acquired resistance to gefitinib.Proc Natl Acad Sci USA 2005; 102(21): 7665–7670.

114. Balak MN, Gong Y, Riely GJ, et al. Novel D761Y and commonsecondary T790M mutations in epidermal growth factor receptor-mutant lung adenocarcinomas with acquired resistance to kinaseinhibitors. Clin Cancer Res 2006; 12(21): 6494–6501.

115. Tokumo M, Toyooka S, Ichihara S, et al. Double mutation andgene copy number of EGFR in gefitinib refractory non-small-celllung cancer. Lung Cancer (Amsterdam, Netherlands) 2006; 53(1):117–121.

116. Xu Y, Liu H, Chen J, et al. Acquired resistance of lungadenocarcinoma to EGFR–tyrosine kinase inhibitors gefitiniband erlotinib. Cancer Biol Therap Landes Biosci 2010; 9(8):572–582.

117. Li S, Schmitz KR, Jeffrey PD, et al. Structural basis for inhibitionof the epidermal growth factor receptor by cetuximab. Cancer Cell

2005; 7(4): 301–311.118. Brand TM, Iida M, Wheeler DL. Molecular mechanisms of

resistance to the EGFR monoclonal antibody cetuximab. Cancer

Biol Therap 2011; 11(9): 777–792.119. Di Fiore F, Blanchard F, Charbonnier F, et al. Clinical relevance

of KRAS mutation detection in metastatic colorectal cancer treatedby Cetuximab plus chemotherapy. Br J Cancer 2007; 96(8):1166–1169.

120. De Roock W, Piessevaux H, De Schutter J, et al. KRAS wild-type state predicts survival and is associated to early radiologicalresponse in metastatic colorectal cancer treated with cetuximab.Ann Oncol 2008; 19(3): 508–515.

121. Amado RG, Wolf M, Peeters M, et al. Wild-type KRAS is requiredfor panitumumab efficacy in patients with metastatic colorectalcancer. J Clin Oncol 2008; 26(10): 1626–1634.

122. Diaz LA, Williams RT, Wu J, et al. The molecular evolutionof acquired resistance to targeted EGFR blockade in colorectalcancers. Nature 2012; 486(7404): 537–540.

123. Misale S, Yaeger R, Hobor S, et al. Emergence of KRAS mutationsand acquired resistance to anti-EGFR therapy in colorectal cancer.Nature 2012; 486(7404): 532–536.

Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com

Page 11: Somatic alterations as the basis for resistance to targeted therapies

254 BG Blair et al

124. Di Nicolantonio F, Martini M, Molinari F, et al. Wild-type BRAF

is required for response to panitumumab or cetuximab in metastaticcolorectal cancer. J Clin Oncol 2008; 26(35): 5705–5712.

125. Sartore-Bianchi A, Martini M, Molinari F, et al. PIK3CA muta-tions in colorectal cancer are associated with clinical resistance toEGFR-targeted monoclonal antibodies. Cancer Res 2009; 69(5):1851–1857.

126. Bardelli A, Corso S, Bertotti A, et al. Amplification of the METreceptor drives resistance to anti-EGFR therapies in colorectalcancer. Cancer Discov 2013; 3(6): 658–673.

127. Davies H, Bignell GR, Cox C, et al. Mutations of the BRAF genein human cancer. Nature 2002; 417(6892): 949–954.

128. Tsai J, Lee JT, Wang W, et al. Discovery of a selective inhibitor ofoncogenic B-Raf kinase with potent antimelanoma activity. ProcNatl Acad Sci USA 2008; 105(8): 3041–3046.

129. Hatzivassiliou G, Song K, Yen I, et al. RAF inhibitors prime wild-type RAF to activate the MAPK pathway and enhance growth.Nature 2010; 464(7287): 431–435.

130. Flaherty KT, Puzanov I, Kim KB, et al. Inhibition of mutated,activated BRAF in metastatic melanoma. N Engl J Med 2010;363(9): 809–819.

131. Flaherty KT, Infante JR, Daud A, et al. Combined BRAF andMEK inhibition in melanoma with BRAF V600 mutations. N Engl

J Med 2012; 367(18): 1694–1703.132. Nazarian R, Shi H, Wang Q, et al. Melanomas acquire resistance to

B-RAF(V600E) inhibition by RTK or N-RAS upregulation. Nature

2010; 468(7326): 973–977.133. Trunzer K, Pavlick AC, Schuchter L, et al. Pharmacody-

namic effects and mechanisms of resistance to vemurafenib inpatients with metastatic melanoma. J Clin Oncol 2013; 31(14):1767–1774.

134. Straussman R, Morikawa T, Shee K, et al. Tumour micro-environment elicits innate resistance to RAF inhibitors throughHGF secretion. Nature 2012; 487(7408): 500–504.

135. Prahallad A, Sun C, Huang S, et al. Unresponsiveness of coloncancer to BRAF(V600E) inhibition through feedback activation ofEGFR. Nature 2012; 483(7387): 100–103.

136. Soda M, Choi YL, Enomoto M, et al. Identification of thetransforming EML4–ALK fusion gene in non-small-cell lungcancer. Nature 2007; 448(7153): 561–566.

137. Sasaki T, Rodig SJ, Chirieac LR, et al. The biology and treatmentof EML4–ALK non-small cell lung cancer. Eur J Cancer 2010;46(10): 1773–1780.

138. Peters S, Taron M, Bubendorf L, et al. Treatment and detection ofALK-rearranged NSCLC. Lung Cancer 2013; 81: 145–154.

139. Christensen JG, Zou HY, Arango ME, et al. Cytoreductiveantitumor activity of PF-2341066, a novel inhibitor of anaplasticlymphoma kinase and c-Met, in experimental models of anaplasticlarge-cell lymphoma. Mol Cancer Therap 2007; 6(12, pt 1):3314–3322.

140. Zou HY, Li Q, Lee JH, et al. An orally available small-moleculeinhibitor of c-Met, PF-2341066, exhibits cytoreductive antitumorefficacy through antiproliferative and antiangiogenic mechanisms.Cancer Res 2007; 67(9): 4408–4417.

141. Choi YL, Soda M, Yamashita Y, et al. EML4–ALK mutations inlung cancer that confer resistance to ALK inhibitors. N Engl J Med

2010; 363(18): 1734–1739.142. Katayama R, Shaw AT, Khan TM, et al. Mechanisms of acquired

crizotinib resistance in ALK-rearranged lung cancers. Sci Translat

Med 2012; 4(120): 120ra17.143. Tanizaki J, Okamoto I, Okabe T, et al. Activation of HER family

signaling as a mechanism of acquired resistance to ALK inhibitorsin EML4–ALK -positive non-small cell lung cancer. Clin Cancer

Res 2012; 18(22): 6219–6226.144. Samuels Y, Wang Z, Bardelli A, et al. High frequency of mutations

of the PIK3CA gene in human cancers. Science (New York, NY)

2004; 304(5670): 554.145. Janku F, Wheler JJ, Naing A, et al. PIK3CA mutation H1047R is

associated with response to PI3K/AKT/mTOR signaling pathwayinhibitors in early-phase clinical trials. Cancer Res 2013; 73(1):276–284.

146. Janku F, Wheler JJ, Westin SN, et al. PI3K/AKT/mTOR inhibitorsin patients with breast and gynecologic malignancies harboringPIK3CA mutations. J Clin Oncol 2012; 30(8): 777–782.

147. Marone R, Cmiljanovic V, Giese B, et al. Targeting phospho-inositide 3-kinase: moving towards therapy. Biochim Biophys Acta

2008; 1784(1): 159–185.148. Carpten JD, Faber AL, Horn C, et al. A transforming mutation in

the pleckstrin homology domain of AKT1 in cancer. Nature 2007;448(7152): 439–444.

Copyright 2013 Pathological Society of Great Britain and Ireland. J Pathol 2014; 232: 244–254Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk www.thejournalofpathology.com