当前位置:文档之家› Gastric cancer—molecular and clinical dimensions

Gastric cancer—molecular and clinical dimensions

Gastric cancer—molecular and clinical dimensions
Gastric cancer—molecular and clinical dimensions

Department of Gastrointestinal Medical Oncology (R. Wadhwa, S. Song, Y. Yao, J. A. Ajani ),

Department of Systems Biology (J.?S. Lee ), Department of

Epidemiology (Q. Wei ), University of Texas MD Anderson Cancer Center, 1515 Holcombe Boulevard (FC10.3022), Houston, TX 77030, USA.

Correspondence to: J. A. Ajani

jajani@https://www.doczj.com/doc/8d14967670.html,

Gastric cancer—molecular and clinical dimensions

Roopma Wadhwa, Shumei Song, Ju-Seog Lee, Yixin Yao, Qingyi Wei and Jaffer A. Ajani

Abstract | Gastric cancer imposes a considerable health burden around the globe despite its declining incidence. The disease is often diagnosed in advanced stages and is associated with a poor prognosis for patients. An in-depth understanding of the molecular underpinnings of gastric cancer has lagged behind many other cancers of similar incidence and morbidity, owing to our limited knowledge of germline susceptibility traits for risk and somatic drivers of progression (to identify novel therapeutic targets). A few germline (PLCE1) and somatic (ERBB2, ERBB3, PTEN, PI3K/AKT/mTOR, FGF , TP53, CDH1 and MET ) alterations are emerging and some are being pursued clinically. Novel somatic gene targets (ARID1A, FAT4, MLL and KMT2C ) have also been identified and are of interest. Variations in the therapeutic approaches dependent on geographical region are evident for localized gastric cancer—differences that are driven by preferences for the adjuvant strategies and the extent of surgery coupled with philosophical divides. However, greater uniformity in approach has been noted in the metastatic cancer setting, an incurable condition. Having realized only modest successes, momentum is building for carrying out more phase III comparative trials, with some using biomarker-based patient selection strategies. Overall, rapid progress in biotechnology is improving our molecular understanding and can help with new drug discovery. The future prospects are excellent for defining biomarker-based subsets of patients and application of specific therapeutics. However, many challenges remain to be tackled. Here, we review representative molecular and clinical dimensions of gastric cancer.

Wadhwa, R. et al. Nat. Rev. Clin. Oncol. advance online publication 24 September 2013; doi:10.1038/nrclinonc.2013.170

Introduction

Globally, the incidence of gastric cancer ranks as the fourth most frequent cancer in men and fifth in women, but its death rate is comparable to lung cancer—the most common cancer worldwide.1 Histologically, gastric cancer can be mostly diffuse or intestinal type. Diffuse gastric cancer (DGC) represents poorly differentiated cancer cells interspersed with stromal cells. By contrast, intestinal gastric cancer (IGC) comprises cancer cells that form gland-like tubular structures with limited stromal components.2

In 2008, approximately 989,600 (8% of all cancers) new cases of gastric cancer worldwide and 738,000 (10% of all cancer deaths) gastric-cancer-specific deaths were reported. Of these, 70% of deaths occurred in develop-ing nations—with China reporting approximately 40%—owing to the high incidence of gastric cancer in these regions,1 with Asia, Eastern Europe and South America being most affected. Despite these seemingly worrying numbers, the incidence of gastric cancer globally has been declining since World War II;3 several factors that have contributed to the decline include improved living standards 3–6 and early detection strategies, which have reduced the death rate in Japan.7 Helicobacter pylori infection as a risk factor is of importance for preven-tive strat e gies by developing strategic elimination of

the bacteria in high-risk areas.8 Furthermore, major molecular biology advances have identified small subsets of gastric cancers defined by biomarkers. Of these, the over e xpression of HER2 protein and amplification of its gene ERBB2 have led to new treatment approaches in gastric cancer, such as trastuzumab. At present, addi-tional biomarkers (for example, c-MET, PI3K/Akt/mTOR and HER3) are currently being explored. Greater understanding of the molecular biology and immune biology of gastric cancer is still lacking, but likely to be highly rewarding.

This Review focuses on the aspects of the emerging molecular biology of gastric cancer that might lead to improved therapeutics. We discuss the complexity of genomic alterations that drive the disease and show how these present potential opportunities for exploit-ation. The details of molecular signalling are largely not covered if their translational value is unclear. We discuss the role of tumour-initiating cells and the role of many oncogenes in conferring secondary resistance to therapy. In addition, we emphasize the progress—or lack thereof—made in the clinical arena by outlining recent phase III trial results. We aim to highlight advances in molecular and clinical research that reflect the current understanding of gastric cancer, rather than provide an encyclopaedic reference. Details of preventive strategies, impact of new classifications and nuances of surgery and radiotherapy are beyond the scope of this Review.

Competing interests

The authors declare no competing interests.

REVIEWS

Helicobacter pylori as a risk factor

The risk factors for gastric cancer include old age, smoking, alcohol consumption, above normal body weight, high salt and fat consumption, low vegetable and fruit consumption, low economic status, perni-cious anaemia, other chronic gastric diseases and H. pylori infection.6,9 Of these risk factors, H. pylori is the most fascinating (Figure 1) and best studied. The bacterium is associated with a range of effects, from chronic inflamma t ion to epigenetic modulations that can promote t r ansformation, invasion and metastasis of host cells.

H. pylori infection increases the risk of gastric cancer threefold to sixfold,10 and is associated with distal gastric cancer and IGC.11 Chronic active gastritis is an integral part of bacteria-related gastric cancer:11,12 the organism attaches to gastric epithelial cells, leading to inflammation and an increase in the reactive oxygen and nitrogen species, which causes tissue damage.13,14 Furthermore, H. pylori also induces the expression of Toll-like receptors (TLRs), which leads to proliferation of the gastric epithelium.15 The oncogene CagA (now known as S100A8) is produced by carcinogenic H. pylori strains;12,16,17 the gene is included in the cag patho g enicity island and is translocated by the bacterium into the host epithelial cytosol.16,18 Once in the host cell, CagA is phosphory l ated by Src and c-Abl kinases19,20 to its active form to exert a range of effects,such as forming a complex with the SRC homology 2-domain (SH2)-containing tyrosine phosphatase SHP-2 (encoded by PTPN11), resulting in cytoskeletal reorganization that can induce malignant transformation.21 CagA activates the ERK/MAP kinase cascade, resulting in ETS domain-containing protein Elk-1 (Elk-1) phosphorylation and increased proto-oncogene c-Fos (c-Fos) transcription.22 In addition, CagA promotes invasion through activation of the hepatocyte growth factor receptor c-MET, which leads to an oncogenic response.23 CagA also induces E-cadherin-mediated impairment of cell adhesion junctions leading to cytoplasmic and nuclear accumu-lation of β-catenin.24 The protein CagA can bind Crk adaptor (p38) proteins (Crk-II, Crk-I and Crk-L)25 and kinase PAR1,26 eliciting loss of cell polarity, which is an oncogenic event. Furthermore, CagA stimulates the inflammatory pathway by inducing the secretion of cytokines IL-8, IL-1 and TNF-α27–29 through NF-κB in the epi t helium.30 Proinflammatory IL-1 gene cluster polymorphisms (in IL1B, encoding IL-1β, and in the loci for IL1RN and its receptor antagonist) increase the risk of noncardia gastric cancer (that is, cancer that affects the lower stomach).31,32 CagA also upregulates cyclo-oxygenase 2 (COX-2),27,33 which in turn induces the expression of oncogenic prostaglandins.34,35

Aside from its CagA-mediated effects, H. pylori reduces the expression of the proapoptotic FAS-associated factor 1 (encoded by FAF1).36 The bacteri u m also mediates increases in another oncoprotein, a q uaporin-3 (encoded by AQP3),37alters DNA methyl-ation of E-cadherin (encoded by CDH1), which is an oncogenic event,38,39 and promotes methylation of the tumour suppressor TFF240 and RUNX3,41 as well as likely promoting the methylation of six other tumour suppres-sors (FLNC, HAND1, THBD, ARPC1B, HRASLS and LOX).42 Finally, H. pylori can lead to aberrant expres-sion of the nucleic acid editing enzyme activation-induced cytidine deaminase (AID), which can promote carcino g enesis through gene editing.43 These elucidated pathways and mechanisms show that H. pylori is carcino-genic, albeit in susceptible individuals. These improved and detailed understandings of H.pylori-induced host events will lead to plausible targets for drug development to reduce the risk of gastric cancer.

and its several virulence factors, such as CagA, interact with gastric epithelial cells to induce chronic inflammation, mucosal damage and multiple alterations in gene expression and genetic and epigenetic changes, eventually leading to gastric carcinogenesis. Abbreviations: COX-2, cyclooxygenase-2; CpG island, areas of cytosine and guanine repeats; LPS, lipopolysaccharide; RNS, reactive nitrogen species; ROS, reactive oxygen species; VacA, vacuolating cytoxin A.

The genomic profile of gastric cancer

The regions that have a high prevalence of H. pylori—for example, South America, Asia, Africa and Eastern Europe—have few cases of gastric cancer, which implies that genetic susceptibility might be a critical factor to consider.44 Indeed, several studies of single nucleotide polymorphisms (SNPs) and genome-wide association studies (GWAS) have revealed some plausible genes implicated in gastric cancer.

Rare germline mutations (<0.001% of the general popu lation)in CDH1 have been implicated in familial cases of gastric cancer.45,46 Intuitively, SNPs can facilitate gastric cancer such that one adverse allele contributes weakly, but multiple adverse alleles can considerably increase the risk.47 Prior investigations of SNPs have focused on genes involved in mucosal protection

against H. pylori infection (for example, IL1B, IL1RN and TNFA), carcino g en metabolism (for example, CYP2E1 and GSTM1), deoxynucleotide synthesis (for example, MTHFR and TYMS), DNA repair (for example, MTHFR and XRCC1) and tumour suppressors (for example, TP53 and CDH1). However, these studies have had limited yield and none can be used clinically because SNP studies require customized approaches (with a priori assumptions that alterations in certain functional SNPs would increase susceptibility to gastric cancer). In spite of the correlations between certain SNPs and gastric cancer, prospective validation requiring large, population-based studies have been lacking as they are labour-intensive and resource-intensive.

GWAS have been used to scan the whole genome to identify SNPs that are implicated in the disease (Table 1). Such studies have identified genes not previously known to be involved in gastric cancer—for example, PLCE1 (encoding pancreas-enriched phospholipase C) has an oncogenic role in skin and intestinal cancers,48 but is now thought to have a role in gastric cancer. A Japanese research group documented that SNPs in PSCA (encod-ing prostate stem-cell antigen) was associated with diffuse gastric cancer.49,50 The researchers genotyped 188 cases and 752 controls for 85,576 SNPs and then valid data were obtained from 749 cases and 750 controls for 2,753 SNPs. Although the intronic rs2976392 SNP in PSCA was identi f ied as the risk allele and the SNP was in disequilibrium with rs2294008 (located in exon 1), the function of PSCA in gastric tissue is unclear, although its possible tumour-suppressing function in the gastric epi-thelium has been suggested.51 A second study included 1,077 oesophageal cancer (which bears many similari-ties to gastric cardia cancer in terms of environmental risk factors) cases and 1,733 controls and identified 18 ‘hits’ that were validated in 2,766 cases of gastric cardia cancer; PLCE1 rs2274223 and C20orf54 rs1304295 SNPs were shown to be associated with gastric cancer risk.48 Another study of 2,240 gastric cancer cases and 3,302 controls identified the PLCE1 rs2274223 SNP as a risk allele for gastric cardia cancer.52 Indeed, PLCE1 SNPs have been shown to be associated with risk of gastric cancer53,54 and poor prognosis in Chinese patients,55 but not white patients.56 The fourth GWAS included 1,006 cases and 2,273 controls in China and validated

the genomic data in 3,288 cases and 3,069 controls; SNP

rs13361707 (located between PTGER4 and PRKAA1)

and the ZBTB20 rs9841504 SNP were associated with

increased risk of gastric cancer.57

Cancer stem cells and aberrant pathways

Gastric carcinogenesis is complex and not fully character-

ized.58 Although IGC develops after systematic progres-

sion from the preneoplastic stages, DGC is thought to

arise de novo as the result of downregulation (mutation or

promoter methylation) of CDH1,59,60 thereby permitting tumorigenesis and progression. Nevertheless, accumu-

lated genetic alterations—mutations, amplifications,

insertions, deletions and recombinations—contribute to

gastric cancer,61,62 with more alterations accumulating as

the cancer progresses (Table 2). Furthermore, increasing

evidence points to the existence of gastric cancer stem

cells (CSCs) that initiate malignancy by self-renewal and differentiation. Although the origin of human gastric

CSCs is still unclear, the mesenchymal stem cells in the

bone marrow have been suggested to have a role.63,64

Despite the efforts to genotype IGC and DGC (and iden-

tify novel subtypes),65–76 clinically relevant and robust

molecular subtypes have yet to emerge; in one study, IGC

and DGC were found not to be genotypically distinct, but

did respond differently on the basis of their histopatho-

logical characteristics.73 However, these are preliminary

results. Further detailed analyses of the implicated cells

and pathways might point to viable targets that can

be exploited in gastric cancer; indeed, some of these

s t rategies have been studied in other cancer types.

CSCs are known to undergo epithelial– m esenchymal

transition (EMT), activate oncogenic pathways77 and

activate embryogenic signalling pathways,78 which

are essential for the self-renewal and maintenance of

tumours. EMT leads to the CSC-like phenotypes.79 CSCs

are generally resistant to stress caused by reactive oxygen

species, which in turn makes them relatively resistant to

radiotherapy and chemotherapy. Additionally, CSCs have

self-renewal capacities. These cells depend on the Wnt,

Notch and Hedgehog pathways for maintenance.80 Four

pleiotropic transcriptional factors (Snail, Slug, Twist and

REVIEWS

Zeb1/2) orchestrate the EMT and related processes,81 whereas c-MET and TGF-β signalling can be critical for transformation to EMT.82 c-MET activation can induce the reprogramming transcription factors known to support embryonic stem cells and induce differentiated cells to form the pluripotent stem (iPS) cells.83

TGF-β can be pro-oncogenic by inducing matrix deposition, immunosuppression and EMT.84–86 TGF-β signalling drives EMT and CSC self-renewal by target-ing microRNAs (miRNAs),81 upregulating Snail family members and repressing E-cadherin.87 Downstream of c-MET and the TGF-β receptor, PI3K/Akt/mTOR signal l ing conveys prosurvival messages that lead to CSC expansion and maintenance.88 Consequently, there has been interest in targeting mTOR with metformin to inhibit CSCs in breast cancer (and possibly other cancers).89 Ras, along with Hedgehog, helps to maintain CSCs:90,91 Ras alters cellular metabolism to provide an advantage to CSCs to gather biomass to support cellular division. Ras also facilitates tolerance to DNA damage during cell cycling by inhibiting the activity of tumour suppressors. Hedgehog components facilitate rapid cell division in CSCs, as is their function in develop i ng embryos. Furthermore, TGF-β is overexpressed in DGC92 and stimulates collagen synthesis and subsequent fibro-sis. TGF-β can be antiapoptotic through trans a ctivation of EGFR.93 The bone morphogenetic proteins (BMPs), which are members of the TGF-β superfamily, activate PI3K/Akt.94 Given the considerable interest in the inhibi-tion of angiogenesis, VEGF and VEGFR overexpression, which is common in IGC through activation of NF-κB by H. pylori, is a potential target.95,96

G astric CSCs express CD133, CD44, aldehyde dehydro g enase 1 (ALDH1) and ATP-binding cassette sub-family G member 2 (ABCG2); CD44 and ALDH1 are associated with therapy resistance and might be valid therapeutic targets.97 CD44 imparts stem-cell-like proper-ties by promoting synthesis of intracellular reduced gluta-thione.98 Among the other plausible targets in gastric cancer, HER family members have been of interest.99 The oncogenic properties of HER are conferred through the RAS/MEK/MAPK and PI3K/Akt/mTOR pathways.100,101 Overexpression of HER2 caused by ERBB2 amplifica-tion is more prevalent in IGC than DGC.102,103 Thus, confirm i ng the HER2 status in patients with IGC is more important than in patients with DGC; in DGC HER over-expression is rare and might not be driving the cancer. HER2 interacts with EGFR (also known as HER1), HER3104 and insulin-like growth factor 1 (IGF-1R),105 which are amplified, overexpressed106,107 or acquire acti-vating mutations in cancer.108–110ERBB3 mutations have been reported in 10% of gastric cancers.111 Constitutive activation of c-MET can trigger prolifer-ation and antiapoptotic signals.112 Amplification or over-expression of c-MET, rather than mutation of its gene, can activate receptor tyrosine kinase function.113,114 c-MET overexpression or amplification is common in DGC as well as IGC.115–117 Amplified c-MET crosstalk can acti-vate EGFR, HER2 and HER3 to establish a s i gnalling network leading to constitutive PI3K/Akt signal-ling.118–120 This information enables the understanding of the mechanisms of primary and second a ry resistance when the c-MET pathway is inhibited and can lead to ration a l therapeutic strategies in the future, including dual pathway inhibition. Gastric cancers with c-MET overexpression coexpress EGFR, HER3 or both,121 which is clinically relevant for strategies looking to inhibit both targets concurrently.117,120,121 Indeed, the PI3K/Akt/mTOR pathway is frequently altered as a result of amplification or overexpression (PIK3CA, AKT1), activating muta-tions (PIK3CA)59,122 or loss of PTEN.123 Overexpression of phosphory l ated mTOR can occur in DGC;59 HER3 and FGFR amplification in DGC is another mechanism for PI3K/Akt activation.124,125

Although HER, c-MET, PI3K/Akt/mTOR, VEGFR and VEGF have been targeted in gastric and other cancer types, TGF-β and E-cadherin are not targetable because loss of function of a gene is difficult to restore in patients (although it can be achieved in cell lines). Several novel targets worth mentioning include chromatin modifiers, such as ARID1A, MLL3, MLL and FAT4 (cell adhesion), which are of increasing interest and importance.75,126 Frequent mutations in the genes mentioned above have been observed in 47% of gastric cancers and can poten-tially be therapeutic targets. As we uncover the complex-ity of RNA regulation of intracellular processes, more targets are likely to emerge, for example, targeting the mutations in the spliceosome machinery.

MicroRNAs

miRNAs have a role in tumorigenesis, tumour pro-gression and metastasis, and will—in all likelihood—c o ntinue to gain importance in diagnostics, monitoring during therapy and as therapeutic targets (Figure 2).127 Oncogenic microRNAs (oncomiRs) are overexpressed and inhibit tumour suppressors, leading to cell prolifer-ation, invasion and reduced apoptosis. For example, over-expression of miR-296-5p in gastric cancer cells increased cell prolifer a tion and inhibition of apoptosis via the repres-sion of tumour suppressor CDX1.128 Furthermore, over-expression of miR-301a directly targets tumour suppressor RUNX3,129 whereas miR-17-5p/20a targets tumour protein p53-inducible nuclear protein 1 (TP53INP1).130 miR-18a levels were correlated with those of survivin (also known as baculoviral inhibitor of apoptosis repeat-containing 5; BIRC5), Bcl-xL (B-cell lymphoma-extra large) and c-Myc, all three being downstream transcriptional targets of STAT3. STAT3-induced transcription can be negatively regulated by PIAS3. Thus, miR-18a acts as an oncomiR by negatively regulating PIAS3.131 miRNA-372 is onco-genic as it targets TNFAIP1 and modulates NF-κB signal-ling in gastric cancer cells.132IRX1, a newly identified tumour suppressor gene, is inacti v ated by miR-544.133 miR-10b is highly expressed in IGC and is associated with the depth of invasion, lymph-node involvement and metastatic progression.134

Tumour suppressor microRNAs (tsmiRs), by contrast,

are downregulated miRNAs that facilitate the activity of target oncogenes. miR-195 and miR-378 have been shown to be downregulated in gastric cancer (in vivo and in vitro); their target oncogenes are CDK6 and VEGF.135 The oncogene target of miR-133b is FGFR1, which is often amplified in DGC.136 The target of miR-29c is Mcl?1; acti-vation of this miRNA by celecoxib represses Mcl?1 and promotes apoptosis of gastric cancer cells.137 miR-34a is down r egulated in gastric cancer cells; its oncogene target is BIRC5.138 miR-145 suppresses protein C-ets-1 (ETS1), also known as p54, by binding to its three-prime untranslated region (3'-UTR) and reducing the onco-genic processes.139 Let-7i is frequently down r egulated in most tumours and is prognostic of lymphatic invasion, nodal metastasis and poor pathological tumour response in patients with gastric cancer.140 Furthermore, ZFX has a role the maintenance of CSCs and is the target of miR-144 in gastric cancer.141 miR-101 targets the 3'-UTR of PTGS2 (also known as COX2) mRNA and its down-regulation in gastric cancer correlates with over e xpression of COX-2 and cell prolifer a tion.142 Finally, miRNA-146a inhibits NF-κB by targeting CARD10 and COPS8 in gastric cancer.143

All miRNAs are stable in serum, plasma, gastric juice and other bodily fluids, which makes them ideal bio-markers for the disease.144 Indeed, miR-21 and miR-106a were shown to be overexpressed in the gastric juice of patients compared with normal controls.145 Additionally, levels of miR-421 in the gastric juice of patients was higher than in controls (P <0.001) and it resulted in the earlier diagnosis of gastric cancer than by serum carcino-embryonic antigen.146 Similarly, plasma miR-106b, miR-20a and miR-221 levels were elevated in patients

compared with healthy controls (P <0.05).147 In another

study, plasma levels of miRNA-199a-3p were associ-

ated with tumour invasion, lymph-node involvement

and

metasta ses.148 Circulating miR-17-5p and miR-20a

(miR-17-5p/20a) have also been detected in patients

with gastric cancer and both of these miRNAs corre-

lated with overall survival and relapse-free survival.149

Finally, the miR-200c expression level in the blood in

patients with gastric cancer were significantly higher

than in the blood of normal controls (P =0.018).150 Clearly,

research on several miRNAs holds promise and needs to

be fully developed.

Mechanisms of resistance

Treatment resistance is inevitable; the major reasons

for treatment failure in patients are the occurrence of

primary and secondary resistance. For example, HER2

is the only validated biomarker in gastric cancer, and

HER2-overexpressing tumours regularly become resist-

ant to targeted agents, such as trastuzumab. Accordingly, understanding the underlying mechanisms of resistance

is important. In HER2-overexpressing tumours, the

predominant mechanism of resistance is compensa-

tory signalling by other cell surface receptors.151 HER2-overexpressing cells—when inhibited—reprogramme

other oncogenes, including IGF1R, MET, GDF15 (encod-

ing growth differentiation factor 15) and other members

of the ERBB family.152 The IGF1R-mediated resistance

to HER2 targeting involves the PI3K pathway, leading to

enhanced degradation of p27Kip1,105,153,154 whereas activated

Figure 2 | microRNA targets and functions in gastric cancer. Some oncomiRs are overexpressed in tumours and inhibit tumour suppressors, leading to cell proliferation, invasion and reduced apoptosis. By contrast, tsmiRs normally target oncogenes that are downregulated in tumours and facilitate the activity of their target oncogenes. Abbreviations: EMT, epithelial–mesenchymal transition; miR, microRNA; oncomiRs, oncogenic microRNAs; tsmiRs, tumour-suppressor microRNAs.

REVIEWS

c-MET mediates resistance in gastric cancer cells155,156 by restoring shared downstream signalling in the MAPK and Akt pathways.155 Increased levels of EGFR and HER3 ligands can overcome HER2 inhibition.88,124,152 Constitutively activated p95HER2, a truncated HER2 receptor, is the most intriguing mechanism of resistance in response to the blockade of the extracellular domain of HER2.157 This method of conferring resistance to trastu-zumab is mediated by the regulatory DNA that engages a specific protease gene to truncate the receptor. Thus, resist-ance is not caused by acquiring a mutation in the ERBB2 gene itself, rather another gene modifies the receptor and renders the truncated receptor constitutively active. Membrane mucins, such as mucin-4, interact with HER2 in HER2-overexpressing breast cancer cells, resulting in epitope masking that blocks trastuzumab binding.158 Other factors that confer resistance to trastuzumab include focal adhesion kinase (FAK), Src and altera-tions in cell-cycle regulators.159 Constitutive activation of PI3K caused by activating PIK3CA mutations,61 reduced PTEN expression160 or deregulated signalling can induce resistance to HER2 inhibition.61,161,162 Finally, STAT3 activation can mediate resistance as a result of produc-tion of IL-6.163 These data suggest that cancer cells have many redundant mechanisms to overcome resistance to therapy. Accordingly, we have considerable work ahead of us, including exploring antibody–drug c o njugates and immune modulation therapies.

Clinical considerations

Much clinical development is underway in the setting of gastric cancer, which includes many of the targets dis-cussed herein. Many of the recently completed phase III trials should have an impact on patient care and future research directions (Figure 3). Notably, any discussion regarding clinical developments and the future of gastric cancer care must be made with consideration of the regional differ e nces, both in terms of epidemiology and standard practice. The epidemiology and predominant location of primary gastric cancer varies geographically164 because of variations in genetic susceptibility, predomi-nance of certain histological phenotypes (for example,

FGFR2 amplification (9%), VEGF/VEGFR overexpression (36–40%), EGFR amplification and overexpression (27–44%), HER-2 amplification and overexpression (7–34%), c-MET amplification (10–15%), kRAS mutation (2–20%), Raf mutation (0–3%),

PI3K mutation (4–36%), phospho-Akt expression (29–86%), phospho-mTOR expression (60–88%), PTCH1 overexpression (16%), SMO overexpression (12%) and HER3 mutations (10%, not shown). *No clinical trials of these agents have yet been reported in gastric cancer. ?No known numbers or percentages for these genes and pathways. Abbreviations: EGFR, epidermal growth factor receptor; FGFR, fibroblast growth factor receptor; GLI, glioma-associated oncogene family zinc finger 1; HDAC, histone deacetylase; HER, human epidermal growth factor receptor; HGF, hepatocyte growth factor;

Hh, Hedgehog; IGFR, insulin-like growth factor receptor; MMP, matrix metalloproteinase; mTOR, mammalian target

of rapamycin; PDGFR, platelet-derived growth factor receptor; Ptch-1, protein patched homolog 1; Smo, smoothened; VEGF, vascular endothelial growth factor; VEGFR, vascular endothelial growth factor receptor.

IGC is frequent in the endemic areas) and carcinogenic determinants, including H. pylori.165 Gastric cardia cancer is more common in the West, whereas noncardia gastric cancer is more common in the endemic regions. Besides these differ e nces, the surgical approach is more comprehensive in Asia—where a D2 dissection (lymph node dissection that includes regional nodes outside the perigastric area) is routine166—than in the West, where D0 (without formal node dissection) or D1 (dissection limited to perigastric nodes) approaches are preferred. Other regional differences are reflected in the adjuvant strategy for localized gastric cancer (LGC). In the West, postoperative chemoradiotherapy or preoperative chemo-therapy has been embraced as the adjunctive strategy for patients with localized gastric cancer.167 By contrast, in Southeast Asia, two prospective trials (ACTS-GC168 and CLASSIC)169 established the benefit of post o perative adju-vant chemo t herapy. Many of these factors might account for d i fferences in survival of patients in different regions.166

Localized gastric cancer

LGC can be staged as clinical T1 (cT1) or higher with or without regional nodes. Once a patient is diagnosed with LGC, a multidisciplinary evaluation (by medical oncolo-gists, surgical oncologists, surgeons, gastroenterologists, pathologists, radiation oncologists, geneticists [if appro-priate] and nutritionists) is highly recommended prior to initiating any therapy.170 Endoscopic resection for a T1 lesion, when feasible, is recommended. For tumours not amenable to effective endoscopic therapy, surgery should be considered. However, adjunctive therapies (preoperative and postoperative) have contributed to the higher cure rates than those obtained by surgery.167,171–173 Indeed, in North America and Europe, results from the INT-0116167 and Medical Research Council Adjuvant Gastric Infusional Chemotherapy (MAGIC)174 trials have established specific adjunctive strategies. In Asia, post-operative adjuvant chemotherapy has been established as the standard of care.168,169

Postoperative adjuvant chemoradiotherapy

Despite safety concerns, such as haematological effects (leukopenia) and treatment-related deaths (cardiac events, pulmonary fibrosis and sepsis as a complication of myelosuppression) during the INT-0116 phase III trial (conducted in the USA) of observation after surgery versus adjuvant chemoradiotherapy (5- f luorouracil, leuco v orin and radiotherapy) after surgery, chemoradio-therapy improved the 5-year cure rate by ~10%.167 This advantage prevailed with longer follow-up durations.175 However, a subsequent US-based trial coordinated by Cancer and Leukemia Group B produced negative results when investigating the adjuvant chemotherapy strategy (epirubicin, cisplatin and 5- f luorouracil).176 Similarly, the ARTIST trial used the INT-0116 strategy, but differed in two aspects: the control group was treated with chemo-therapy and all patients had undergone a D2 gastrec-tomy.177 The trial failed to demonstrate a benefit for those receiving c hemoradiotherapy,178 raising more questions than it answered.Postoperative chemotherapy

The benefits of adjuvant chemotherapy after D2 gastrec-

tomy were first established in Japan using S-1 (an oral fluoropyrimidine) as the adjuvant.168 The Adjuvant Chemotherapy Trial of TS-1 for Gastric Cancer (ACTS-

GC)168 randomly assigned 1,059 patients to 1 year of S-1

or observation. The primary analysis demonstrated a 33%

improvement in overall survival for the S-1 group, results

that prevailed after a longer follow-up duration (5 years).178

A second Asian study, the Capecitabine and Oxaliplatin

Adjuvant Study in Stomach Cancer (CLASSIC trial) ran-

domly assigned 1,035 patients who had undergone D2

gastrectomy to capecitabine plus oxaliplatin for 6 months

or observation,169 documenting benefit for chemotherapy

for the end point of disease- f ree survival (at 3 years; hazard

ratio [HR]=0.56, 95% CI, 0.44–0.72; P <0.0001). A meta-

analysis based on data from 3,710 patients showed a 7%

improvement in overall survival for fluorouracil-based

postoperative chemotherapy when compared with surgery

alone,173 but this evidence is not robust; the analysis com-

bined many studies from different time periods with

differing eligibility criteria and therapeutic approaches,

making it difficult to make a firm conclusion.

Perioperative or preoperative chemotherapy

The MAGIC trial, which randomly assigned 504 patients

(74% of whom had gastric cancer), established the evi-

dence for perioperative chemotherapy (epirubicin, cispla-

tin and 5-fluorouracil) for gastroesophageal cancer in

the West.174 A second, French trial (in which 25% of the

patients had gastric cancer) terminated early because of

the lack of interest, but demonstrated benefit for preopera-

tive chemotherapy (5-fluorouracil and cisplatin).172Several

trials using a variety of adjuvant strategies are currently

ongoing (Table 3).179

Advanced gastric cancer

Most targeted therapies investigated in unenriched (that

is, patients are not selected based on a biomarker) patient

populations have produced disappointing results (Table 3).

Despite these results, the infrastructure for conduct i ng

regional or global phase III trials has ripened and timely

patient accrual is less of a problem. However, many ques-

tions regarding the lack of progress for patients with

advanced gastric cancer remain. Glaring unanswered ques-

tions include: would further understanding of molecular

biology of gastric cancer lead to selection of appropriate

patient subsets and, therefore, more effective therapy? Are

regional differ e nces in gastric cancers (and patient genet-

ics) so substantial that global trials in unenriched popula-

tions will not be fruitful? Do second-line and third-line

therapies produce substantial survival advantages as

to jeopardize the primary end point of overall survival

required by most regulatory agencies for first-line phase III

trials? Certainly, many other unknown factors are likely to

contribute to the disappointing results observed thus far in

advanced gastric cancer. We stress that use of a single bio-

logical agent (such as trastuzumab) in enriched populations

(HER2-overexpressing tumours) provides only marginal

advantage because of the presence of primary resistance

REVIEWS

or development of secondary resistance. In today’s era of advanced bio t echnology, we are perhaps witnessing the end of empirical investigative strat e gies whereby the new agent is investigated in patients with similar stages with the assumption that the genomic makeup of all tumours and genetics of all patients is similar, which will certainly impact future trials in this cancer type.

First-line therapy

Level 1 evidence for an advantage in overall survival is available for only a few therapeutic agents (docetaxel,180 cisplatin181 and trastuzumab).182 Most trials have been disappointing with the exception of the trastuzumab trial that investigated a HER2-enriched population.182 However, the longer follow-up duration has reduced the hazard ratio for overall survival.183 In the primary analysis,182 the hazard ratio for overall survival was 0.74, but reduced to 0.80. The difference in median overall survival was reduced from 2.7 months to merely 1.4 months, representing an approxi-mate 50% decrease in the effect of trastuzumab, which suggests that only a few patients benefit. Notable among trials with disappointing results are two randomized studies investi g ating the value of EGFR inhibition (EXPAND and REAL-3).184,185 In the EXPAND trial,184 patients were

administered capecitabine and cisplatin, with (experimen-tal) or without (control) cetuximab. Median progression-free survival was 4.4 months in the experimental group compared with 5.6 months in the control group. In the REAL-3 trial,185 panitumumab was added to epicrubicin, oxaliplatin and capecitabine. The experimental (four-agent) group had shorter overall survival (8.8 months) com-pared with the control (three-agent) group (11.3 months; HR 1.37). In both trials, the results demonstrated that the addition of anti-EGFR antibodies (cetuximab or panitu-mumab) to chemo t herapy did not confer a benefit. Except for the ToGA trial,182 all studies investigating a biological agent were c o nducted in an unenriched population. Amongst the failed trials in advanced gastric cancer, was the AV AGAST trial, which was conducted in 774 patients who were randomly assigned to chemotherapy with or without bevacizumab (which targets VEGF-A). This trial did not meet its primary end point of overall survival.186 Retrospective biomarker evaluation of plasma VEGF-A and neuropilin-1 levels seemed to be associated with beva-cizumab efficacy; high plasma VEGF-A and low tumour n e uropilin- 1 levels seemed to correlate with improved overall survival, but neither of these biomarkers is vali-dated.187 Another noteworthy trial is the First line Advanced Gastric Cancer Study (FLAGS), conducted in >1,000 patients who were randomly assigned to receive S-1 plus cisplatin versus 5-fluorouracil plus cisplatin, which also did not meet its primary end point of an overall survival advantage.188

Second-line therapy

A small phase III Arbeitsgemeinschaft Internistische Onkologie (AIO) trial of 40 patients randomly assigned to irinotecan with best supportive care (BSC) or BSC alone showed that overall survival could be extended using irino-tecan plus BSC.189 The randomized GRANITE-1 study (NCT00879333)190 included >600 patients and compared everolimus with placebo in the second-line or third-line settings, but did not achieve its primary end point of improved overall survival.191 However, the REGARD trial (NCT00917384)192 that compared BSC with or without ramucirumab (which targets VEGFR2) in 345 patients demonstrated a marginal improvement in overall survival for ramucirumab (P =0.047),193 suggesting that ramuci-rumab has only a borderline effect. More impressively, the randomized Cougar-02 trial of approximately 186 patients assigned to BSC or docetaxel plus BSC demon s trated a signifi c ant prolongation of overall survival in the combi-nation arm. The patients who received docetaxel survived 44% longer than those who did not receive docetaxel; overall survival was 5.2 months versus 3.6 months, respec-tively (P =0.01, HR 0.67).194 Lapatinib—a dual inhibitor of HER2 and EGFR—was investigated in a phase III study (TYTAN) in >300 patients assigned to either lapatinib or placebo, but the primary end point of prolonged overall survival was not achieved.195

Targeting the c-MET pathway in these patients is also of interest. In a small study with crizotinib, two out of four patients with five or more MET copy number gains had a longer response duration than those with fewer than five copies of the gene.196 Furthermore, rilotumumab (AMG 102), a fully human monoclonal antibody, demon-

strated longer overall survival for patients whose tumours

had high total c-MET expression.197

Conclusions

Despite considerable advances in biotechnology that have

improved our understanding of cancer, immense com-

plexities confront us. Although gastric cancer lags behind

many other tumour types in terms of genetic sequenc-

ing and specially designed therapies, more progress is

anticipated. Greater understanding of the pathogenesis

of IGC by H. pylori is poised to help with increasingly

sophisticated preventive strategies. Germline susceptibil-

ity investigations have uncovered novel genes, but clinical implementation has proven problematic and more work

is needed. Clinically, pro g ress has been slow, but adjuvant

strategies are now frequently employed in patients with

LGC around the world, which should be considered an

important advancement. Future progress will be propelled

by further improvements in biotechnologies to produce

better b i omarkers and drugs.

Given our current knowledge, patients who are ≤45 years

old should be recommended genetic counselling followed

by genetic testing, if needed. Patients with advanced gastric

cancer could benefit from HER2/neu testing.151 The value

of assessing other somatic alterations in gastric cancer is

currently ongoing, but results from ongoing large-scale

genomic sequencing projects will yield insights into the

landscape of genomic alterations. Banking of research blood

(and other bodily fluids) as well as tumour tissue from all

patients for studies should enable detailed examination of

this heterogeneous cancer, and is highly desirable but chal-

lenging. Interrogation of blood for specific alterations in

circulating DNA, RNA, peptides, proteins and tumour cells

holds particular promise for the early detection of gastric

cancer and its m o nitoring during therapy and surveillance.

All patients with LGC must undergo multidisciplinary

evaluation and discussion by experts representing each

major discipline. LGC is a complex disease and patients

should be treated only by experienced high-volume physi-

cians and at centres with excellent infrastructure. Patients

with clinical T1aN0 gastric cancer should be evaluated for

endoscopic therapy. Patients with clinical >T1bN0 gastric

cancer should be offered adjunctive therapy (preoperative,

postoperative or both). Suitable patients with advanced

disease should be offered second- l ine systemic therapy. Investigations of therapeutic agents that take advantage

of the host immune system and mechanisms (including

a n

tibody–drug conjugates) hold immense promise.

REVIEWS

1. Jemal, A. et al. Global cancer statistics. CA Cancer

J. Clin.61, 69–90 (2011).

2. National Cancer Institute at the National Institutes

of Health. Cellular Classification of Gastric Cancer

[online], https://www.doczj.com/doc/8d14967670.html,/cancertopics/

pdq/treatment/gastric/HealthProfessional/

page2 (2013).

3. Bertuccio, P. et al. Recent patterns in gastric

cancer: a global overview. Int. J. Cancer125,

666–673 (2009).

4. Chen, J., Bu, X. L., Wang, Q. Y., Hu, P. J. &

Chen, M. H. Decreasing seroprevalence of

Helicobacter pylori infection during 1993–2003

in Guangzhou, southern China. Helicobacter12,

164–169 (2007).

5. Kawakami, E., Machado, R. S., Ogata, S. K.

& Langner, M. Decrease in prevalence of

Helicobacter pylori infection during a 10-year

period in Brazilian children. Arq. Gastroenterol.45,

147–151 (2008).

6. Tkachenko, M. A. et al. Dramatic changes in the

prevalence of Helicobacter pylori infection during

childhood: a 10-year follow-up study in Russia.

J. Pediatr. Gastroenterol. Nutr.45, 428–432

(2007).

7. Lee, K. J. et al. Gastric cancer screening and

subsequent risk of gastric cancer: a large-scale

population-based cohort study, with a 13-year

follow-up in Japan. Int. J. Cancer118, 2315–2321

(2006).

8. Parkin, D. M. The global health burden of

infection-associated cancers in the year 2002.

Int. J. Cancer118, 3030–3044 (2006).

9. Axon, A. Review article: gastric cancer and

Helicobacter pylori. Aliment. Pharmacol. Ther.

16 (Suppl. 4), 83–88 (2002).

10. Kim, S. S., Ruiz, V. E., Carroll, J. D. & Moss, S. F.

Helicobacter pylori in the pathogenesis of gastric

cancer and gastric lymphoma. Cancer Lett.305,

228–238 (2011).

11. Conteduca, V. et al.H. pylori infection and gastric

cancer: state of the art (review). Int. J. Oncol.42,

5–18 (2013).

12. Rizzato, C. et al. Risk of advanced gastric

precancerous lesions in Helicobacter pylori

infected subjects is influenced by ABO blood group and cagA status. Int. J. Cancer 133, 315–322

(2013).

13. Correa, P. & Houghton, J. Carcinogenesis of

Helicobacter pylori. Gastroenterology133,

659–672 (2007).

14. Tsugawa, H. et al. Reactive oxygen species-

induced autophagic degradation of Helicobacter

pylori CagA is specifically suppressed in cancer

stem-like cells. Cell Host Microbe12, 764–777

(2012).

15. Pimentel-Nunes, P. et al.Helicobacter pylori

induces increased expression of Toll-like receptors and decreased toll-interacting protein in gastric

mucosa that persists throughout gastric

carcinogenesis. Helicobacter18, 22–32 (2013). 16. Peek, R. M. Jr et al.Helicobacter pylori cagA+

strains and dissociation of gastric epithelial cell

proliferation from apoptosis. J. Natl Cancer Inst.

89, 863–868 (1997).

17. Graham, D. Y. & Yamaoka, Y. H. pylori and cagA:

relationships with gastric cancer, duodenal ulcer,

and reflux esophagitis and its complications.

Helicobacter3, 145–151 (1998).

18. Hatakeyama, M. Helicobacter pylori cagA—a

potential bacterial oncoprotein that functionally

mimics the mammalian Gab family of adaptor

proteins. Microbes Infect.5, 143–150 (2003). 19. Mueller, D. et al. c-Src and c-Abl kinases control

hierarchic phosphorylation and function of the

CagA effector protein in Western and East Asian

Helicobacter pylori strains. J. Clin. Invest.122,

1553–1566 (2012).20. Müller, A. Multistep activation of the

Helicobacter pylori effector CagA. J. Clin. Invest.

122, 1192–1195 (2012).

21. Higashi, H. et al. SHP-2 tyrosine phosphatase as

an intracellular target of Helicobacter pylori CagA

protein. Science295, 683–686 (2002).

22. Meyer-ter-Vehn, T., Covacci, A., Kist, M. &

Pahl, H. L. Helicobacter pylori activates mitogen-

activated protein kinase cascades and induces

expression of the proto-oncogenes c-fos and c-jun.

J. Biol. Chem.275, 16064–16072 (2000).

23. Schirrmeister, W. et al. Ectodomain shedding of

E-cadherin and c-Met is induced by Helicobacter

pylori infection. Exp. Cell Res.315, 3500–3508

(2009).

24. Murata-Kamiya, N. et al.Helicobacter pylori cagA

interacts with E-cadherin and deregulates the

β-catenin signal that promotes intestinal

transdifferentiation in gastric epithelial cells.

Oncogene26, 4617–4626 (2007).

25. Suzuki, M. et al. Interaction of CagA with Crk plays

an important role in Helicobacter pylori-induced

loss of gastric epithelial cell adhesion. J. Exp. Med.

202, 1235–1247 (2005).

26. Hatakeyama, M. Linking epithelial polarity and

carcinogenesis by multitasking Helicobacter pylori

virulence factor cagA. Oncogene27, 7047–7054

(2008).

27. Bartchewsky, W. Jr et al. Effect of

Helicobacter pylori infection on IL-8, IL-1β and

COX-2 expression in patients with chronic gastritis

and gastric cancer. Scand. J. Gastroenterol.44,

153–161 (2009).

28. Li, C. Q., Pignatelli, B. & Ohshima, H.

Coexpression of interleukin-8 and inducible nitric

oxide synthase in gastric mucosa infected with

cagA+ Helicobacter pylori. Dig. Dis. Sci.45, 55–62

(2000).

29. Suganuma, M. et al. TNF-α-inducing protein, a

carcinogenic factor secreted from H. pylori, enters

gastric cancer cells. Int. J. Cancer 123, 117–122

(2008).

30. Brandt, S., Kwok, T., Hartig, R., Konig, W. &

Backert, S. NF-κB activation and potentiation of

proinflammatory responses by the Helicobacter

pylori cagA protein. Proc. Natl Acad. Sci. USA102,

9300–9305 (2005).

31. Rad, R. et al. Synergistic effect of Helicobacter

pylori virulence factors and interleukin-1

polymorphisms for the development of severe

histological changes in the gastric mucosa.

J. Infect. Dis.188, 272–281 (2003).

32. Zambon, C. F. et al. Pro- and anti-inflammatory

cytokines gene polymorphisms and

Helicobacter pylori infection: interactions influence

outcome. Cytokine29, 141–152 (2005).

33. Chang, Y. J. et al. Induction of cyclooxygenase-2

overexpression in human gastric epithelial cells by

Helicobacter pylori involves TLR2/TLR9 and

c-Src-dependent nuclear factor-κB activation.

Mol. Pharmacol.66, 1465–1477 (2004).

34. Walduck, A. K. et al. Identification of novel

cyclooxygenase-2-dependent genes in

Helicobacter pylori infection in vivo. Mol. Cancer8,

22 (2009).

35. Seo, J. H., Kim, H. & Kim, K. H. Cyclooxygenase-2

expression by transcription factors in

Helicobacter pylori-infected gastric epithelial cells:

comparison between HP 99 and NCTC 11637.

Ann. NY Acad. Sci.973, 477–480 (2002).

36. Liu, A. Q., Ge, L. Y., Ye, X. Q., Luo, X. L. & Luo, Y.

Reduced FAF1 expression and Helicobacter

infection: correlations with clinicopathological

features in gastric cancer. Gastroenterol. Res.

Pract. 2012, 153219 (2012).

37. Wang, G. et al. Involvement of aquaporin 3 in

Helicobacter pylori-related gastric diseases.

PLoS ONE7, e49104 (2012).

38. Tamura, G. et al. E-cadherin gene promoter

hypermethylation in primary human gastric

carcinomas. J. Natl Cancer Inst.92, 569–573

(2000).

39. Becker, K. F. & H?fler, H. Frequent somatic allelic

inactivation of the E-cadherin gene in gastric

carcinomas. J. Natl Cancer Inst.87, 1082–1084

(1995).

40. Peterson, A. J. et al.Helicobacter pylori infection

promotes methylation and silencing of trefoil

factor 2, leading to gastric tumor development

in mice and humans. Gastroenterology139,

2005–2017 (2010).

41. Katayama, Y., Takahashi, M. & Kuwayama, H.

Helicobacter pylori causes RUNX3 gene

methylation and its loss of expression in gastric

epithelial cells, which is mediated by nitric oxide

produced by macrophages. Biochem. Biophys.

Res. Commun.388, 496–500 (2009).

42. Yoshida, T. et al. Altered mucosal DNA methylation

in parallel with highly active Helicobacter pylori-

related gastritis. Gastric Cancer https://www.doczj.com/doc/8d14967670.html,/

10.1007/s10120-012-0230-x.

43. Matsumoto, Y. et al.Helicobacter pylori infection

triggers aberrant expression of activation-induced

cytidine deaminase in gastric epithelium.

Nat. Med.13, 470–476 (2007).

44. Holcombe, C. Helicobacter pylori: the African

enigma. Gut33, 429–431 (1992).

45. Guilford, P. et al. E-cadherin germline mutations in

familial gastric cancer. Nature392, 402–405

(1998).

46. Huntsman, D. G. et al. Early gastric cancer in

young, asymptomatic carriers of germ-line

E-cadherin mutations. N. Engl. J. Med.344,

1904–1909 (2001).

47. Hu, Z., Ajani, J. A. & Wei, Q. Molecular

epidemiology of gastric cancer: current status and

future prospects. Gastrointest. Cancer Res.1,

12–19 (2007).

48. Wang, L. D. et al. Genome-wide association study

of esophageal squamous cell carcinoma in

Chinese subjects identifies susceptibility loci at

PLCE1 and C20orf54. Nat. Genet.42,

759–763 (2010).

49. Study Group of Millennium Genome Project for

Cancer. Genetic variation in PSCA is associated

with susceptibility to diffuse-type gastric cancer.

Nat. Genet.40, 730–740 (2008).

50. Wang, T., Zhang, L., Li, H., Wang, B. & Chen, K.

Prostate stem cell antigen polymorphisms and

susceptibility to gastric cancer: a systematic

review and meta-analysis. Cancer Epidemiol.

Biomarkers Prev.21, 843–850 (2012).

51. Saeki, N., Gu, J., Yoshida, T. & Wu, X. Prostate

stem cell antigen: a Jekyll and Hyde molecule?

Clin. Cancer Res.16, 3533–3538 (2010).

52. Abnet, C. C. et al. A shared susceptibility locus in

PLCE1 at 10q23 for gastric adenocarcinoma and

esophageal squamous cell carcinoma. Nat. Genet.

42, 764–767 (2010).

53. Zhang, H. et al. Genetic variants at 1q22 and

10q23 reproducibly associated with gastric cancer

susceptibility in a Chinese population.

Carcinogenesis32, 848–852 (2011).

54. Wang, M. et al. Potentially functional variants of

PLCE1 identified by GWASs contribute to gastric

adenocarcinoma susceptibility in an eastern

Chinese population. PLoS ONE7, e31932 (2012).

55. Luo, D. et al. Genetic variation in PLCE1 is

associated with gastric cancer survival in a

Chinese population. J. Gastroenterol.46,

1260–1266 (2011).

56. Palmer, A. J. et al. Genetic variation in C20orf54,

PLCE1 and MUC1 and the risk of upper

gastrointestinal cancers in Caucasian

populations. Eur. J. Cancer Prev.21, 541–544

(2012).

REVIEWS

57. Shi, Y. et al. A genome-wide association study

identifies new susceptibility loci for non-cardia

gastric cancer at 3q13.31 and 5p13.1.

Nat. Genet.43, 1215–1218 (2011).

58. Wu, K. C. et al. Molecular basis of therapeutic

approaches to gastric cancer. J. Gastroenterol.

Hepatol.24, 37–41 (2009).

59. Wong, H. & Yau, T. Targeted therapy in the

management of advanced gastric cancer: are we

making progress in the era of personalized

medicine? Oncologist17, 346–358 (2012).

60. Schneider, B. G. et al. Promoter DNA

hypermethylation in gastric biopsies from subjects at high and low risk for gastric cancer.

Int. J. Cancer127, 2588–2597 (2010).

61. Markman, B., Dienstmann, R. & Tabernero, J.

Targeting the PI3K/Akt/mTOR pathway—beyond

rapalogs. Oncotarget1, 530–543 (2010).

62. Chalhoub, N. & Baker, S. J. PTEN and the PI3-

kinase pathway in cancer. Annu. Rev. Pathol.4,

127–150 (2009).

63. Bergfeld, S. A. & DeClerck, Y. A. Bone marrow-

derived mesenchymal stem cells and the tumor

microenvironment. Cancer Metastasis Rev.29,

249–261 (2010).

64. Takaishi, S., Okumura, T. & Wang, T. C. Gastric

cancer stem cells. J. Clin. Oncol.26, 2876–2882

(2008).

65. Chen, C. N. et al. Gene expression profile predicts

patient survival of gastric cancer after surgical

resection. J. Clin. Oncol.23, 7286–7295 (2005).

66. Cho, J. Y. et al. Gene expression signature-based

prognostic risk score in gastric cancer.

Clin. Cancer Res.17, 1850–1857 (2011).

67. Deng, N. et al. A comprehensive survey of genomic

alterations in gastric cancer reveals systematic

patterns of molecular exclusivity and co-

occurrence among distinct therapeutic targets.

Gut61, 673–684 (2012).

68. Ivanova, T. et al. Integrated epigenomics identifies

BMP4 as a modulator of cisplatin sensitivity in

gastric cancer. Gut62, 22–33 (2013).

69. Kim, H. K. et al. A gene expression signature of

acquired chemoresistance to cisplatin and

fluorouracil combination chemotherapy in gastric

cancer patients. PLoS ONE6, e16694 (2011). 70. Kim, H. K. et al. Three-gene predictor of clinical

outcome for gastric cancer patients treated with

chemotherapy. Pharmacogenomics J.12,

119–127 (2012).

71. Ooi, C. H. et al. Oncogenic pathway combinations

predict clinical prognosis in gastric cancer.

PLoS Genet.5, e1000676 (2009).

72. Takeno, A. et al. Gene expression profile

prospectively predicts peritoneal relapse after

curative surgery of gastric cancer. Ann. Surg.

Oncol.17, 1033–1042 (2010).

73. Tan, I. B. et al. Intrinsic subtypes of gastric cancer,

based on gene expression pattern, predict

survival and respond differently to chemotherapy.

Gastroenterology141, 476–485 (2011).

74. Wu, Y. et al. Comprehensive genomic meta-

analysis identifies intra-tumoural stroma as a

predictor of survival in patients with gastric

cancer. Gut26, 1100–1111 (2013).

75. Zang, Z. J. et al. Exome sequencing of gastric

adenocarcinoma identifies recurrent somatic

mutations in cell adhesion and chromatin

remodeling genes. Nat. Genet.44, 570–574

(2012).

76. Zouridis, H. et al. Methylation subtypes and large-

scale epigenetic alterations in gastric cancer.

Sci. Transl. Med.4, 156ra140 (2012).

77. Tan, D. S., Gerlinger, M., Teh, B. T. & Swanton, C.

Anti-cancer drug resistance: understanding the

mechanisms through the use of integrative

genomics and functional RNA interference.

Eur. J. Cancer46, 2166–2177 (2010).78. Subramaniam, D., Ramalingam, S.,

Houchen, C. W. & Anant, S. Cancer stem cells:

a novel paradigm for cancer prevention and

treatment. Mini Rev. Med. Chem.10, 359–371

(2010).

79. Mani, S. A. et al. The epithelial–mesenchymal

transition generates cells with properties of stem

cells. Cell133, 704–715 (2008).

80. Peacock, C. D. & Watkins, D. N. Cancer stem cells

and the ontogeny of lung cancer. J. Clin. Oncol. 26,

2883–2889 (2008).

81. Peinado, H., Olmeda, D. & Cano, A. Snail, Zeb and

bHLH factors in tumour progression: an alliance

against the epithelial phenotype? Nat. Rev. Cancer

7, 415–428 (2007).

82. Voutsadakis, I. A. The ubiquitin-proteasome

system and signal transduction pathways

regulating epithelial mesenchymal transition

of cancer. J. Biomed. Sci.19, 67 (2012).

83. Li, Y. Q. et al. c-Met signaling induces a

reprogramming network and supports the

glioblastoma stem-like phenotype. Proc. Natl Acad.

Sci. USA 108, 9951–9956 (2011).

84. Roberts, A. B. & Wakefield, L. M. The two faces of

transforming growth factor β in carcinogenesis.

Proc. Natl Acad. Sci. USA100, 8621–8623

(2003).

85. Bierie, B. & Moses, H. L. TGFβ: the molecular

Jekyll and Hyde of cancer. Nat. Rev. Cancer6,

506–520 (2006).

86. Watabe, T. & Miyazono, K. Roles of TGF-β family

signaling in stem cell renewal and differentiation.

Cell Res.19, 103–115 (2009).

87. Singh, A. & Settleman, J. EMT, cancer stem cells

and drug resistance: an emerging axis of evil in

the war on cancer. Oncogene29, 4741–4751

(2010).

88. Zhou, J. B. et al. Activation of the PTEN/mTOR/

STAT3 pathway in breast cancer stem-like cells is

required for viability and maintenance. Proc. Natl

Acad. Sci. USA 104, 16158–16163 (2007).

89. Hirsch, H. A., Iliopoulos, D., Tsichlis, P. N. &

Struhl, K. Metformin selectively targets cancer

stem cells, and acts together with chemotherapy

to block tumor growth and prolong remission.

Cancer Res.69, 7507–7511 (2009).

90. Singh, B. N., Fu, J., Srivastava, R. K. & Shankar, S.

Hedgehog signaling antagonist GDC-0449

(vismodegib) inhibits pancreatic cancer stem cell

characteristics: molecular mechanisms.

PLoS ONE6, e27306 (2011).

91. Heindl, S. et al. Relevance of MET activation and

genetic alterations of KRAS and E-cadherin for

cetuximab sensitivity of gastric cancer cell lines.

J. Cancer Res. Clin. Oncol.138, 843–858 (2012).

92. Komuro, A. et al. Diffuse-type gastric carcinoma:

progression, angiogenesis, and transforming

growth factor β signaling. J. Natl Cancer Inst.101,

592–604 (2009).

93. Ebi, M., Kataoka, H., Higashiyama, S. & Joh, T.

TGFβ induces EGFR transactivation and HB-EGF

C-terminal fragment nuclear translocation through

ADAM17 activation in gastric cancer cells.

Gastroenterology140 (Suppl. 1), S629 (2011).

94. Kang, M. H. et al. Inhibition of PI3 kinase/Akt

pathway is required for BMP2-induced EMT and

invasion. Oncol. Rep.22, 525–534 (2009).

95. Takahashi, Y. et al. Significance of vessel count

and vascular endothelial growth factor and its

receptor (KDR) in intestinal-type gastric cancer.

Clin. Cancer Res.2, 1679–1684 (1996).

96. Sharma, V. K., Vasudeva, R. & Howden, C. W.

Changing demographics of gastric cancer: 15 year

experience. Gastroenterology114 (Suppl. 1), A40

(1998).

97. Takaishi, S. et al. Identification of gastric cancer

stem cells using the cell surface marker CD44.

Stem Cells27, 1006–1020 (2009).

98. Ishimoto, T. et al. CD44 variant regulates redox

status in cancer cells by stabilizing the xCT

subunit of system xc(-) and thereby promotes

tumor growth. Cancer Cell19, 387–400 (2011).

99. Fornaro, L. et al. Anti-HER agents in gastric cancer:

from bench to bedside. Nat. Rev. Gastroenterol.

Hepatol.8, 369–383 (2011).

100. Schlessinger, J. Common and distinct elements in

cellular signaling via EGF and FGF receptors.

Science306, 1506–1507 (2004).

101. Dhanasekaran, D. N. & Johnson, G. L. MAPKs:

function, regulation, role in cancer and therapeutic

targeting. Oncogene26, 3097–3099 (2007).

102. Lordick, F. et al. HER2 status of advanced gastric

cancer is similar in Europe and Asia

[abstract 253]. Ann. Oncol.18 (Suppl. 7),

vii95–vii96 (2007).

103. Zhang, X. L. et al. Comparative study on

overexpression of HER2/neu and HER3 in gastric

cancer. World J. Surg.33, 2112–2118 (2009).

104. Terashima, M. et al. Impact of expression of

human epidermal growth factor receptors EGFR

and ERBB2 on survival in stage II/III gastric

cancer. Clin. Cancer Res.18, 5992–6000 (2012).

105. Browne, B. C. et al. Inhibition of IGF1R activity

enhances response to trastuzumab in

HER-2-positive breast cancer cells. Ann. Oncol.22,

68–73 (2011).

106. Wilkinson, N. W. et al. Epidermal growth factor

receptor expression correlates with histologic

grade in resected esophageal adenocarcinoma.

J. Gastrointest. Surg.8, 448–453 (2004).

107. Isinger-Ekstrand, A. et al. Genetic profiles of

gastroesophageal cancer: combined analysis

using expression array and tiling array-comparative

genomic hybridization. Cancer Genet. Cytogenet.

200, 120–126 (2010).

108. Moutinho, C. et al. Epidermal growth factor

receptor structural alterations in gastric cancer.

BMC Cancer8, 10 (2008).

109. Liu, Z. M. et al. Epidermal growth factor receptor

mutation in gastric cancer. Pathology43,

234–238 (2011).

110. Pinto, C. et al. Phase II study of cetuximab in

combination with cisplatin and docetaxel in

patients with untreated advanced gastric or

gastro-oesophageal junction adenocarcinoma

(DOCETUX study). Br. J. Cancer101, 1261–1268

(2009).

111. Jaiswal, B. S. et al. Oncogenic ERBB3 mutations in

human cancers. Cancer Cell23, 603–617 (2013).

112. Migliore, C. & Giordano, S. Molecular cancer

therapy: can our expectation be MET?

Eur. J. Cancer44, 641–651 (2008).

113. El-Rifai, W. & Powell, S. M. Molecular biology

of gastric cancer. Semin. Radiat. Oncol. 12,

128–140 (2002).

114. Smolen, G. A. et al. Amplification of MET may

identify a subset of cancers with extreme

sensitivity to the selective tyrosine kinase inhibitor

PHA-665752. Proc. Natl Acad. Sci. USA103,

2316–2321 (2006).

115. Janjigian, Y. Y. et al. MET expression and

amplification in patients with localized gastric

cancer. Cancer Epidemiol. Biomarkers Prev.20,

1021–1027 (2011).

116. Lee, J. et al. Impact of MET amplification on

gastric cancer: possible roles as a novel

prognostic marker and a potential therapeutic

target. Oncol. Rep.25, 1517–1524 (2011).

117. Guo, A. et al. Signaling networks assembled by

oncogenic EGFR and c-Met. Proc. Natl Acad.

Sci. USA 105, 692–697 (2008).

118. Arteaga, C. L. HER3 and mutant EGFR meet MET.

Nat. Med.13, 675–677 (2007).

119. Corso, S., Comoglio, P. M. & Giordano, S. Cancer

therapy: can the challenge be MET? Trends Mol.

Med.11, 284–292 (2005).

REVIEWS

120. Corso, S. et al. Activation of HER family members in gastric carcinoma cells mediates resistance to

MET inhibition. Mol. Cancer9, 121 (2010).

121. Bachleitner-Hofmann, T. et al. HER kinase

activation confers resistance to MET tyrosine

kinase inhibition in MET oncogene-addicted

gastric cancer cells. Mol. Cancer Ther. 7,

3499–3508 (2008).

122. Byun, D. S. et al. Frequent monoallelic deletion of PTEN and its reciprocal association with PIK3CA

amplification in gastric carcinoma. Int. J. Cancer

104, 318–327 (2003).

123. Oki, E. et al. Impact of PTEN/AKT/PI3K signal pathway on the chemotherapy for gastric cancer

[abstract 4034]. J. Clin. Oncol. 24 (Suppl. 18),

187s (2006).

124. Garrett, J. T. et al. Transcriptional and

posttranslational up-regulation of HER3 (ErbB3)

compensates for inhibition of the HER2 tyrosine

kinase. Proc. Natl Acad. Sci. USA 108,

5021–5026 (2011).

125. Shin, E. Y. et al. Up-regulation and co-expression of fibroblast growth factor receptors in human gastric cancer. J. Cancer Res. Clin. Oncol. 126, 519–528

(2000).

126. Wang, K. et al. Exome sequencing identifies frequent mutation of ARID1A in molecular

subtypes of gastric cancer. Nat. Genet.43,

1219–1223 (2011).

127. Song, S. & Ajani, J. A. The role of microRNAs in cancers of the upper gastrointestinal tract.

Nat. Rev. Gastroenterol. Hepatol.10, 109–118

(2013).

128. Li, T. et al. MicroRNA-296-5p increases

proliferation in gastric cancer through repression

of Caudal-related homeobox 1. Oncogene http://

https://www.doczj.com/doc/8d14967670.html,/10.1038/onc.2012.637.

129. Wang, M. et al. Overexpressed miR-301a

promotes cell proliferation and invasion by

targeting RUNX3 in gastric cancer. J. Gastroenterol.

https://www.doczj.com/doc/8d14967670.html,/10.1007/s00535-012-0733–6. 130. Wang, M. et al. miR-17-5p/20a are important markers for gastric cancer and murine double

minute 2 participates in their functional

regulation. Eur. J. Cancer 49, 2010–2021 (2013). 131. Wu, X. et al. MicroRNA expression signatures during malignant progression from Barrett’s

esophagus to esophageal adenocarcinoma.

Cancer Prev. Res. (Phila.)6, 196–205 (2013). 132. Zhou, C. et al. microRNA-372 maintains oncogene characteristics by targeting TNFAIP1 and affects

NFκB signaling in human gastric carcinoma cells.

Int. J. Oncol.42, 635–642 (2013).

133. Zhi, Q. et al. Oncogenic miR-544 is an important molecular target in gastric cancer. Anticancer

Agents Med. Chem.13, 270–275 (2013).

134. Wang, Y. Y. et al. Clinicopathologic significance of miR-10b expression in gastric carcinoma.

Hum. Pathol. 44, 1278–1285 (2013).

135. Deng, H. et al. MicroRNA-195 and microRNA-378 mediate tumor growth suppression by epigenetical regulation in gastric cancer. Gene 518, 351–359

(2013).

136. Wen, D. et al. miR-133b acts as a tumor

suppressor and negatively regulates FGFR1 in

gastric cancer. Tumour Biol. 34, 793–803(2013). 137. Saito, Y. et al. The tumor suppressor

microRNA-29c is downregulated and restored by

celecoxib in human gastric cancer cells.

Int. J. Cancer 132, 1751–1760(2012).

138. Cao, W. et al. Expression and regulatory function of miRNA-34a in targeting survivin in gastric

cancer cells. Tumour Biol. 34, 963–971(2012). 139. Zheng, L. et al. miRNA-145 targets v-ets

erythroblastosis virus E26 oncogene homolog 1 to suppress the invasion, metastasis, and

angiogenesis of gastric cancer cells. Mol. Cancer

Res. 11, 182–193 (2013).140. Liu, K. et al. Decreased expression of microRNA

let-7i and its association with chemotherapeutic

response in human gastric cancer. World J. Surg.

Oncol.10, 225 (2012).

141. Akiyoshi, S. et al. Clinical significance of

miR-144-ZFX axis in disseminated tumour cells in

bone marrow in gastric cancer cases.

Br. J. Cancer107, 1345–1353 (2012).

142. He, X. P. et al. Downregulation of miR-101 in

gastric cancer correlates with cyclooxygenase-2

overexpression and tumor growth. FEBS J. 279,

4201–4212 (2012).

143. Crone, S. G. et al. microRNA-146a inhibits G.

protein-coupled receptor-mediated activation of

NF-κB by targeting CARD10 and COPS8 in gastric

cancer. Mol. Cancer11, 71 (2012).

144. Ichikawa, D., Komatsu, S., Konishi, H. & Otsuji, E.

Circulating microRNA in digestive tract cancers.

Gastroenterology142, 1074–1078.e1 (2012).

145. Cui, L. et al. Gastric juice microRNAs as potential

biomarkers for the screening of gastric cancer.

Cancer 119, 1618–1626 (2013).

146. Zhang, X. et al. Gastric juice microRNA-421 is a

new biomarker for screening gastric cancer.

Tumour Biol.33, 2349–2355 (2012).

147. Cai, H. et al. Plasma microRNAs serve as novel

potential biomarkers for early detection of gastric

cancer. Med. Oncol.30, 452 (2013).

148. Li, C. et al. MiRNA-199a-3p in plasma as a

potential diagnostic biomarker for gastric cancer.

Ann. Surg. Oncol. https://www.doczj.com/doc/8d14967670.html,/10.1245/

s10434-012-2600–3.

149. Wang, M. et al. Circulating miR-17-5p and

miR-20a: molecular markers for gastric cancer.

Mol. Med. Report5, 1514–1520 (2012).

150. Valladares-Ayerbes, M. et al. Circulating

miR-200c as a diagnostic and prognostic

biomarker for gastric cancer. J. Transl. Med.10,

186 (2012).

151. Stern, H. M. Improving treatment of HER2-positive

cancers: opportunities and challenges. Sci. Transl.

Med.4, 127rv2 (2012).

152. Nahta, R. Molecular mechanisms of trastuzumab-

based treatment in HER2-overexpressing breast

cancer. ISRN Oncol. 2012, 428062 (2012).

153. Kim, J. W. et al. The growth inhibitory effect of

lapatinib, a dual inhibitor of EGFR and HER2

tyrosine kinase, in gastric cancer cell lines.

Cancer Lett.272, 296–306 (2008).

154. Lu, Y., Zi, X. & Pollak, M. Molecular mechanisms

underlying IGF-I-induced attenuation of the growth-

inhibitory activity of trastuzumab (Herceptin) on

SKBR3 breast cancer cells. Int. J. Cancer 108,

334–341 (2004).

155. Chen, C. T. et al. MET activation mediates

resistance to lapatinib inhibition of HER2-

amplified gastric cancer cells. Mol. Cancer Ther.

11, 660–669 (2012).

156. Shattuck, D. L., Miller, J. K., Carraway, K. L.

& Sweeney, C. Met receptor contributes to

trastuzumab resistance of HER2-overexpressing

breast cancer cells. Cancer Res.68, 1471–1477

(2008).

157. Molina, M. A. et al. Trastuzumab (herceptin), a

humanized anti-HER2 receptor monoclonal

antibody, inhibits basal and activated HER2

ectodomain cleavage in breast cancer cells.

Cancer Res.61, 4744–4749 (2001).

158. Nagy, P. et al. Decreased accessibility and lack of

activation of ErbB2 in JIMT-1, a herceptin-

resistant, MUC4-Expressing breast cancer cell

line. Cancer Res.65, 473–482 (2005).

159. Gong, S. J., Jin, C. J., Rha, S. Y. & Chung, H. C.

Growth inhibitory effects of trastuzumab and

chemotherapeutic drugs in gastric cancer cell

lines. Cancer Lett.214, 215–224 (2004).

160. Arcaro, A. & Guerreiro, A. S. The phosphoinositide

3-kinase pathway in human cancer: genetic

alterations and therapeutic implications.

Curr. Genomics8, 271–306 (2007).

161. Engelman, J. A. et al. Effective use of PI3K and

MEK inhibitors to treat mutant Kras G12D and

PIK3CA H1047R murine lung cancers. Nat. Med.

14, 1351–1356 (2008).

162. Eichhorn, P. J. et al. Phosphatidylinositol 3-kinase

hyperactivation results in lapatinib resistance that

is reversed by the mTOR/phosphatidylinositol

3-kinase inhibitor NVP-BEZ235. Cancer Res.68,

9221–9230 (2008).

163. Korkaya, H. et al. Activation of an IL6 inflammatory

loop mediates trastuzumab resistance in HER2+

breast cancer by expanding the cancer stem cell

population. Mol. Cell47, 570–584 (2012).

164. Shah, M. A. & Ajani, J. A. Gastric cancer—an

enigmatic and heterogeneous disease. JAMA303,

1753–1754 (2010).

165. Bickenbach, K. & Strong, V. E. Comparisons of

gastric cancer treatments: East vs West. J. Gastric

Cancer12, 55–62 (2012).

166. Tamura, S., Takeno, A. & Miki, H. Lymph node

dissection in curative gastrectomy for advanced

gastric cancer. Int. J. Surg. Oncol. 2011, 748745

(2011).

167. Macdonald, J. S. et al. Chemoradiotherapy after

surgery compared with surgery alone for

adenocarcinoma of the stomach or

gastroesophageal junction. N. Engl. J. Med.345,

725–730 (2001).

168. Sakuramoto, S. et al. Adjuvant chemotherapy for

gastric cancer with S-1, an oral fluoropyrimidine.

N. Engl. J. Med.357, 1810–1820 (2007).

169. Bang, Y. J. et al. Adjuvant capecitabine and

oxaliplatin for gastric cancer after D2 gastrectomy

(CLASSIC): a phase 3 open-label, randomised

controlled trial. Lancet379, 315–321 (2012).

170. Ajani, J. A. et al. Gastric cancer. J. Natl Compr.

Canc. Netw.8, 378–409 (2010).

171. Lim, L., Michael, M., Mann, G. B. & Leong, T.

Adjuvant therapy in gastric cancer. J. Clin. Oncol.

23, 6220–6232 (2005).

172. Ychou, M. et al. Perioperative chemotherapy

compared with surgery alone for resectable

gastroesophageal adenocarcinoma: an FNCLCC

and FFCD multicenter phase III trial. J. Clin. Oncol.

29, 1715–1721 (2011).

173. Paoletti, X. et al. Benefit of adjuvant chemotherapy

for resectable gastric cancer: a meta-analysis.

JAMA303, 1729–1737 (2010).

174. Cunningham, D. et al. Perioperative chemotherapy

versus surgery alone for resectable

gastroesophageal cancer. N. Engl. J. Med.355,

11–20 (2006).

175. Smalley, S. R. et al. Updated analysis of SWOG-

directed intergroup study 0116: a phase III trial of

adjuvant radiochemotherapy versus observation

after curative gastric cancer resection. J. Clin.

Oncol.30, 2327–2333 (2012).

176. Fuchs, C. S. et al. Postoperative adjuvant

chemoradiation for gastric or gastroesophageal

junction (GEJ) adenocarcinoma using epirubicin,

cisplatin, and infusional (CI) 5-FU (ECF) before and

after CI 5-FU and radiotherapy (CRT) compared

with bolus 5-FU/L V before and after CRT:

Intergroup trial CALGB 80101 [abstract 4003].

J. Clin. Oncol.29 (Suppl.), 4003 (2011).

177. Lee, J. et al. Phase III trial comparing capecitabine

plus cisplatin versus capecitabine plus cisplatin

with concurrent capecitabine radiotherapy in

completely resected gastric cancer with D2 lymph

node dissection: the ARTIST trial. J. Clin. Oncol.30,

268–273 (2012).

178. Sasako, M. et al. Five-year outcomes of a

randomized phase III trial comparing adjuvant

chemotherapy with S-1 versus surgery alone in

stage II or III gastric cancer. J. Clin. Oncol.29,

4387–4393 (2011).

REVIEWS

179. Dikken, J. L. et al. Neo-adjuvant chemotherapy followed by surgery and chemotherapy or by

surgery and chemoradiotherapy for patients with

resectable gastric cancer (CRITICS). BMC Cancer

11, 329 (2011).

180. Van Cutsem, E. et al. Phase III study of docetaxel and cisplatin plus fluorouracil compared with

cisplatin and fluorouracil as first-line therapy for

advanced gastric cancer: a report of the V325

Study Group. J. Clin. Oncol.24, 4991–4997

(2006).

181. Koizumi, W. et al. S-1 plus cisplatin versus S-1 alone for first-line treatment of advanced gastric

cancer (SPIRITS trial): a phase III trial.

Lancet Oncol.9, 215–221 (2008).

182. Bang, Y. J. et al. Trastuzumab in combination with chemotherapy versus chemotherapy alone for

treatment of HER2-positive advanced gastric or

gastro-oesophageal junction cancer (ToGA): a

phase 3, open-label, randomised controlled trial.

Lancet376, 687–697 (2010).

183. FD A. Trastuzumab. Office of Medical Products and Tobacco [online], https://www.doczj.com/doc/8d14967670.html,/

AboutFDA/CentersOffices/

OfficeofMedicalProductsandTobacco/CDER/

ucm230418.htm (2010).

184. Lordick, F. et al. Capecitabine and cisplatin with or without cetuximab for patients with previously

untreated advanced gastric cancer (EXPAND):

a randomised, open-label phase 3 trial.

Lancet Oncol.14, 490–499(2013).

185. Waddell, T. et al. Epirubicin, oxaliplatin, and capecitabine with or without panitumumab for

patients with previously untreated advanced

oesophagogastric cancer (REAL3): a randomised,

open-label phase 3 trial. Lancet Oncol. 14,

481–489(2013).

186. Ohtsu, A. et al. Bevacizumab in combination with chemotherapy as first-line therapy in advanced

gastric cancer: a randomized, double-blind,

placebo-controlled phase III study. J. Clin. Oncol.

29, 3968–3976 (2011).

187. Van Cutsem, E. et al. Bevacizumab in combination with chemotherapy as first-line therapy in

advanced gastric cancer: a biomarker evaluation

from the AVAGAST randomized phase III trial.

J. Clin. Oncol.30, 2119–2127 (2012).

188. Ajani, J. A. et al. Multicenter phase III comparison of cisplatin/S-1 with cisplatin/infusional

fluorouracil in advanced gastric or

gastroesophageal adenocarcinoma study: the

FLAGS trial. J. Clin. Oncol.28, 1547–1553 (2010). 189. Thuss-Patience, P. C. et al. Survival advantage for irinotecan versus best supportive care as second-

line chemotherapy in gastric cancer--a randomised phase III study of the Arbeitsgemeinschaft

Internistische Onkologie (AIO). Eur. J. Cancer47,

2306–2314 (2011).

190. US National Library of Medicine. https://www.doczj.com/doc/8d14967670.html, [online], https://www.doczj.com/doc/8d14967670.html,/show/

NCT00879333 (2013).

191. Everolimus for advanced gastric cancer: results of the randomized, double-blind, phase III GRANITE-1

study. J. Clin. Oncol. (in press).

192. US National Library of Medicine. https://www.doczj.com/doc/8d14967670.html, [online], https://www.doczj.com/doc/8d14967670.html,/show/

NCT00917384/ (2013).

193. Fuchs, C. S. et al. REGARD: a phase III,

randomized, double-blind trial of ramucirumab and

best supportive care (BSC) versus placebo and

BSC in the treatment of metastatic gastric or

gastroesophageal junction (GEJ) adenocarcinoma

following disease progression on first-line

platinum- and/or fluoropyrimidine-containing

combination therapy [abstract LBA5]. J. Clin. Oncol.

30(Suppl. 34), LBA5 (2012).

194. Ford, H. et al. Cougar-02: a randomized phase III study of docetaxel versus active symptom control

in advanced esophagogastric adenocarcinoma

[abstract LBA4]. J. Clin. Oncol. 30(Suppl. 34),

LBA4 (2012).

195. Bang, Y.-J. et al. A randomized, open-label,

phase III study of lapatinib in combination with

weekly paclitaxel versus weekly paclitaxel alone

in the second-line treatment of HER2 amplified

advanced gastric cancer (AGC) in Asian

population: Tytan study [abstract 11]. J. Clin.

Oncol. 30(Suppl. 34), 11 (2012).

196. Lennerz, J. K. et al. MET amplification identifies a

small and aggressive subgroup of

esophagogastric adenocarcinoma with evidence

of responsiveness to crizotinib. J. Clin. Oncol.29,

4803–4810 (2011).

197. Oliner, K. S. et al. Evaluation of MET pathway

biomarkers in a phase II study of rilotumumab

(R, AMG 102) or placebo (P) in combination with

epirubicin, cisplatin, and capecitabine (ECX) in

patients (pts) with locally advanced or metastatic

gastric (G) or esophagogastric junction (EGJ)

cancer [abstract 4005]. J. Clin. Oncol. 30(Suppl.),

4005 (2012).

198. Hwang, J. Y. et al. Recapitulation of previous

genome-wide association studies with two

distinct pathophysiological entities of gastric

cancer in the Korean population. J. Hum. Genet.

58, 233–235(2013).

199. Saeki, N. et al. A functional single nucleotide

polymorphism in mucin 1, at chromosome

1q22, determines susceptibility to diffuse-type

gastric cancer. Gastroenterology140, 892–902

(2011).

200. Gu, H. et al. Replication study of PLCE1 and

C20orf54 polymorphism and risk of esophageal

cancer in a Chinese population. Mol. Biol. Rep.

39, 9105–9111 (2012).

201. Mammano, E. et al. Epidermal growth factor

receptor (EGFR): mutational and protein

expression analysis in gastric cancer.

Anticancer Res.26, 3547–3550 (2006).

202. Hong, S. P. et al. Overexpression of Notch 1

signaling associates with the tumorigenesis of

gastric adenorna and intestinal type of gastric

cancer. Gastroenterology132, A617 (2007).

203. Yu, G. Z. et al. Overexpression of phosphorylated

mammalian target of rapamycin predicts lymph

node metastasis and prognosis of Chinese

patients with gastric cancer. Clin. Cancer Res. 15,

1821–1829 (2009).

204. Lang, S. A. et al. Mammalian target of rapamycin

is activated in human gastric cancer and serves

as a target for therapy in an experimental model.

Int. J. Cancer 120, 1803–1810 (2007).

205. Zhou, Y. N. et al. Clinicopathological significance

of E-cadherin, VEGF, and MMPs in gastric cancer.

Tumor Biol.31, 549–558 (2010).

206. Kitoh, T. et al. Increased expression of matrix

metalloproteinase-7 in invasive early gastric

cancer. J. Gastroenterol.39, 434–440 (2004).

207. Fukaya, M. et al. Hedgehog signal activation in

gastric pit cell and in diffuse-type gastric cancer.

Gastroenterology131, 14–29 (2006).

208. Hattori, Y. et al. Immunohistochemical detection

of K-sam protein in stomach cancer. Clin. Cancer

Res. 2, 1373–1381 (1996).

209. Pereira, P. S. et al. E-cadherin missense

mutations, associated with hereditary diffuse

gastric cancer (HDGC) syndrome, display distinct

invasive behaviors and genetic interactions with

the Wnt and Notch pathways in Drosophila

epithelia. Hum. Mol. Genet. 15, 1704–1712

(2006).

210. Kobayashi, M., Kawashima, A., Mai, M. & Ooi, A.

Analysis of chromosome 17p13 (p53 locus)

alterations in gastric carcinoma cells by dual-

color fluorescence in situ hybridization.

Am. J. Pathol. 149, 1575–1584 (1996).

211. Guo, C. Y., Xu, X. F., Wu, L. Y. & Liu, S. F.

PCR-SSCP-DNA sequencing method in detecting

PTEN gene mutation and its significance in human

gastric cancer. World J. Gastroenterol.14,

3804–3811 (2008).

212. Li, H. et al. Insulin-like growth factor-I receptor

blockade reduces tumor angiogenesis and

enhances the effects of bevacizumab for a human

gastric cancer cell line, MKN45. Cancer 117,

3135–3147 (2011).

213. Cepero, V. et al. MET and KRAS gene amplification

mediates acquired resistance to MET tyrosine

kinase inhibitors. Cancer Res.70, 7580–7590

(2010).

214. Fuchs, C. et al. Postoperative adjuvant

chemoradiation for gastric or gastroesophageal

adenocarcinoma using epirubicin, cisplatin, and

infusional (CI) 5-FU (ECF) before and after CI 5-FU

and radiotherapy (RT): Interim toxicity results from

Intergroup trial CALGB 80101 [abtract 4003].

J. Clin. Oncol. 29 (Suppl.),4003 (2011).

215. Tsuburaya, A. et al. SAMIT: Preliminary safety data

from a 2 × 2 factorial randomized phase III trial to

investigate weekly paclitaxel (PTX) followed by oral

fluoropyrimidines (FPs) versus FPs alone as

adjuvant chemotherapy in patients (pts) with

gastric cancer [abstract 4017]. J. Clin. Oncol.

29 (Suppl.),4017 (2011).

216. Leong, T. L. et al. TOPGEAR: An international

randomized phase III trial of preoperative

chemoradiotherapy versus preoperative

chemotherapy for resectable gastric cancer

(AGITG/TROG/EORTC/NCIC CTG)

[abstract TPS4141]. J. Clin. Oncol.30 (Suppl.),

TPS4141 (2012).

217. Lordick, F. et al. Capecitabine and cisplatin with or

without cetuximab for patients with previously

untreated advanced gastric cancer (EXPAND):

a randomised, open-label phase 3 trial.

Lancet Oncol.14, 490–499 (2013).

218. US National Library of Medicine. https://www.doczj.com/doc/8d14967670.html,

[online], https://www.doczj.com/doc/8d14967670.html,/show/

NCT00450203 (2013).

219. Waddell, T. S. et al. A randomized, multicenter trial

of epirubicin, oxaliplatin, and capecitabine (EOC)

plus panitumumab in advanced esophagogastric

cancer (REAL3) [abstract LBA4000]. J. Clin. Oncol.

30 (Suppl.),LBA4000 (2012).

220. US National Library of Medicine. https://www.doczj.com/doc/8d14967670.html,

[online], https://www.doczj.com/doc/8d14967670.html,/show/

NCT00680901 (2013).

221. Kang, J. H. et al. Salvage chemotherapy for

pretreated gastric cancer: a randomized phase III

trial comparing chemotherapy plus best

supportive care with best supportive care alone.

J. Clin. Oncol.30, 1513–1518 (2012).

222. US National Library of Medicine. https://www.doczj.com/doc/8d14967670.html,

[online], https://www.doczj.com/doc/8d14967670.html,/show/

NCT00978549 (2009).

Acknowledgements

The authors are supported by grants from the

Caporella, Cantu, Smith, Park, Fairman, Frazier, Sultan,

Oaks, Vansteklenberg and Milrod families, by the

Kevin Fund, Schecter Private Fund and the Rivercreek

Foundation and a Multidisciplinary Research Grant from

University of Texas MD Anderson Cancer Center.

J. A. Ajani is also supported by grants RO1CA138671

and ROaCA172741 awarded by the National Cancer

Institute, Bethesda, MD, USA. The authors also thank

the editorial staff of Nature Reviews Clinical Oncology for

their editorial assistance in considerably improving the

clarity and organization of the manuscript.

Author contributions

All authors researched the data, discussed the

article’s content, wrote the manuscript and edited

it before submission.

REVIEWS

相关主题
文本预览
相关文档 最新文档