Top Banner
Tumor Microenvironment in Head and Neck Squamous Cell Carcinoma Joseph M. Curry, a John Sprandio, b David Cognetti, a Adam Luginbuhl, a Voichita Bar-ad, c Edmund Pribitkin, a and Madalina Tuluc d The tumor microenvironment (TME) of head and neck squamous cell carcinoma (HNSCC) is comprised of cancer-associated fibroblasts (CAFs), immune cells, and other supporting cells. Genetic changes in the carcinoma cells, such as alterations to TP53, NOTCH1, and specific gene expression profiles, contribute to derangements in cancer and microenvironment cells such as increased ROS, overproduction of cytokines, and epithelial to mesenchymal transition (EMT). CAFs are among the most critical elements of the TME contributing to proliferation, invasion, and metastasis. The adaptive immune response is suppressed in HNSCC through overexpression of cytokines, triggered apoptosis of T cells, and alterations in antigen processing machinery. Overexpression of critical cytokines, such as transforming growth factor-β (TGF-β), contributes to EMT, immune suppression, and evolution of CAFs. Inflammation and hypoxia are driving forces in angiogenesis and altered metabolism. HNSCC utilizes glycolytic and oxidative metabolism to fuel tumorigenesis via coupled mechanisms between cancer cell regions and cells of the TME. Increased understanding of the TME in HNSCC illustrates that the long-held notion of condemned mucosareflects a process that extends beyond the epithelial cells to the entire tissue comprised of each of these elements. Semin Oncol 41:217-234 & 2014 Elsevier Inc. S quamous cell carcinoma comprises more than 90% of cancers of the head and neck and arises from the squamous lining of the mucosal surfa- ces of the upper aerodigestive tract, including the oral cavity, pharynx, larynx, and sinonasal tract. Head and neck squamous cell carcinoma (HNSCC) is the sixth most common cancer worldwide, and only 50%60% of patients are alive at 5 years after diagnosis. 1,2 Treatment can be quite morbid and result in significant functional as well as aesthetic deficits, such as impair- ment of speech and swallowing and facial deformity. Treatment failure and locoregional recurrence are common and occur in up to 30% of patients and account for the majority of deaths. 3 The high rate of local recurrence produced the long-held notion of condemned mucosaor field cancerizationinitially described in the 1950s. 4 This concept underscores not only the difficulty in treating HNSCC but also denotes the complexity of the molecular conditions under which HNSCC develops and recurs. It is clear that the notion of the condemned mucosa reflects a condemned tissuecomposed of the cancerous cells, adjacent epithelial, stromal, and immune cells and their surrounding matrix. Together these elements comprise the tumor microenvironment (TME). In fact, this shift in thought from the concept that cancer is derived from a single cell type, to a disease occurring in a complex tissue, has led some investigators to suggest that the very definition of carcinoma be changed. 5 Tumorigenesis requires multiple elements outlined by Hanahan and Weinberg: (1) limitless replicative potential, (2) self-sufficiency in growth signals, (3) insen- sitivity to anti-growth signals, (4) ability to evade apoptosis, (5) increased angiogenesis, and (6) invasion and metastasis. 6 Knowledge of the mechanisms through which the cancer cells use the TME to execute these processes continues to evolve. 7,8 There is great interest in the downstream paracrine interactions with 0093-7754 & 2014 Elsevier Inc. http://dx.doi.org/10.1053/j.seminoncol.2014.03.003 Conflicts of interest: none. a Department of Otolaryngology Head and Neck Surgery, Thomas Jefferson University, Philadelphia, PA. b Department of Medical Oncology, Thomas Jefferson University, Philadelphia, PA. c Department of Radiation Oncology, Thomas Jefferson University, Philadelphia, PA. d Department of Pathology, Thomas Jefferson University, Philadelphia, PA. Address correspondence to Joseph M. Curry, MD, Department of Otolaryngology Head and Neck Surgery, Thomas Jefferson Uni- versity, Philadelphia, PA 19107. E-mail: joseph.curry@jefferson. edu Seminars in Oncology, Vol 41, No 2, April 2014, pp 217-234 217 Open access under CC BY-NC-ND license. Open access under CC BY-NC-ND license.
18

Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

Jun 24, 2018

Download

Documents

trantu
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

Tumor Microenvironment in Head and Neck SquamousCell Carcinoma

Joseph M. Curry,a John Sprandio,b David Cognetti,a Adam Luginbuhl,a Voichita Bar-ad,c

Edmund Pribitkin,a and Madalina Tulucd

The tumor microenvironment (TME) of head and neck squamous cell carcinoma (HNSCC) is

comprised of cancer-aGenetic changes in the

expression profiles, co

increased ROS, overprCAFs are among the m

metastasis. The adaptiv

cytokines, triggered aOverexpression of crit

to EMT, immune supp

forces in angiogenesmetabolism to fuel tu

cells of the TME. Incr

notion of “condemnedentire tissue comprise

Semin Oncol 41:217-2

0093-7754& 2014 Elshttp://dx.doi

Conflicts o

aDepartmenJefferson

bDepartmePhiladelp

cDepartmenPhiladelp

dDepartmePA.

Address coOtolarynversity,edu

Seminars

ssociated fibroblasts (CAFs), immune cells, and other supporting cells.carcinoma cells, such as alterations to TP53, NOTCH1, and specific gene

ntribute to derangements in cancer and microenvironment cells such as

oduction of cytokines, and epithelial to mesenchymal transition (EMT).ost critical elements of the TME contributing to proliferation, invasion, and

e immune response is suppressed in HNSCC through overexpression of

poptosis of T cells, and alterations in antigen processing machinery.ical cytokines, such as transforming growth factor-β (TGF-β), contributesression, and evolution of CAFs. Inflammation and hypoxia are driving

is and altered metabolism. HNSCC utilizes glycolytic and oxidativemorigenesis via coupled mechanisms between cancer cell regions and

eased understanding of the TME in HNSCC illustrates that the long-held

mucosa” reflects a process that extends beyond the epithelial cells to thed of each of these elements.

34 & 2014 Elsevier Inc. Open access under CC BY-NC-ND license.

Squamous cell carcinoma comprises more than90% of cancers of the head and neck and arises

from the squamous lining of the mucosal surfa-

ces of the upper aerodigestive tract, including the oralcavity, pharynx, larynx, and sinonasal tract. Head and

neck squamous cell carcinoma (HNSCC) is the sixth

most common cancer worldwide, and only 50%–60%of patients are alive at 5 years after diagnosis.1,2

Treatment can be quite morbid and result in significant

functional as well as aesthetic deficits, such as impair-ment of speech and swallowing and facial deformity.

evier Inc..org/10.1053/j.seminoncol.2014.03.003

f interest: none.

t of Otolaryngology Head and Neck Surgery, ThomasUniversity, Philadelphia, PA.

nt of Medical Oncology, Thomas Jefferson University,hia, PA.t of Radiation Oncology, Thomas Jefferson University,hia, PA.nt of Pathology, Thomas Jefferson University, Philadelphia,

rrespondence to Joseph M. Curry, MD, Department ofgology Head and Neck Surgery, Thomas Jefferson Uni-Philadelphia, PA 19107. E-mail: joseph.curry@jefferson.

in Oncology, Vol 41, No 2, April 2014, pp 217-234

Open access under CC BY-NC-ND license.

Treatment failure and locoregional recurrence arecommon and occur in up to 30% of patients and

account for the majority of deaths.3 The high rate of

local recurrence produced the long-held notion of“condemned mucosa” or “field cancerization” initially

described in the 1950s.4 This concept underscores not

only the difficulty in treating HNSCC but also denotesthe complexity of the molecular conditions under

which HNSCC develops and recurs. It is clear that

the notion of the condemned mucosa reflects a“condemned tissue” composed of the cancerous cells,

adjacent epithelial, stromal, and immune cells and their

surrounding matrix. Together these elements comprisethe tumor microenvironment (TME). In fact, this shift

in thought from the concept that cancer is derived

from a single cell type, to a disease occurring in acomplex tissue, has led some investigators to suggest

that the very definition of carcinoma be changed.5

Tumorigenesis requires multiple elements outlinedby Hanahan and Weinberg: (1) limitless replicative

potential, (2) self-sufficiency in growth signals, (3) insen-

sitivity to anti-growth signals, (4) ability to evadeapoptosis, (5) increased angiogenesis, and (6) invasion

and metastasis.6 Knowledge of the mechanisms

through which the cancer cells use the TME to executethese processes continues to evolve.7,8 There is great

interest in the downstream paracrine interactions with

217

Page 2: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

J.M. Curry et al218

the stroma, immune interactions, and metabolic

changes and the role each plays in tumorigenesis.HNSCC is genetically heterogeneous, but a num-

ber of pathways have been found to be commonly

involved; the impact of several critical abnormalitieson the TME is highlighted below. The cellular

elements of the TME often coevolve with the tumor.

Stromal fibroblasts, T cells, macrophages, and othercell types develop abnormal phenotypes in a disor-

ganized response to the cancer (Figure 1). These non-

cancerous cells provide many of the paracrine signalsnecessary to turn on the pleotrophic abilities of

cancer cells.9 For example, fibroblasts become

cancer-associated fibroblasts and secrete factors suchas matrix metalloproteins (MMPs), contributing to

tumor invasiveness. Furthermore, as the chronic

inflammation of the TME remains unresolved, alter-ations in adaptive immune response such as apopto-

sis of cytotoxic T cells and activation of suppressor T

cells occurs.10 Additionally, tumors reprogram theirsurroundings creating a metabolically fertile environ-

ment to meet their high energy and anabolic require-

ments. This process was aptly described by Paget asthe “seed and soil” hypothesis.11 Fundamental tumor–non-tumor microenvironmental interactions such as

these represent potential points of intervention fortherapeutic strategies. Many critical targets, such as

nuclear factor-κB (NF-κB), hypoxia-inducible factor

(HIF)-1α, and vascular endothelial growth factor(VEGF), have been, and continue to be, explored as

therapeutic targets in the TME12–14 (Table 1).

IMPACT OF GENETIC AND EPIGENETICCHANGES OF THE EPITHELIUM ON THE TME

The initiating genetic alterations in the epithelial

cells of HNSCC are primarily the result of thecarcinogenic properties of tobacco and alcohol, and

in the oropharynx, oncogenic strains of the human

papilloma virus (HPV). Classically, HNSCC has beenthought of as a disease caused by tobacco and

alcohol, yet tobacco-related cancers are decreasing

Figure 1. Select elements and interaciotns of the TME. Theperivascular niche that commonly contain cancer stem cells ancompartment contains tumor cells that are glycolytic and less proDDR is shown between the leading edge tumor edge and normaadjacent to normal stroma and fibroblasts (yellow). CAFs expreprotein inducer (EMMPRIN) on cancer cells to activate MMP2activate TGF-β. CAFs and tumor cells produce elements like VEGinteractions between regulatory T cells (pink), cytotoxic T cells (r(blue/green) are shown. TGF-β and IL-10 produced by TAMs anMIF that recruits neutrophils. Regulatory T cells induce toleracytotoxic T cells in utlitzed by cancer cells to induce apoptosincrease angiogenesis and invasion by production of MMP-9, Vorange) are recruited to the TME by GM-CSF produced by cancTGF-β.

in incidence.15 Over the past several decades, onco-

genic strains of HPV have become apparent as anetiology for oropharyngeal squamous cell carcinoma

(OPSCC). HPV-related OPSCC accounts for up to 60%

of cases of oropharyngeal cancer in some regions;this has resulted in an increased incidence among

younger nonsmokers, and has been equated to an

epidemic by some investigators.16 Currently, this isthe second most common malignancy caused by

HPV.17 OPSCC is caused primarily by HPV16 (but

also HPV18, HPV31, and others), via the E6 and E7mechanisms established in cervical cancer.18

The most widely identified mutation in non–HPV-related HNSCC occurs in the tumor-suppressor geneTP53. This has been identified to occur in approx-

imately 50% of HNSCCs and is likely an early event, as

it is commonly found in premalignant lesions aswell.19–21 Mutations also have been shown to corre-

late with aggression and poor outcomes; for exam-

ple, p53 mutations have been found in 95% ofradioresistant tumors.22–24 Histologically negative

margins with p53 mutations have been shown to be

associated with a greater incidence of local recur-rence.25 Mutation of TP53 in tumor cells is associated

with increased migration of cancer-associated fibro-

blasts (CAFs) to the TME, while intact TP53 inhibitsmigration.26 Loss of functional p53 increases reactive

oxygen species (ROS) and reactive nitrogen species

(RNS) and may drive carcinogenesis via NF-κB andother inflammatory-mediated mechanisms. Altera-

tions in TP53 induce a DNA damage response

(DDR) in adjacent non-tumoral cells via productionof ROS. This effect was recently demonstrated in

esophageal SCC, and it increases with proximity to

and size of the primary tumor, with effects beingidentified several centimeters from the tumor.27–30

TP53 mutations also have been linked to abnormal

tumor metabolism, contributing to the Warburgeffect through increased activity of glucose trans-

porters and glycolytic enzymes furthering the pro-

duction of an acidic environment and high levels ofROS toxic to normal cells31 (Figure 1).

tumor is shown here with the leading tumor edge andd highly replicating tumor cells (blue). The more centralliferative (orange). Peritumoral epithelium demonstratingl epithelium. CAFs (purple) are shown in the tumor stromass MT-MMP that interacts with extrcellular metallomatrix. CSCs express CD144, which interacts with MMP9 toF, PGE2, and CXCL12 that trigger angiogenesis. Immuneed), M2 TAMs (green), and tumor-associated neurtrophilsd cancer cells suppress T-cell activity. TAMs also producence by cytotoxic T cells. The Fas receptor on activatedis. Tumor-associated neutrophils produce ROS, and alsoEGF, and HGF. CD34þ myeloid progenitor cells (yellow/er cells which in turn induce immunosuppression through

Page 3: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

Tumor microenvironment in HNSCC 219

NOTCH1, the second most commonly mutated

gene in HNSCC occurring in approximately 15% ofcases, functions as a tumor-suppressor.21 It encodes a

transmembrane receptor that regulates cell differ-

entiation and embryonic development.21,32 InHNSCC, it is dependent on intercellular signaling in

the TME and contributes to proliferation and invasive-

ness through the pro-inflammatory cytokine, tumornecrosis factor-α (TNF-α). This TNF-α mechanism acts

on Slug and Twist, two other important transcription

factors that act as regulators of invasion and epithelialto mesenchymal transition (EMT).9,33 Evidence also

Page 4: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

Table 1. Critical Cells of the TME in HNSCC

Cell Type Markers Secreted Factors Metabolism References

Squamous Cell CarcinomaKey Genetic Alterations: TP53,NOTCH1, EGFR, CDKN2a, STAT3,Cyclin D1, Rb

E-cadherin, cytokeratins,PD-L1, FasL

MMP 2, MMP 9, MMP 13,ROS,VEGF, CXCL1,CXCL8, PDGF, IL-8,FGF-2, TGF-β, TNF-α,IL-1, GMCSF

Tan et al,Koontongkaew S,Zhang Z et al,Smith A et al,Curry J et al, Feron O.

Central tumor compartment Glycolytic: (MCT4þ,MCT1�, TOMM20�,COX�)OXPHOS: (MCT1þ,MCT4-,TOMM20þ,COXþ)

Leading edge/invasive front,perivascular niche (proliferativecancer cells: high Ki-67)

Cancer stem cells CD33, CD144, ALDHEpithelial to mesenchymal

transitionN-cadherin, vimentin

Cancer-Associated Fibroblast α-SMA, integrin α6 HGF, CXCL12, TGF-β,MMP2, MMP9, PMF,PDGF, Type IV collagen,Col15-binding integrins,PGE2

Glycolytic: (MCT4þ,MCT1-LDH-Bþ)

Leef G, Curry J et al,Wheeler SE et al,Marsh D et al

Tumor-Infiltrating LymphocytesRegulatory T cells CD4þCD25þFoxP3þ IL-10, IL 12, TGF-β Young MR, Ferris RL

et al, Whiteside TLCytotoxic T cells CD8þ, TCR, Fas, PD-1 Perforin, granzymes,granulysin

Th2 suppressor cells CD4þ IL-4, IL-6, IL-10Myeloid progenitor cells CD34þ TGF-β

Tumor-Associated Macrophages(M2)

IL-10, TGF-β, MIF, EGF, CSF-1, MMP9, CXCL2,CXCL8,VEGF, ROS, RNS, PGEs

Lago Costa N et al,Dumitru C et al,Galdiero MR et al

Tumor-Associated Neutrophils MMP9, VEGF, HGF, elastase,ROS, PGEs

Galdiero MR et al,Dumtru C et al

Endothelial Cells Endothelins, CXCL1, CXCL8 Neiva KG et alAbbreviations: Rb, retinoblastoma gene; EGFR, epidermal growth factor receptor; CDKN2a, cyclin-dependent kinase inhibitor 2a; STAT 3, signal transducer and activator of transcription 3;

PD-L1, programmed death ligand-1; FasL, Fas ligand; MMP, matrix metalloprotein; ALDH, aldehyde dehydrogenase; ROS, reactive oxygen species; VEGF, vascular endothelial growthfactor; IL, inteleukin; FGF,fibroblast growth factor; TGF-β, transforming growth factor-β; TNF-α, tumor necrosis factor-α; PDGF, platelet-derived growth factor; GM-CSF, granulocye-macrophage colony-stimulating factor; EGF, epidermal growth factor; CSF-1, colony-stimulating factor-1; TCR, T-cell receptor; FoxP3, forked/winghead transcription factor; RNS, reactivenitrogen species; MCT, monocarboxylate transporter; PGE, protaglandin; PD-1, programmed death-1; TOMM20, translocase of outer mitochondrial membrane 20; COX, cytochrome Coxidase complex; LDH-B, lactate dehydrogenase B.

J.M.Curry

etal

220

Page 5: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

Tumor microenvironment in HNSCC 221

has suggested that its activity is mediated through

MMPs and the inflammatory transcription factor, NF-κB, among other critical mechanisms34,35 (Figure 1).

EGFR is a membrane-bound tyrosine kinase recep-

tor that binds epidermal growth factor (EGF) and TGF-α. It is the target for the effective and widely used

monoclonal antibody, cetuximab. Mutation of the

EGFR gene is only present in about 10% of cases,but gene amplification is present in about 30% of cases

and overexpression has been identified in up to 90%

of cases. Increased expression and gene copy numbercorrelate with poor prognosis.36–39 After binding

one of its ligands, EGFR triggers multiple intracellular

signaling cascades that activate cell proliferation,survival, invasion, metastasis, and angiogenesis.21,40–43

It also allows for decreased response to radiotherapy

by enhancing proliferation, DNA-repair, and hypoxicresponses within the TME.44–46 Activation triggers

increased interleukin-8 (IL-8) and VEGF production,

promoting inflammation and angiogenesis.47,48

Cyclin-dependent kinase inhibitor 2a (CDKN2a) is

an important tumor-suppressor that is mutated in

9%–12% of HNSCC patients. The function of itsproduct, p16, is to block cell cycle progression from

the G1 to S phase, and is typically upregulated under

stress conditions in the cellular microenvironmentsuch as hypoxia.49 Mutations are common but alone

likely to be insufficient to result in tumorigenesis.21

Mutational loss of function correlates with a wors-ened prognosis, while overexpression of p16 is

common in HPV-related OPSCC and correlates with

an improved prognosis.45,48,50

Signal transducer and activation of transcription

(STAT) proteins are transcription factors that are

commonly overexpressed in cancer, and in HNSCC,STAT3 has been found to be commonly mutated and

overexpressed.5 STAT3 activation is linked to numer-

ous pathways, including TGF-β, IL-6, and EGFR, andit is involved in EMT, proliferation, apoptosis, and

inflammation.51 STAT3 is also central to mainte-

nance of self-renewal in cancer stem cells (CSCs).52

The STAT/JAK pathway is one of the critical targets

of cetuximab.53 Numerous other abnormalities have

been found to be prevalent in HNSCC; Tan et al hasaddressed many of these in an excellent review.21

Given the great variety of genetic abnormalities

that have been identified in HNSCC, gene expressionprofiles may offer greater accuracy for characteriza-

tion and diagnosis than analysis of single loci. Several

groups have established profiles that can differentiateHNSCC from surrounding normal tissues. For exam-

ple, in a early study Chung et al identified four

subtypes of SCC based on gene expression profiles,each with different survival and recurrence rates.54

They identified patterns that they classified as

(1) EGFR pathway subtype, (2) mesenchymal-enriched subtype, (3) normal-epithelium–like

subtype, and (4) high-antioxidant enzyme subtype.

The EGFR group had the worst outcome. The secondsubgroup had a high fibroblast component and

demonstrated evidence of EMT. The third group

demonstrated gene expression closest to normaltonsillar epithelium and had the best outcome. The

fourth group demonstrated patterns similar to that

induced by exposure to cigarette smoking with highlevels of antioxidant enzymes being expressed.54

Clatot et al recently published a series in which they

used high-throughput reverse transcriptase polymer-ase chain reaction to create a nine-gene model with

which they were able to classify patients with 90%

accuracy. Those in a cluster with higher expressionof the chemokine CXCL12 had significantly greater

disease-free survival compared to those in a low-

expression CXCL12 cluster. Among these nine genes,a high-fold change in survival was seen in the group

comprised of CXCL12, SCL16A4 (monocarboxylate

transporter 4, MCT4), and carbonic anhydrase IX(CA9). CXCL12 is an important cytokine in HNSCC

implicated in angiogenesis and other processes.55

SCL16A4/MCT4 is a lactate transporter that has beenshown to be overexpressed in response to hypoxia.

CA9 is also upregulated by hypoxia and functions to

regulate intracellular pH.56

A number of epigenetic changes have been found

to be common to HNSCC, including DNA methyla-

tion, histone modification, microRNA interference,and small interfering RNA. Epigenetic regulation such

as methylation of CDK2a and other genes has been

shown to occur.57 Methylation of death-associatedprotein kinase (DAPK) is associated with resistance

to anti-EGFR agents, like cetuximab.58 Jung et al

performed a combined analysis of the transcriptome,methylome, and miRNome of metastatic HNSCC and

non-metastatic HNSCC and identified a signature that

correlated with lower survival and metastatic pheno-type. The pathways involved in this group were

specifically related to cell–cell adhesion, EMT,

immune response, and apoptosis. For example, theyidentified decreased expression of desmoglein 3 (DSG

3), a component of desmosomes critical for cell–celladhesion. Desmosomes also have been shown to havetumor-suppressor function, and decreased expression

of DSG3 has been linked to a poor prognosis.59 They

also identified several elements significant in EMT,including upregulation of vimentin and downregula-

tion of cytokeratin intermediate fibers and activation

of TGF-β–related EMT pathways. Analysis of miRNAdemonstrated upregulation of pathways related to

DDR and immune response.60

CANCER-ASSOCIATED FIBROBLASTS

Normal squamous mucosal lining of the upperaerodigestive tract is organized into distinct

Page 6: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

J.M. Curry et al222

compartments: the upper layer of differentiated

squamous or respiratory epithelial cells, a basalepithelial layer, the underlying basement membrane,

and stromal layer. Fibroblasts are abundant in the

stroma and are the primary element responsible forsecretion of the basement membrane proteins. They

secrete structural proteins such as type IV collagen

and laminin and also produce numerous cytokinesand paracrine signals. Accordingly, tumor- or cancer-

associated fibroblasts (CAFs) are among the most

critical cellular elements of the TME. CAFs arephenotypically altered fibroblasts, which are active

participants in the process of tumorigenesis, promot-

ing growth and metastasis.61

CAFs arise from the population of circulating

fibroblasts and co-evolve with the tumor developing

a distinct phenotype, and playing an active role incarcinogenesis.62–64 They produce a variety of con-

tractile proteins, giving them an “active” phenotype.Frequently, they demonstrate ultrastructural accu-mulation of α-smooth muscle actin (α-SMA), charac-

teristic of myofibroblastic (MF) differentiation.65–67

In HNSCC, CAFs frequently have this MF phenotypeand are associated with dense collagen deposition

and stromal desmoplasia.68,69 CAFs are also charac-

terized by expression of integrin α6, which is criticalto cell adhesion and surface signaling. It complexes

to bind laminins, components of the extrcellular

matrix, and interacts with CDKN1A, altering cellcycle progression. Lim et al demonstrated that

upregulation of α-SMA and integrin-α6 correlated

with worsened prognosis in oral cancer.70

CAFs express a variety of factors critical to

carcinogenesis, promoting cell motility by upregula-

tion of cytokines, such as paracrine motility factor,hepatocyte growth factor (HGF), CXCL12, and

TGF-β.71 HGF secreted by CAFs has been shown to

promote invasion and angiogenesis in HNSCC andesophageal SCC.65,72–74 CXCL12 binds to CXCR4;

this interaction plays a role in upregulation of MMP9,

EMT, and HIF-1α expression.75 TGF-β is a criticalelement in the TME that serves numerous functions,

including immunosuppression. Additionally, CAFs

directly contribute to extracellular matrix remodel-ing by secreting MMPs.76,77

Marsh et al demonstrated that the MF phenotype

seen in some oral carcinomas was strongly prognos-tic of a negative outcome.78 This study evaluated 282

oral HNSCC specimens and found that the presence

of MF stroma was the strongest prognostic variableassessed, as compared to surgical margins, extracap-

sular spread, and stage, among others. MF stroma

correlated with depth of invasion and with extrac-apsular spread in nodal metastasis. Interestingly,

tumor-containing lymph nodes with extranodal

spread were also surrounded with MF stroma. Inoral and lingual carcinoma cell lines, Lin et al were

able to demonstrate increased proliferation in asso-

ciation with CAFs.79,80 In a mouse model usingheterotopic injection of HNSCC cells with normal

fibroblasts or CAFs, Wheeler et al demonstrated that

HNSCC cells with CAFs resulted in increased growthof the primary tumor and nodal and distant meta-

stases compared to co-injection with normal

fibroblasts.61

CAFs are also critical to tumor metabolism. Recent

studies indicate that epithelial cancer cells may

derive nutrients from the CAFs via a coupled meta-bolic mechanism. Cancerous cells induce glycolysis

in adjacent stromal cells such as CAFs and then use

their high-energy byproducts, such as lactate andpyruvate.81 This is somewhat contrarian to the long

held belief of the Warburg effect, whereby tumors

are thought to rely on aerobic glycolysis to produceenergy for rapid growth. This has been labeled the

“reverse Warburg effect” and has been shown to be a

critical prognostic indicator in breast and otherhuman cancers. There is some evidence that sug-

gests this occurs in HNSCC as well.82

THE IMMUNE RESPONSE IN THE TME

The persistent unresolved inflammation associ-

ated with cancer results in a eventual decay and

malfunction of the normal immune processes, whichin turn contributes to tumorigenesis through

immune tolerance and suppression and also to

angiogenesis and production of ROS. Essentially,tumorigenesis is at least in part a byproduct of a

failure of the immune system.10,12,83,84 The adaptive

immune response contributes in a variety of ways totumorigenesis through the immune interactions in

the TME involving T lymphocytes, macrophages,

dendritic cells, and others.85

T Lymphocytes

T lymphocytes are the central component of theanti-tumor response. They serve to initiate and regulate

the adaptive immune response and to elicit the cyto-

toxic response to tumors.85 There is evidence thatdysfunction occurs at the local, regional, and systemic

levels in HNSCC. While a strong lymphocytic host

presence at the tumor interface is indicative of anadaptive immune response and correlates with an

improved survival,86–88 dysfunctional circulating T cells

and tumor-infiltrating T cells have been identified inHNSCC, suggesting that tumors can suppress a previ-

ously intact local and systemic immune response.84,89–93 Moreover on a regional level, metastatic lymph nodesof HNSCC show significantly decreased levels of CD8þ

lymphocytes.87,94 Common functional deficits of

tumor-infiltrating T cells include: (1) absent or lowexpression of a key molecule in the signaling receptor

Page 7: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

Tumor microenvironment in HNSCC 223

receptor chain (CD3ζ), (2) decreased proliferation in

response to mitogens, (3) inability to kill tumor celltargets, (4) imbalance of their cytokine profile, and

(5) evidence of profound apoptotic features.84

Evasion of the adaptive response is executedthrough a variety of mechanisms such as decreased

expression of major histocompatibility complexes

(MHC I) or induction of apoptosis in T cells.Decreased expression of antigen-processing machi-

nery such as MHC glycoprotein allows escape of

subpopulations of tumor cells by avoiding activationof cell mediated immunity.95–97 This mechanism has

been demonstrated in HNSCC whereby tumor cells

produce gangliosides, which downregulate MHC I.98

Another means of evading detection is to induce

apoptosis in cytotoxic T cells. The FasL receptor

mechanism is expressed by activated cytotoxic T cells,which bind to FasL and typically result in triggering

the cytotoxic response. However, this also predis-

poses the T cell to apoptosis. Oral SCC cells have beenshown to contain membranous FasL-positive vesicles,

which trigger induction of T-cell apoptosis, circum-

venting the cytotoxic response.84,85,95

The cytotoxic response also can be dampened by

suppression. Intratumoral cytotoxic CD8þ T cells in

HNSCC show increased expression of programmeddeath-1 (PD-1), a marker of suppressed function.87,99

Its ligand, programmed death receptor ligand-1 (PD-

L1), is a surface protein that blocks function ofT lymphocytes and is expressed on malignant oral

SCC cells and also on CAFs.100 Cho et al demon-

strated that increased PD-L1 expression resulted inincreased apoptosis of intratumoral CD8þ TILs.101

Moreover, cytokines like, TGF-β, IL-10, and others

allow local naı̈ve T cells to be triggered to becomesuppressor T cells, while also exploiting the suppres-

sive functions of existing regulatory T cells.102PD-1 is

of particular interest in HPV-associated HNSCC, as alymphocytic infiltrate is one of the common features

of HPV-related OPSCC. Infiltration of the TME by PD-

1–positive T lymphocytes was correlated withimproved prognosis.103 While this is contrary to the

above findings, in the case of HPV-related OPSCC, the

PD-1–positive T lymphocytes, likely reflect an acti-vated chronic immune response due to long-standing

viral infection.103

ANTIGEN-PRESENTING CELLS AND TUMOR-ASSOCIATED MACROPHAGES

Dendritic cells are specialized antigen-presenting

cells (APCs) common in the TME of HNSCC.84,98,104

They have a high a capacity for antigen capture andalso stimulate T-cell maturation. In contrast, when

exposed to TGF-β and IL-10, they can promote

immune tolerance and differentiation of CD4þ Tcells into suppressive regulatory T cells.84,105–107

Langerhans cells are APCs located within the skin

and mucous membranes of the upper aerodigestivetract. They detect antigens in the mucosa and then

migrate to regional lymph nodes where they initiate

a primary immune response. Some evidence sug-gests that greater infiltration of HNSCC tumor sam-

ples with Langerhans cells correlates with improved

prognosis.84,108–110

Tumor-associated macrophages (TAMs) are

present with varying frequency in tumors, and are

common in HNSCC. TAMs are classified into twovarieties: proinflammatory (M1) and suppressive

(M2). Accordingly, studies in various cancers have

shown that TAMs can be associated with positive ornegative prognosis. M1 TAMs contribute to the anti-

tumor immune response via the production of

proinflammatory cytokines IL-12, IL-23, andinterferon-γ.84,111–113 While the M2 TAMs appear to

accumulate near blood vessels, promote angiogene-

sis,114,115 and produce a variety of suppressivecytokines such as IL-10 and TGF-β. They also serve

to promote tissue remodeling and inhibit anti-tumor

cytotoxic effects of M1 TAMs.84,111–113,116 Data inoral SCC suggest that TAMs are largely of the M2

type, as tumors with high levels of TAM infiltration

correlate with higher stage, lymph node metastasis,and extracapsular spread.114,117,118 Lago Costa et al

demonstrated that macrophages were increased in

the TME and the peripheral blood in HNSCC, andthat samples with increased TAMs showed increased

levels of TGF-β and its correlated immunosuppres-

sive effects.119 They produce ROS, RNS, and prosta-glandins (PGs), all of which can contribute to

inflammation and tumorigenesis. COX2 inhibitors

and nitric oxide synthase inhibitors (iNOS) havebeen used to antagonize these inflammatory agents

and their cytokines.120,121 TAMs in HNSCC also

produce significant levels of macrophage migrationinhibitory factor (MIF), which is an inflammatory

cytokine that stimulates neutrophils. MIF recruits

neutrophils to the tumor via a CXCR2 mechanismand then by feedback mechanisms increases inva-

siveness of the tumor cells.122 Neutrophils act on the

tumor in a variety of ways: inducing genetic insta-bility via ROS, increasing angiogenesis via MMP9 and

VEGF, and increasing invasion via HGF.123

THE BASEMENT MEMBRANE, INVASION, ANDMATRIX METALLOPROTIENASES

The basement membrane is barrier to tumor pro-

gression, and its degradation facilitates tumor invasion

and metastasis. For this to occur, cancer cells must(1) develop motility, (2) alter cell–cell adhesion, and(3) remodel the ECM.124 The basement membrane not

only serves as a structural framework for the overlyingepithelial cells but also provides paracrine signals that

Page 8: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

J.M. Curry et al224

affect their behaviors such as differentiation and migra-

tion.125 Many of the key elements of the basementmembrane, including collagen type IV and fibronectin,

have been shown to be disregulated in HNSCC. MMPs

are most important group of proteolytic enzymes usedby cancer to degrade the ECM. MMPs in normal tissues

are expressed in balance with their inhibitors to main-

tain a well-organized system. MMPs are upregulated byNOTCH1 pathways, EGFR, TGF-β, HGF, and

granulocyte-macrophage colony-stimulating factor (GM-

CSF), which are commonly overexpressed inHNSCC.126–129 Among the most commonly identified

metalloproteinases in HNSCC are MMP-2, MMP-9 and

membrane-bound MMP (MT-MMP). MMP-2 and MMP-9are gelatinases and degrade collagen type IV, the most

critical step in degrading the BM.130 Increased levels of

MMP-2 and MMP-9 correlate with increased nodalmetastasis and poor prognosis.48 MMP-9 is the most

structurally complex and can degrade numerous ele-

ments of the TME, including elastin, fibrillin, laminin,gelatin, and types IV, V, XI, and XVI collagen.131,132

MMPs were initially thought to be produced solely by

the tumor cell, but further investigation has shownproduction also by the CAFs and surrounding inflam-

matory cells.48,133,134 CAFs are primarily responsible for

the increased production of MMP-2 in co-culture experi-ments.130 MT-MMP is critical in activating MMP-2.135,136

There are numerous other significant MMPs, such as

MMP-13, which participates in angiogenesis increasingthe level of VEGF at the invasive front.137

Importantly, the functions of MMPs extend

beyond protein degradation and invasion, as theytarget growth factors, growth factor receptors, and

cytokines.8 For example, MMP-9 also produces a

tolerogenic effect on dendritic APCs and also onregulatory T cells.43 Release of MMP-9 results in

endothelial cell invasion and vessel formation.12,55

MMPs impact differentiation and maturation of bonecells into osteoclasts, which is critical to the process

of bony invasion.138 HNSCC CSCs are characterized

by expression of CD44; CD44 is a surface proteinthat functions as a receptor for hyaluronic acid and

also is the docking receptor necessary for MMP-9

function.139 Given the broad significance of MMPsthey represent a possible target for therapy directed

at the TME. Interestingly, quercitin, a flavonoid

isolated from onions, inhibits MMP-2 and -9pathways.140

TGF-b AND EPITHELIAL–MESENCHYMALTRANSITION

A number of chemokines and cytokines providecritical paracrine signaling in the TME; here we focus

on TGF-β, as it broadly impacts many cellular behav-

iors in the TME. TGF-β has both growth-promotingand -suppressive effects on cells, and for some time

the role of TGF-β in malignancy had been controver-

sial. It typically inhibits epithelial cell proliferationand promotes secretion of matrix proteins and

proteases. Currently it is understood to act as a

tumor-suppressor early in tumorigenesis, then in laterphases it enhances the malignant phenotype.141 TGF-

β primarily acts through the SMAD family of tran-

scription factors and works in concert with mitogen-activated protein kinases (MAPKs), which regulate

diverse cellular activities such as mitosis, differentia-

tion, proliferation, cell survival, and apoptosis. Dys-regulation of TGF-β in malignancy occurs through

several mechanisms, including loss of response to its

ligand, defects in the transduction pathway, andothers.141 Oral SCC has been shown to be resistant

to the suppressive effects of TGF-β, secondary to

downregulation of TGF-β receptor II (TBRII).141,142

TGF-β is a primary factor triggering EMT in

HNSCC. EMT contributes to invasion allowing for

enhanced mobility via expression of a protein expres-sion patterns more characteristic of a mesenchymal

phenotype. Once established, nests of metastatic

tumor can transition back to a phenotype recapitulat-ing the original tumor in a distant site. EMT is

mediated through disruption of epithelial cell junc-

tions, remodeling of the actin cytoskeleton, andupregulation of mesenchymal markers like vimentin

and firbonectin.141 TGF-β pathways as well as those

triggered by the inflammatory cytokines TNF-α andIL-6 converge upon STAT3, upregulating it.52,143,144

STAT3 proteins are commonly overexpressed in

HNSCC. STAT3 interacts with Twist, Snail, and Slug(Snail2), transcription factors that contribute to EMT

in various cancers.52,144,145 Twist increases expres-

sion of N cadherin, a marker of a mesenchymalmotile phenotype, and decreases expression of E

cadherin, a marker of an epithelial phenotype. Slug

also decreases expression of E-cadherin. Loss ofE-cadherin and gained expression of N-cadherin is

critical to invasion and is referred to as cadherin

switching.146 Prime et al were able to demonstratemorphologic evidence of EMT and cadherin switch-

ing after several days exposure to TGF-β.138,147

Emerging evidence suggests that EMT is fundamentalto gaining “stemness” or the transition of cancer cells

to becoming CSCs. CSCs are thought to serve as a

fountainhead for tumors as they give rise to theremaining population of tumor cells, and contribute

to treatment resistance. CSCs accumulate at the

invasive front and perivascular spaces and are demar-cated by expression of markers such as CD133 and

CD44, and by aldehyde dehydrogenase activity.139

On the surface of CSCs, CD44 interacts with MMP-9and this allows for proteolytic activation of TGF-β.148

TGF-β extends beyond the epithelial cancer cells of

a tumor, and many of the effects have been describedabove. Lewis et al showed that TGF-β produced at the

Page 9: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

Tumor microenvironment in HNSCC 225

invasive leading edge of the tumor induced a MF

phenotype in primary fibroblasts. They also showedthat this effect resulted in secretion of HGF by

myofibrblasts, which in turn promoted invasion

through the basement membrane. TGF-β serves toinhibit TH1 lymphocytes and cytotoxic T lympho-

cytes and the functions of natural killer cells.84

ANGIOGENESIS, INFLAMMATION, ANDHYPOXIA

Small tumor deposits of 1–3 mm can be supplied

by diffusion of nutrients from the surrounding tissue;

beyond this, the tumor is dependent on angiogenesisto supply its needs.149 A number of studies have

shown that angiogenesis is correlated with tumor

aggression.48,150–154 HNSCC often has large hypoxicareas of tumor necrosis where growth exceeds

angiogenesis.155–157 Hypoxic response and inflam-

mation are driving forces in angiogenesis.12,158 More-over, CSCs in HNSCC appear to be concentrated

along the invasive front of the tumor and in the

perivascular niche, an area within 100 μm of themicrovasculature. A variety of factors in the TME,

such as VEGF, NF-κB, and HIF-1α play central roles in

this process.VEGF enhances endothelial growth, migration of

endothelial precursors, and their differentiation. High

VEGF expression in oral SCC has been correlatedwith a poor prognosis, and a recent meta-analysis

suggested that VEGF overexpression could be a

useful prognostic marker.159 VEGF binds to its recep-tor, VEGFR1 in tumor cells, and induces expression

of Bcl-2, inducing chemokines like CXCL1 and

CXCL8. CXCL1 and CXCL8 promote endothelial cellproliferation and survival.160 Endothelial cells in turn

produce factors like EGF, which significantly increase

tumor cell survival and migration.161 VEGF and otherangiogenic factors such as IL-6 and IL-8 are increased

by a number of chemokines such as CXCL12, which

binds to chemokine receptors CXCR2 and CXCR4.High CXCR2 and CXCR4 levels have been shown to

be associated with increased microvessel density

within tumors.55,77,162

Chronic inflammation of the TME contributes to

tumor progression through a variety of mechanisms,

including production ROS and angiogenic factors.NF-κβ is an inflammatory signal transcription factor

playing a variety roles in invasion, proliferation, and

angiogenesis. Constitutive activation NF-κβ results inoverexpression of a variety of factors, including Il-6,

IL-8, and VEGF.43

There are many downstream inflammatorymarkers expressed as a result of NF-κβ and other

mechanisms, such as cyclooxygenases like COX-2.163

COX enzymes catalyze the production of PGs andlikely are the rate-limiting step in their synthesis.

COX-2 is usually overexpressed in inflammation and

preneoplastic lesions. PGs are increased in HNSCC,and PGE2 promotes invasion and angiogenesis and

inhibits apoptosis of cancer cells. COX-2 acts on

VEGF, fibroblast growth factor, and MMPs and is alsopro-angiogenic. COX-2 levels have been found to be

prognostic and selective COX-2 inhibitors have

been shown to increase the efficacy of radio-therapy in vitro.51,164 NF-κβ is the target of many

therapeutic interventions, such as curcumin, n-acetyl

cysteine (NAC), epigallocatechin gallate (EGCG), andothers.5

Intratumoral hypoxia is a key characteristic of

HNSCC, and is a negative prognostic factor, contri-buting to both chemotherapy and radiotherapy resist-

ance. Intratumoral hypoxia is generally accepted to

be a pO2 o10 mm Hg, and intratumoral pO2 levels≤2.5 mm Hg correlate with a worsened prognosis, as

does the overall volume of hypoxic tumor at the

primary site.165 HIF-1α is the most important factorinduced in adaptive response to hypoxia, and ele-

vated expression is also directly associated with a

poor prognosis.166 This transcription factor interactswith more than 100 genes to alter expression of

VEGF, CA9, lysyl oxidase, and many others.48,167 It

has been shown to alter cellular metabolism, and toincrease lymphatic vessel density and blood vessel

density in oral SCC.168,169 CA9 functions to regulate

pH homeostasis and alter the uptake of chemother-apeutic drugs, and also is purported to play a role in

proliferation and cell adhesion.165,170 Lysyl oxidase

catalyzes the crosslinking of collagens and elastins,and overexpression increases microvascular den-

sity.171,172 Agents such as reseveratrol, EGCG and

others may act by promoting degradation of HIF-1α.173–175 Resveratrol has been shown to decrease

expression of HIF-1α and VEGF in vitro.176

METABOLISM IN THE TME

Cancer cells have high bioenergetic requirementsneeded to maintain tumor growth. Tumor cells in

culture have long been demonstrated to rely heavily

on glycolysis with decreased, dysfunctional, or absentmitochondrial OXPHOS. Reliance on glycolysis in the

presence of oxygen is referred to as the Warburg

effect.177 This results in the generation of less ATPthan OXPHOS and yields high levels of pyruvate and

lactate. This is somewhat counterintuitive as there is

such a high bioenergetics requirement, yet OXPHOSis a more efficient means of energy generation than

glycolysis. Thus it is unclear why tumor cells would

thrive with a less efficient mechanism. It has beenhypothesized that glycolysis may confer a growth

advantage.178–180 Some normal, highly proliferative

cells, such as lymphocytes, favor aerobic glycolysisover oxidative metabolism, providing a rationale for

Page 10: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

J.M. Curry et al226

the Warburg effect.181 Many cancer cells have defects

in critical components of the OXPHOS pathway, suchas the mitochondrial B-catalytic subunit of Hþ-ATPsynthase.182,183 Furthermore, when glycolytic flux is

high, the ATP yield can exceed that produced byOXPHOS.182,183 Additionally, the intermediates of

glycolytic metabolism can provide substrates for

amino acid, fatty acid, and nucleotide synthesis.167

The metabolic pressures induced by hypoxia in the

setting of rapid growth may then in turn select for

tumor cells which favor glycolytic metabolism evenin the presence of oxygen, as is suggested by the

frequent overexpression of HIF-1α in many cancers.

Hypoxic induction of HIF-1α favors this processspecifically inducing pyruvate dehydrogenase kinase

(PDK) and lactate dehydrogenase a (LDH-A). PDK

inactivates pyruvate dehydrogenase preventingimport of pyruvate to the mitochondria. LDH-A

restores NAD positivity and also uses pyruvate in

the cytosol, which together can reduce electron flowthough OXPHOS and also reduce oxidative stress.

Additionally, glycolytic metabolism results in the

acidotic efflux into the TME that assists in breakdownof the ECM and kills non-adapted normal cells.184

However, this is not likely the whole picture:

much recent evidence suggests that a metabolicsymbiosis exists within tumors cell between differ-

ent populations. Feron has likened this to the

coupling between fast and slow-twitch musclefibers. Fast-twitch glycolytic fibers release lactate

that is then taken up and utilized by slow twitch

fibers. MCT1 is a high-affinity transporter of lactate,which mediates influx into the cell; MCT4 is a low-

affinity transporter of lactate, which primarily medi-

ates efflux of lactate from cells. These transporterscouple cancer cells, so that hypoxic cells maintain

functioning glycolytic metabolism while aerobic

tumor cells recycle and utilize lactate and otherhigh-energy substrates produced by them. A similar

process in cancer would allow for an efficient intra-

tumoral metabolic coupling mechanism betweenoxygenated cells and hypoxic cells.179

Additional evidence favors multicompartmental

metabolism between the cancer cells and CAFs.Numerous co-culture experiments and in situ tumor

analyses have demonstrated this effect in breast and

other cancers.11,81,185 This work has brought to lighta “reverse Warburg effect”, where oxidative stresses

exerted by tumor cells induce aerobic glycolysis and

autophagy in CAFs. This, in turn, results in increasedlevels of intermediate catabolites such as lactate,

glutamine, and ketone bodies. These catabolites are

released into the TME and used for OXPHOS incarcinoma cells. This metabolically enriches the TME

and creates an environment that favors growth,

apoptosis resistance, invasion, and metastasis.81,185

This is Paget’s seed and soil hypothesis, a

phenomenon that may have been unnoticed in

previous homotypic culture experiments.11

Most studies on HNSCC cellular metabolism sug-

gest that the carcinoma cells are highly glycolytic

with high L-lactate generation, yet recent studiessuggest that metabolic heterogeneity and metabolic

coupling occur. Most HNSCC cells generate signifi-

cantly higher levels of lactate compared with normalhuman oral keratinocytes (NHOK), although several

cell lines generate significantly lower lactate levels

than NHOKs.186 It has been postulated that the cellswith decreased lactate production have increased

lactate uptake via MCTs, allowing them to utilize

OXPHOS.186 When some HNSCC cell lines that aretypically glycolytic are supplemented with excess

pyruvate, some of the effects were reversed, which

suggests that OXPHOS is important to supportHNSCC cell proliferation in the presence of a

catabolite-rich microenvironment.187 High tumor

lactate concentrations in HNSCC are associated withsubsequent nodal and distant metastatses.188,189

In our previously published work on oral SCC, we

demonstrated evidence of this multicompartmentmodel of metabolism. We have suggested that there

may be three metabolic compartments in HNSCC,

where the leading tumor edge relies on OXPHOSand the deeper layers of the tumor are more

glycolytic (aerobic or anaerobic) and tumor stroma

represents a third compartment undergoing aerobicglycolysis (Figure 2). This was demonstrated through

high expression of MCT4 in the stroma and deeper

tumor, while MCT1 was more highly expressed bythe leading tumor edge. We also confirmed OXPHOS

in the leading tumor edge with assays for TOMM20

and LDHb, both functional markers for mitochon-drial metabolism. This pattern of metabolic coupling

was demonstrated in a subset of our oral SCC

patients, and correlated with aggressive behaviorincluding a worsened disease-free survival and peri-

neural invasion. Interestingly, it also correlated with

increased specific uptake values (SUV) on positronemission tomography/computed tomography. We

further tested this metabolic coupling theory with

a squamous cell carcinoma line co-culture experi-ment. Using immortalized squamous cell lines we

were able to generate two divergent SCC popultions,

one RAS-dependent and another NF-κB–dependent.These cell lines were each able to induce metabolic

reprogramming of CAFs via oxidative stress. This

resulted in a lactate shuttling process that feeds thecancer cells fueling anabolic growth via and MCT1/

MCT4 metabolic couple between the tumor and the

stroma. Interestingly, this model also demonstratedthat the CAFs protected the cancer cells against

oxidative stress by reducing oxidative stresses within

the carcinoma cells. RAS-transformed cells were ableto reprogram adjacent epithelial cells, as well as

Page 11: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

Figure 2. Metabolic coupling in HNSCC. The leading edge of the tumor relies on OXPHOS while the inner compartmentand CAFs rely on glycolytic metabolism. Higher expression of MCT1 is seen in the leading tumor edge, while higherexpression of MCT4 is seen in the central compartment and stromal CAFs.

Tumor microenvironment in HNSCC 227

fibroblasts, suggesting that cancer cells can subju-gate either group.

MCT4 may represent a possible target for meta-

bolic interruption and uncoupling of the tumor andstroma. In an animal model, we were able to

demonstrate that NAC was able to selectively inhibit

MCT4 induction in CAFs, halting mitochondrial bio-genesis in cancer cells but not in normal epithelial

cells. This may allow targeted therapy that selec-

tively starves cancer cells. MCT1 also has beenproposed as a possible target to prevent uptake of

lactate, forcing aerobic cells to use glucose and

depriving or decreasing availability to hypoxiccells.167 In fact, the MCT1 inhibitor, a-cyano-

hydroxycinnamate has been shown to slow tumorgrowth and potentiate the effect of radiotherapy in

MCT1-expressing tumors in mice.179,184

Hypoxia contributes to chemotherapy and radio-therapy resistance.155 Inhibition of HIF-1α can pre-

vent the induction of the hypoxic response blocking

angiogenesis.190 Zhang et al inhibited HIF-1α withsiRNA and oligonucleotides, which increased apop-

tosis in oral SCC.191 The EGFR inhibitor cetuximab

blocks downstream signaling activated by EGFR; thistriggers G1 phase arrest and can also trigger apop-

tosis.192 In addition, it has been shown to down-

regulate HIF-1α; this, in turn, downregulates theirLDH-a and glycolytic potential. This inhibition of

Page 12: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

J.M. Curry et al228

glycolytic potential leads to inhibition of

proliferation.192

Metformin is a commonly used antihyperglycemic

drug in type 2 diabetics and has been proposed as a

potential anticancer therapy also that may impacttumor metabolism in the TME. Metformin has been

shown to inhibit cancer cell proliferation in several

human cancers, such as gastric, medullary thyroid,breast, and pancreatic cancers.193–196 Epidemiologic

studies also have shown significant effects from

metformin use in diabetics, lowering the risk ofcancer incidence and mortality.197 In oral SCC, Luo

et al demonstrated that metformin blocked cell cycle

progression at the G0/G1 phase and induced apop-tosis. Metformin triggered alterations in multiple

other pathways as well: increasing activation of the

adenosine monophosphate (AMP) kinase pathway,suppressing the mammalian target of rapamycin

(mTOR) pathway, decreasing cyclin D1 levels and

retinoblastoma (Rb) phosphorylation, and downregu-lating Bcl 2.198 They also were able to demonstrate

in vivo evidence of increased apoptosis in a xenograft

model. While this study demonstrated various effectson the cell cycle; the metabolic effects of metformin

on cancer have yet to be investigated.

CONCLUSION

Many elements of the TME beyond the cancerousepithelial cells impact progression of HNSCC.

Genetic alterations induced by tobacco and alcohol

or the HPV virus initiate the sequence of events thattrigger transformation of stromal cells, immune sup-

pression, and chronic inflammation. In turn,

unchecked growth, invasion, and metastasis prevail.The complexity of these processes reveals that the

long-held notion of “condemned mucosa” actually

reflects a “condemned tissue” comprised of many celltypes which have co-evolved during tumorigenesis.

REFERENCES1. Kamangar F, Dores GM, Anderson WF. Patterns of

cancer incidence, mortality, and prevalence across

five continents: defining priorities to reduce cancer

disparities in different geographic regions of the

world. J Clin Oncol. 2006;24:2137–50.

2. Marur S, Forastiere AA. Head and neck cancer:

changing epidemiology, diagnosis, and treatment.

Mayo Clin Proc. 2008;83:489–501.

3. Leemans CR, Tiwari R, Nauta JJ, van der Waal I, Snow

GB. Recurrence at the primary site in head and neck

cancer and the significance of neck lymph node

metastases as a prognostic factor. Cancer. 1994;73:

187–90.

4. Slaughter DP, Southwick HW, Smejkal W. Field

cancerization in oral stratified squamous epithelium;

clinical implications of multicentric origin. Cancer.

1953;6:963–8.

5. Albini A, Sporn MB. The tumour microenvironment

as a target for chemoprevention. Nat Rev Cancer.

2007;7:139–47.

6. Hanahan D, Weinberg RA. The hallmarks of cancer.

Cell. 2000;100:57–70.

7. Haddad RI, Shin DM. Recent advances in head and

neck cancer. N Engl J Med. 2008;359:1143–54.

8. Mbeunkui F, Johann DJ, Jr. Cancer and the tumor

microenvironment: a review of an essential relation-

ship. Cancer Chemother Pharmacol. 2009;63:571–82.

9. Weinberg RA. Twisted epithelial-mesenchymal tran-

sition blocks senescence. Nat Cell Biol. 2008;10:

1021–3.

10. Cavallo F, De Giovanni C, Nanni P, Forni G, Lollini

PL. The immune hallmarks of cancer. Cancer Immu-

nol Immunother. 2011;60:319–26.

11. Paget S. The distribution of secondary growths in

cancer of the breast. Cancer Metastasis Rev. 1989;8:

98–101.

12. Albini A, Tosetti F, Benelli R, Noonan DM. Tumor

inflammatory angiogenesis and its chemoprevention.

Cancer Res. 2005;65:10637–41.

13. Karin M. NF-kappaB and cancer: mechanisms and

targets. Mol Carcinog. 2006;45:355–61.

14. Bertl E, Bartsch H, Gerhauser C. Inhibition of angio-

genesis and endothelial cell functions are novel

sulforaphane-mediated mechanisms in chemopre-

vention. Mol Cancer Ther. 2006;5:575–85.

15. Leemans CR, Braakhuis BJ, Brakenhoff RH. The

molecular biology of head and neck cancer. Nat

Rev Cancer. 2011;11:9–22.

16. Ramqvist T, Dalianis T. An epidemic of oropharyng-

eal squamous cell carcinoma (OSCC) due to human

papillomavirus (HPV) infection and aspects of treat-

ment and prevention. Anticancer Res. 2011;31:

1515–9.

17. Syrjanen S, Lodi G, von Bultzingslowen I, Aliko A,

Arduino P, Campisi G, et al. Human papillomaviruses

in oral carcinoma and oral potentially malignant

disorders: a systematic review. Oral Dis. 2011;17

(Suppl 1):58–72.

18. Rautava J, Syrjanen S. Biology of human papilloma-

virus infections in head and neck carcinogenesis.

Head Neck Pathol. 2012;6(Suppl 1):S3–15.

19. Boyle JO, Hakim J, Koch W, van der Riet P, Hruban

RH, Roa RA, et al. The incidence of p53 mutations

increases with progression of head and neck cancer.

Cancer Res. 1993;53:4477–80.

20. el-Naggar AK, Lai S, Luna MA, Zhou XD, Weber RS,

Goepfert H, et al. Sequential p53 mutation analysis of

pre-invasive and invasive head and neck squamous

carcinoma. Int J Cancer. 1995;64:196–201.

21. Tan M, Myers JN, Agrawal N. Oral cavity and

oropharyngeal squamous cell carcinoma genomics.

Otolaryngol Clin North Am. 2013;46:545–66.

22. Koch WM, Brennan JA, Zahurak M, Goodman SN,

Westra WH, Schwab D, et al. p53 mutation and

locoregional treatment failure in head and neck

squamous cell carcinoma. J Natl Cancer Inst. 1996;

88:1580–6.

23. Alsner J, Sorensen SB, Overgaard J. TP53 mutation is

related to poor prognosis after radiotherapy, but not

Page 13: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

Tumor microenvironment in HNSCC 229

surgery, in squamous cell carcinoma of the head and

neck. Radiother Oncol. 2001;59:179–85.

24. Skinner HD, Sandulache VC, Ow TJ, Meyn RE, Yordy

JS, Beadle BM, et al. TP53 disruptive mutations lead

to head and neck cancer treatment failure through

inhibition of radiation-induced senescence. Clin Can-

cer Res. 2011;18:290–300.

25. Huang X, Pateromichelakis S, Hills A, Sherriff M,

Lyons A, Langdon J, et al. p53 mutations in deep

tissues are more strongly associated with recurrence

than mutation-positive mucosal margins. Clin Cancer

Res. 2007;13:6099–106.

26. Lin SY, Dolfi SC, Amiri S, Li J, Budak-Alpdogan T, Lee

KC, et al. P53 regulates the migration of mesenchy-

mal stromal cells in response to the tumor micro-

environment through both CXCL12-dependent and

-independent mechanisms. Int J Oncol. 2013;43:

1817–23.

27. He H, Tian D, Guo J, Liu M, Chen Z, Hamdy FC, et al.

DNA damage response in peritumoral regions of

oesophageal cancer microenvironment. Carcinogen-

esis. 2013;34:139–45.

28. Bhowmick NA, Neilson EG, Moses HL. Stromal

fibroblasts in cancer initiation and progression.

Nature. 2004;432:332–7.

29. Hu M, Polyak K. Microenvironmental regulation of

cancer development. Curr Opin Genet Dev. 2008;18:

27–34.

30. Szatrowski TP, Nathan CF. Production of large

amounts of hydrogen peroxide by human tumor

cells. Cancer Res. 1991;51:794–8.

31. Cairns RA, Harris IS, Mak TW. Regulation of cancer

cell metabolism. Nat Rev Cancer. 2011;11:85–95.

32. Bolos V, Grego-Bessa J, de la Pompa JL. Notch

signaling in development and cancer. Endocr Rev.

2007;28:339–63.

33. Yoshida R, Nagata M, Nakayama H, Niimori-Kita K,

Hassan W, Tanaka T, et al. The pathological signifi-

cance of Notch1 in oral squamous cell carcinoma.

Lab Invest. 2013;93:1068–81.

34. Liao S, Xia J, Chen Z, Zhang S, Ahmad A, Miele L,

et al. Inhibitory effect of curcumin on oral carcinoma

CAL-27 cells via suppression of Notch-1 and NF-

kappaB signaling pathways. J Cell Biochem. 2011;

112:1055–65.

35. Subramaniam D, Ponnurangam S, Ramamoorthy P,

Standing D, Battafarano RJ, Anant S, et al. Curcumin

induces cell death in esophageal cancer cells

through modulating Notch signaling. PLoS One.

2012;7:e30590.

36. Temam S, Kawaguchi H, El-Naggar AK, Jelinek J,

Tang H, Liu DD, et al. Epidermal growth factor

receptor copy number alterations correlate with

poor clinical outcome in patients with head and

neck squamous cancer. J Clin Oncol. 2007;25:

2164–70.

37. Chung CH, Ely K, McGavran L, Varella-Garcia M,

Parker J, Parker N, et al. Increased epidermal growth

factor receptor gene copy number is associated with

poor prognosis in head and neck squamous cell

carcinomas. J Clin Oncol. 2006;24:4170–6.

38. Ang KK, Berkey BA, Tu X, Zhang HZ, Katz R,

Hammond EH, et al. Impact of epidermal growth

factor receptor expression on survival and pattern of

relapse in patients with advanced head and neck

carcinoma. Cancer Res. 2002;62:7350–6.

39. Rubin Grandis J, Melhem MF, Gooding WE, Day R,

Holst VA, Wagener MM, et al. Levels of TGF-alpha

and EGFR protein in head and neck squamous cell

carcinoma and patient survival. J Natl Cancer Inst.

1998;90:824–32.

40. Agrawal N, Frederick MJ, Pickering CR, Bettegowda

C, Chang K, Li RJ, et al. Exome sequencing of head

and neck squamous cell carcinoma reveals inactivat-

ing mutations in NOTCH1. Science. 2011;333:

1154–7.

41. Sheu JJ, Hua CH, Wan L, Lin YJ, Lai MT, Tseng HC,

et al. Functional genomic analysis identified epider-

mal growth factor receptor activation as the most

common genetic event in oral squamous cell carci-

noma. Cancer Res. 2009;69:2568–76.

42. Kim S, Grandis JR, Rinaldo A, Takes RP, Ferlito A.

Emerging perspectives in epidermal growth factor

receptor targeting in head and neck cancer. Head

Neck. 2008;30:667–74.

43. Wang BQ, Zhang CM, Gao W, Wang XF, Zhang HL,

Yang PC. Cancer-derived matrix metalloproteinase-9

contributes to tumor tolerance. J Cancer Res Clin

Oncol. 2011;137:1525–33.

44. Nijkamp MM, Span PN, Bussink J, Kaanders JH.

Interaction of EGFR with the tumour microenviron-

ment: implications for radiation treatment. Radiother

Oncol. 2013;108:17–23.

45. Zimmermann M, Zouhair A, Azria D, Ozsahin M. The

epidermal growth factor receptor (EGFR) in head

and neck cancer: its role and treatment implications.

Radiat Oncol. 2006;1:11.

46. Rogers SJ, Harrington KJ, Rhys-Evans P, P OC, Eccles

SA. Biological significance of c-erbB family onco-

genes in head and neck cancer. Cancer Metastasis

Rev. 2005;24:47–69.

47. Kalyankrishna S, Grandis JR. Epidermal growth factor

receptor biology in head and neck cancer. J Clin

Oncol. 2006;24:2666–72.

48. Koontongkaew S. The tumor microenvironment

contribution to development, growth, invasion and

metastasis of head and neck squamous cell carcino-

mas. J Cancer. 2013;4:66–83.

49. Horiguchi M, Koyanagi S, Okamoto A, Suzuki SO,

Matsunaga N, Ohdo S. Stress-regulated transcription

factor ATF4 promotes neoplastic transformation by

suppressing expression of the INK4a/ARF cell sen-

escence factors. Cancer Res. 2012;72:395–401.

50. Bova RJ, Quinn DI, Nankervis JS, Cole IE, Sheridan

BF, Jensen MJ, et al. Cyclin D1 and p16INK4A

expression predict reduced survival in carcinoma

of the anterior tongue. Clin Cancer Res. 1999;5:

2810–9.

51. Wang F, Arun P, Friedman J, Chen Z, Van Waes C.

Current and potential inflammation targeted thera-

pies in head and neck cancer. Curr Opin Pharmacol.

2009;9:389–95.

Page 14: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

J.M. Curry et al230

52. Chen YW, Chen KH, Huang PI, Chen YC, Chiou GY,

Lo WL, et al. Cucurbitacin I suppressed stem-like

property and enhanced radiation-induced apoptosis

in head and neck squamous carcinoma–derived

CD44(þ)ALDH1(þ) cells. Mol Cancer Ther. 2010;9:

2879–92.

53. Du Y, Peyser ND, Grandis JR. Integration of molec-

ular targeted therapy with radiation in head and neck

cancer. Pharmacol Ther. 2013.

54. Chung CH, Parker JS, Karaca G, Wu J, Funkhouser

WK, Moore D, et al. Molecular classification of head

and neck squamous cell carcinomas using patterns of

gene expression. Cancer Cell. 2004;5:489–500.

55. Benelli R, Morini M, Carrozzino F, Ferrari N, Min-

ghelli S, Santi L, et al. Neutrophils as a key cellular

target for angiostatin: implications for regulation of

angiogenesis and inflammation. Faseb J. 2002;16:

267–9.

56. Clatot F, Gouerant S, Mareschal S, Cornic M, Ber-

ghian A, Choussy O, et al. The gene expression

profile of inflammatory, hypoxic and metabolic

genes predicts the metastatic spread of human head

and neck squamous cell carcinoma. Oral Oncol.

2013;50:200–7.

57. Shaw R. The epigenetics of oral cancer. Int J Oral

Maxillofac Surg. 2006;35:101–8.

58. Ogawa T, Liggett TE, Melnikov AA, Monitto CL,

Kusuke D, Shiga K, et al. Methylation of death-

associated protein kinase is associated with cetux-

imab and erlotinib resistance. Cell Cycle. 2012;11:

1656–63.

59. Alaibac M. Targeting DSG3: from pemphigus to

squamous cell carcinoma. Expert Opin Ther Targets.

2013;17:477–9.

60. Jung AC, Job S, Ledrappier S, Macabre C, Abecassis J,

de Reynies A, et al. A poor prognosis subtype of

HNSCC is consistently observed across methylome,

transcriptome, and miRNome analysis. Clin Cancer

Res. 2013;19:4174–84.

61. Wheeler SE, Shi H, Lin F, Dasari S, Bednash J, Thorne

S, et al. Enhancement of head and neck squamous

cell carcinoma proliferation, invasion, and metastasis

by tumor-associated fibroblasts in preclinical models.

Head Neck. 2013.

62. Bucala R, Spiegel LA, Chesney J, Hogan M, Cerami A.

Circulating fibrocytes define a new leukocyte sub-

population that mediates tissue repair. Mol Med.

1994;1:71–81.

63. Abe R, Donnelly SC, Peng T, Bucala R, Metz CN.

Peripheral blood fibrocytes: differentiation pathway

and migration to wound sites. J Immunol. 2001;166:

7556–62.

64. Franco OE, Shaw AK, Strand DW, Hayward SW.

Cancer associated fibroblasts in cancer pathogenesis.

Semin Cell Dev Biol. 2010;21:33–9.

65. Kawashiri S, Tanaka A, Noguchi N, Hase T, Nakaya

H, Ohara T, et al. Significance of stromal desmoplasia

and myofibroblast appearance at the invasive front in

squamous cell carcinoma of the oral cavity. Head

Neck. 2009;31:1346–53.

66. Tomasek JJ, Gabbiani G, Hinz B, Chaponnier C,

Brown RA. Myofibroblasts and mechano-regulation

of connective tissue remodelling. Nat Rev Mol Cell

Biol. 2002;3:349–63.

67. Tlsty TD, Hein PW. Know thy neighbor: stromal cells

can contribute oncogenic signals. Curr Opin Genet

Dev. 2001;11:54–9.

68. Chen Y, Satoh T, Sasatomi E, Miyazaki K, Tokunaga

O. Critical role of type IV collagens in the growth of

bile duct carcinoma. In vivo and in vitro studies.

Pathol Res Pract. 2001;197:585–96.

69. Kunz-Schughart LA, Knuechel R. Tumor-associated

fibroblasts (part I): Active stromal participants in

tumor development and progression? Histol Histo-

pathol. 2002;17:599–621.

70. Lim KP, Cirillo N, Hassona Y, Wei W, Thurlow JK,

Cheong SC, et al. Fibroblast gene expression profile

reflects the stage of tumour progression in oral

squamous cell carcinoma. J Pathol. 2011;223:

459–69.

71. Leef G, Thomas SM. Molecular communication

between tumor-associated fibroblasts and head and

neck squamous cell carcinoma. Oral Oncol. 2013;

49:381–6.

72. Grugan KD, Miller CG, Yao Y, Michaylira CZ, Ohashi

S, Klein-Szanto AJ, et al. Fibroblast-secreted hepato-

cyte growth factor plays a functional role in esoph-

ageal squamous cell carcinoma invasion. Proc Natl

Acad Sci U S A. 2010;107:11026–31.

73. Knowles LM, Stabile LP, Egloff AM, Rothstein ME,

Thomas SM, Gubish CT, et al. HGF and c-Met

participate in paracrine tumorigenic pathways in

head and neck squamous cell cancer. Clin Cancer

Res. 2009;15:3740–50.

74. Rousseau B, Larrieu-Lahargue F, Javerzat S, Guilhem-

Ducleon F, Beermann F, Bikfalvi A. The tyrp1-Tag/

tyrp1-FGFR1-DN bigenic mouse: a model for selec-

tive inhibition of tumor development, angiogenesis,

and invasion into the neural tissue by blockade of

fibroblast growth factor receptor activity. Cancer

Res. 2004;64:2490–5.

75. Ishikawa T, Nakashiro K, Klosek SK, Goda H, Hara S,

Uchida D, et al. Hypoxia enhances CXCR4 expres-

sion by activating HIF-1 in oral squamous cell

carcinoma. Oncol Rep. 2009;21:707–12.

76. De Wever O, Demetter P, Mareel M, Bracke M.

Stromal myofibroblasts are drivers of invasive cancer

growth. Int J Cancer. 2008;123:2229–38.

77. Orimo A, Gupta PB, Sgroi DC, Arenzana-Seisdedos F,

Delaunay T, Naeem R, et al. Stromal fibroblasts

present in invasive human breast carcinomas pro-

mote tumor growth and angiogenesis through ele-

vated SDF-1/CXCL12 secretion. Cell. 2005;121:

335–48.

78. Marsh D, Suchak K, Moutasim KA, Vallath S, Hopper

C, Jerjes W, et al. Stromal features are predictive of

disease mortality in oral cancer patients. J Pathol.

2011;223:470–81.

79. Wu MH, Hong HC, Hong TM, Chiang WF, Jin YT,

Chen YL. Targeting galectin-1 in carcinoma-

associated fibroblasts inhibits oral squamous cell

carcinoma metastasis by downregulating MCP-1/

CCL2 expression. Clin Cancer Res. 2011;17:

1306–16.

Page 15: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

Tumor microenvironment in HNSCC 231

80. Lin J, Liu C, Ge L, Gao Q, He X, Liu Y, et al.

Carcinoma-associated fibroblasts promotes the pro-

liferation of a lingual carcinoma cell line by secreting

keratinocyte growth factor. Tumour Biol. 2011;32:

597–602.

81. Martinez-Outschoorn UE, Sotgia F, Lisanti MP. Power

surge: supporting cells “fuel” cancer cell mitochon-

dria. Cell Metab. 2012;15:4–5.

82. Curry JM, Tuluc M, Whitaker-Menezes D, Ames JA,

Anantharaman A, Butera A, et al. Cancer metabolism,

stemness and tumor recurrence: MCT1 and MCT4

are functional biomarkers of metabolic symbiosis in

head and neck cancer. Cell Cycle. 2013;12:1371–84.

83. Dvorak HF. Angiogenesis: update 2005. J Thromb

Haemost. 2005;3:1835–42.

84. Duray A, Demoulin S, Hubert P, Delvenne P, Saussez

S. Immune suppression in head and neck cancers: a

review. Clin Dev Immunol. 2010;2010:701657.

85. Whiteside TL. Immunobiology of head and neck

cancer. Cancer Metastasis Rev. 2005;24:95–105.

86. Brandwein-Gensler M, Smith RV, Wang B, Penner C,

Theilken A, Broughel D, et al. Validation of the

histologic risk model in a new cohort of patients

with head and neck squamous cell carcinoma. Am J

Surg Pathol. 2010;34:676–88.

87. Maleki S, Schlecht NF, Keller C, Diaz J, Moss J,

Prystowsky MB, et al. Lymphocytic host response

to oral squamous cell carcinoma: an adaptive T-cell

response at the tumor interface. Head Neck Pathol.

2011;5:117–22.

88. Zancope E, Costa NL, Junqueira-Kipnis AP, Valadares

MC, Silva TA, Leles CR, et al. Differential infiltration

of CD8þ and NK cells in lip and oral cavity

squamous cell carcinoma. J Oral Pathol Med. 2010;

39:162–7.

89. Ferris RL, Hunt JL, Ferrone S. Human leukocyte

antigen (HLA) class I defects in head and neck

cancer: molecular mechanisms and clinical signifi-

cance. Immunol Res. 2005;33:113–33.

90. Albers A, Abe K, Hunt J, Wang J, Lopez-Albaitero A,

Schaefer C, et al. Antitumor activity of human

papillomavirus type 16 E7-specific T cells against

virally infected squamous cell carcinoma of the head

and neck. Cancer Res. 2005;65:11146–55.

91. Lopez-Albaitero A, Nayak JV, Ogino T, Machandia A,

Gooding W, DeLeo AB, et al. Role of antigen-

processing machinery in the in vitro resistance of

squamous cell carcinoma of the head and neck cells

to recognition by CTL. J Immunol. 2006;176:3402–9.

92. Hathaway B, Landsittel DP, Gooding W, Whiteside

TL, Grandis JR, Siegfried JM, et al. Multiplexed

analysis of serum cytokines as biomarkers in squa-

mous cell carcinoma of the head and neck patients.

Laryngoscope. 2005;115:522–7.

93. Hoffmann TK, Bier H, Whiteside TL. Targeting the

immune system: novel therapeutic approaches in

squamous cell carcinoma of the head and neck.

Cancer Immunol Immunother. 2004;53:1055–67.

94. Verastegui E, Morales R, Barrera JL, Mueller A, Guz-

man B, Meneses A, et al. Immunological approach in

the evaluation of regional lymph nodes of patients

with squamous cell carcinoma of the head and neck.

Clin Immunol. 2002;102:37–47.

95. Young MR. Protective mechanisms of head and neck

squamous cell carcinomas from immune assault.

Head Neck. 2006;28:462–70.

96. Ogino T, Shigyo H, Ishii H, Katayama A, Miyokawa N,

Harabuchi Y, et al. HLA class I antigen down-

regulation in primary laryngeal squamous cell carci-

noma lesions as a poor prognostic marker. Cancer

Res. 2006;66:9281–9.

97. Grandis JR, Falkner DM, Melhem MF, Gooding WE,

Drenning SD, Morel PA. Human leukocyte antigen

class I allelic and haplotype loss in squamous cell

carcinoma of the head and neck: clinical and immu-

nogenetic consequences. Clin Cancer Res. 2000;6:

2794–802.

98. Tourkova IL, Shurin GV, Chatta GS, Perez L, Finke J,

Whiteside TL, et al. Restoration by IL-15 of MHC

class I antigen-processing machinery in human den-

dritic cells inhibited by tumor-derived gangliosides. J

Immunol. 2005;175:3045–52.

99. Katou F, Ohtani H, Watanabe Y, Nakayama T, Yoshie

O, Hashimoto K. Differing phenotypes between

intraepithelial and stromal lymphocytes in early-

stage tongue cancer. Cancer Res. 2007;67:11195–201.

100. Strome SE, Dong H, Tamura H, Voss SG, Flies DB,

Tamada K, et al. B7-H1 blockade augments adoptive

T-cell immunotherapy for squamous cell carcinoma.

Cancer Res. 2003;63:6501–5.

101. Cho YA, Yoon HJ, Lee JI, Hong SP, Hong SD.

Relationship between the expressions of PD-L1 and

tumor-infiltrating lymphocytes in oral squamous cell

carcinoma. Oral Oncol. 2011;47:1148–53.

102. Ferris RL, Whiteside TL, Ferrone S. Immune escape

associated with functional defects in antigen-

processing machinery in head and neck cancer. Clin

Cancer Res. 2006;12:3890–5.

103. Badoual C, Hans S, Merillon N, Van Ryswick C, Ravel

P, Benhamouda N, et al. PD-1-expressing tumor-

infiltrating T cells are a favorable prognostic bio-

marker in HPV-associated head and neck cancer.

Cancer Res. 2013;73:128–38.

104. Steinbrink K, Mahnke K, Grabbe S, Enk AH, Jonuleit

H. Myeloid dendritic cell: From sentinel of immunity

to key player of peripheral tolerance? Hum Immunol.

2009;70:289–93.

105. Hawiger D, Inaba K, Dorsett Y, Guo M, Mahnke K,

Rivera M, et al. Dendritic cells induce peripheral T

cell unresponsiveness under steady state conditions

in vivo. J Exp Med. 2001;194:769–79.

106. Yamazaki S, Inaba K, Tarbell KV, Steinman RM.

Dendritic cells expand antigen-specific Foxp3þCD25þ CD4þ regulatory T cells including suppres-

sors of alloreactivity. Immunol Rev. 2006;212:

314–29.

107. Jonuleit H, Schmitt E, Schuler G, Knop J, Enk AH.

Induction of interleukin 10-producing, nonproliferat-

ing CD4(þ) T cells with regulatory properties by

repetitive stimulation with allogeneic immature

human dendritic cells. J Exp Med. 2000;192:

1213–22.

Page 16: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

J.M. Curry et al232

108. Yilmaz T, Gedikoglu G, Celik A, Onerci M, Turan E.

Prognostic significance of Langerhans cell infiltration

in cancer of the larynx. Otolaryngol Head Neck Surg.

2005;132:309–16.

109. Ma CX, Jia TC, Li XR, Zhand ZF, Yiao CB. Langerhans

cells in nasopharyngeal carcinoma in relation to

prognosis. In Vivo. 1995;9:225–9.

110. Gallo O, Libonati GA, Gallina E, Fini-Storchi O,

Giannini A, Urso C, et al. Langerhans cells related to

prognosis in patients with laryngeal carcinoma. Arch

Otolaryngol Head Neck Surg. 1991;117:1007–10.

111. Mantovani A, Bottazzi B, Colotta F, Sozzani S, Ruco L.

The origin and function of tumor-associated macro-

phages. Immunol Today. 1992;13:265–70.

112. Sica A, Schioppa T, Mantovani A, Allavena P. Tumour-

associated macrophages are a distinct M2 polarised

population promoting tumour progression: potential

targets of anti-cancer therapy. Eur J Cancer. 2006;42:

717–27.

113. Mills CD, Kincaid K, Alt JM, Heilman MJ, Hill AM. M-

1/M-2 macrophages and the Th1/Th2 paradigm. J

Immunol. 2000;164:6166–73.

114. Li C, Shintani S, Terakado N, Nakashiro K, Hamakawa

H. Infiltration of tumor-associated macrophages in

human oral squamous cell carcinoma. Oncol Rep.

2002;9:1219–23.

115. El-Rouby DH. Association of macrophages with

angiogenesis in oral verrucous and squamous cell

carcinomas. J Oral Pathol Med. 2010;39:559–64.

116. Zamarron BF, Chen W. Dual roles of immune cells

and their factors in cancer development and pro-

gression. Int J Biol Sci. 2011;7:651–8.

117. Marcus B, Arenberg D, Lee J, Kleer C, Chepeha DB,

Schmalbach CE, et al. Prognostic factors in oral cavity

and oropharyngeal squamous cell carcinoma. Cancer.

2004;101:2779–87.

118. Liu SY, Chang LC, Pan LF, Hung YJ, Lee CH, Shieh YS.

Clinicopathologic significance of tumor cell-lined

vessel and microenvironment in oral squamous cell

carcinoma. Oral Oncol. 2008;44:277–85.

119. Costa NL, Valadares MC, Souza PP, Mendonca EF,

Oliveira JC, Silva TA, et al. Tumor-associated macro-

phages and the profile of inflammatory cytokines in

oral squamous cell carcinoma. Oral Oncol. 2013;49:

216–23.

120. Karin M, Greten FR. NF-kappaB: linking inflammation

and immunity to cancer development and progres-

sion. Nat Rev Immunol. 2005;5:749–59.

121. Coussens LM, Werb Z. Inflammation and cancer.

Nature. 2002;420:860–7.

122. Dumitru CA, Gholaman H, Trellakis S, Bruderek K,

Dominas N, Gu X, et al. Tumor-derived macrophage

migration inhibitory factor modulates the biology of

head and neck cancer cells via neutrophil activation.

Int J Cancer. 2011;129:859–69.

123. Galdiero MR, Garlanda C, Jaillon S, Marone G,

Mantovani A. Tumor associated macrophages and

neutrophils in tumor progression. J Cell Physiol.

2013;228:1404–12.

124. Rosenthal EL, Matrisian LM. Matrix metalloproteases

in head and neck cancer. Head Neck. 2006;28:

639–48.

125. DeClerck YA, Mercurio AM, Stack MS, Chapman HA,

Zutter MM, Muschel RJ, et al. Proteases, extracellular

matrix, and cancer: a workshop of the path B study

section. Am J Pathol. 2004;164:1131–9.

126. Gutschalk CM, Yanamandra AK, Linde N, Meides A,

Depner S, Mueller MM. GM-CSF enhances tumor

invasion by elevated MMP-2, -9, and -26 expression.

Cancer Med. 2013;2:117–29.

127. Yu B, Wei J, Qian X, Lei D, Ma Q, Liu Y. Notch1

signaling pathway participates in cancer invasion by

regulating MMPs in lingual squamous cell carcinoma.

Oncol Rep. 2012;27:547–52.

128. Rosenthal EL, Hotary K, Bradford C, Weiss SJ. Role of

membrane type 1-matrix metalloproteinase and gelat-

inase A in head and neck squamous cell carcinoma

invasion in vitro. Otolaryngol Head Neck Surg.

1999;121:337–43.

129. Sinpitaksakul SN, Pimkhaokham A, Sanchavanakit N,

Pavasant P. TGF-beta1 induced MMP-9 expression in

HNSCC cell lines via Smad/MLCK pathway. Biochem

Biophys Res Commun. 2008;371:713–8.

130. Koontongkaew S, Amornphimoltham P, Monthanpi-

sut P, Saensuk T, Leelakriangsak M. Fibroblasts and

extracellular matrix differently modulate MMP acti-

vation by primary and metastatic head and neck

cancer cells. Med Oncol. 2012;29:690–703.

131. Murphy G, Docherty AJ. The matrix metalloprotei-

nases and their inhibitors. Am J Respir Cell Mol Biol.

1992;7:120–5.

132. Price SJ, Greaves DR, Watkins H. Identification of

novel, functional genetic variants in the human

matrix metalloproteinase-2 gene: role of Sp1 in

allele-specific transcriptional regulation. J Biol Chem.

2001;276:7549–58.

133. Rosenthal EL, McCrory A, Talbert M, Carroll W, Magnu-

son JS, Peters GE. Expression of proteolytic enzymes in

head and neck cancer-associated fibroblasts. Arch Oto-

laryngol Head Neck Surg. 2004;130:943–7.

134. Lynch CC, Matrisian LM. Matrix metalloproteinases in

tumor-host cell communication. Differentiation.

2002;70:561–73.

135. Kurahara S, Shinohara M, Ikebe T, Nakamura S,

Beppu M, Hiraki A, et al. Expression of MMPS, MT-

MMP, and TIMPs in squamous cell carcinoma of the

oral cavity: correlations with tumor invasion and

metastasis. Head Neck. 1999;21:627–38.

136. Birkedal-Hansen B, Pavelic ZP, Gluckman JL, Stam-

brook P, Li YQ, MMP Stetler-Stevenson WG. and

TIMP gene expression in head and neck squamous

cell carcinomas and adjacent tissues. Oral Dis. 2000;

6:376–82.

137. Kudo Y, Iizuka S, Yoshida M, Tsunematsu T, Kondo

T, Subarnbhesaj A, et al. Matrix metalloproteinase-13

(MMP-13) directly and indirectly promotes tumor

angiogenesis. J Biol Chem. 2012;287:38716–28.

138. Quan J, Zhou C, Johnson NW, Francis G, Dahlstrom

JE, Gao J. Molecular pathways involved in crosstalk

between cancer cells, osteoblasts and osteoclasts in

the invasion of bone by oral squamous cell carci-

noma. Pathology. 2012;44:221–7.

139. Zhang Z, Filho MS, Nor JE. The biology of head and

neck cancer stem cells. Oral Oncol. 2012;48:1–9.

Page 17: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

Tumor microenvironment in HNSCC 233

140. Lai WW, Hsu SC, Chueh FS, Chen YY, Yang JS, Lin JP,

et al. Quercetin inhibits migration and invasion of

SAS human oral cancer cells through inhibition of NF-

kappaB and matrix metalloproteinase-2/-9 signaling

pathways. Anticancer Res. 2013;33:1941–50.

141. Prime SS, Davies M, Pring M, Paterson IC. The role of

TGF-beta in epithelial malignancy and its relevance to

the pathogenesis of oral cancer (part II). Crit Rev

Oral Biol Med. 2004;15:337–47.

142. Paterson IC, Matthews JB, Huntley S, Robinson CM,

Fahey M, Parkinson EK, et al. Decreased expression

of TGF-beta cell surface receptors during progression

of human oral squamous cell carcinoma. J Pathol.

2001;193:458–67.

143. Cho KH, Jeong KJ, Shin SC, Kang J, Park CG, Lee HY.

STAT3 mediates TGF-beta1-induced TWIST1 expres-

sion and prostate cancer invasion. Cancer Lett.

2013;336:167–73.

144. Na YR, Lee JS, Lee SJ, Twist Seok SH Interleukin-6-

induced. and N-cadherin enhance melanoma cell

metastasis. Melanoma Res. 2013;23:434–43.

145. Whiteman EL, Liu CJ, Fearon ER, Margolis B. The

transcription factor snail represses Crumbs3 expres-

sion and disrupts apico-basal polarity complexes.

Oncogene. 2008;27:3875–9.

146. Nguyen PT, Kudo Y, Yoshida M, Iizuka S, Ogawa I,

Takata T. N-cadherin expression is correlated with

metastasis of spindle cell carcinoma of head and neck

region. J Oral Pathol Med. 2011;40:77–82.

147. Bonde AK, Tischler V, Kumar S, Soltermann A,

Schwendener RA. Intratumoral macrophages contrib-

ute to epithelial-mesenchymal transition in solid

tumors. BMC Cancer. 2012;12:35.

148. Yu Q, Stamenkovic I. Cell surface-localized matrix

metalloproteinase-9 proteolytically activates TGF-

beta and promotes tumor invasion and angiogenesis.

Genes Dev. 2000;14:163–76.

149. Folkman J. Role of angiogenesis in tumor growth and

metastasis. Semin Oncol. 2002;29:15–8.

150. El-Gazzar R, Macluskey M, Ogden GR. Evidence for

a field change effect based on angiogenesis in the

oral mucosa? A brief report Oral Oncol. 2005;41:

25–30.

151. El-Gazzar R, Macluskey M, Williams H, Ogden GR.

Vascularity and expression of vascular endothelial

growth factor in oral squamous cell carcinoma,

resection margins, and nodal metastases. Br J Oral

Maxillofac Surg. 2006;44:193–7.

152. Macluskey M, Chandrachud LM, Pazouki S, Green M,

Chisholm DM, Ogden GR, et al. Apoptosis, prolifer-

ation, and angiogenesis in oral tissues. Possible

relevance to tumour progression. J Pathol. 2000;

191:368–75.

153. Sedivy R, Beck-Mannagetta J, Haverkampf C, Battis-

tutti W, Honigschnabl S. Expression of vascular

endothelial growth factor-C correlates with the lym-

phatic microvessel density and the nodal status in

oral squamous cell cancer. J Oral Pathol Med. 2003;

32:455–60.

154. Williams MD. Integration of biomarkers including

molecular targeted therapies in head and neck can-

cer. Head Neck Pathol. 2010;4:62–9.

155. Perez-Sayans M, Somoza-Martin JM, Barros-Angueira

F, Diz PG, Rey JM, Garcia-Garcia A. Multidrug resist-

ance in oral squamous cell carcinoma: The role of

vacuolar ATPases. Cancer Lett. 2010;295:135–43.

156. Perez-Sayans M, Somoza-Martin JM, Barros-Angueira

F, Reboiras-Lopez MD, Gandara Rey JM, Garcia-Garcia

A. Genetic and molecular alterations associated with

oral squamous cell cancer (Review). Oncol Rep.

2009;22:1277–82.

157. Moulder JE, Rockwell S. Tumor hypoxia: its impact

on cancer therapy. Cancer Metastasis Rev. 1987;5:

313–41.

158. De Palma M, Venneri MA, Galli R, Sergi Sergi L, Politi

LS, Sampaolesi M, et al. Tie2 identifies a hematopoi-

etic lineage of proangiogenic monocytes required for

tumor vessel formation and a mesenchymal popula-

tion of pericyte progenitors. Cancer Cell. 2005;8:

211–26.

159. Tse GM, Chan AW, Yu KH, King AD, Wong KT, Chen

GG, et al. Strong immunohistochemical expression of

vascular endothelial growth factor predicts overall

survival in head and neck squamous cell carcinoma.

Ann Surg Oncol. 2007;14:3558–65.

160. Karl E, Zhang Z, Dong Z, Neiva KG, Soengas MS,

Koch AE, et al. Unidirectional crosstalk between

Bcl-xL and Bcl-2 enhances the angiogenic phenotype

of endothelial cells. Cell Death Differ. 2007;14:

1657–66.

161. Neiva KG, Zhang Z, Miyazawa M, Warner KA, Karl E,

Nor JE. Cross talk initiated by endothelial cells

enhances migration and inhibits anoikis of squamous

cell carcinoma cells through STAT3/Akt/ERK signal-

ing. Neoplasia. 2009;11:583–93.

162. Scapini P, Morini M, Tecchio C, Minghelli S, Di Carlo

E, Tanghetti E, et al. CXCL1/macrophage inflamma-

tory protein-2-induced angiogenesis in vivo is medi-

ated by neutrophil-derived vascular endothelial

growth factor-A. J Immunol. 2004;172:5034–40.

163. Zha S, Yegnasubramanian V, Nelson WG, Isaacs WB,

De Marzo AM. Cyclooxygenases in cancer: progress

and perspective. Cancer Lett. 2004;215:1–20.

164. Camacho M, Leon X, Fernandez-Figueras MT, Quer

M, Vila L. Prostaglandin E(2) pathway in head and

neck squamous cell carcinoma. Head Neck. 2008;30:

1175–81.

165. Li JZ, Gao W, Chan JY, Ho WK, Wong TS. Hypoxia in

head and neck squamous cell carcinoma. ISRN

Otolaryngol. 2012;2012:708974.

166. Roh JL, Cho KJ, Kwon GY, Ryu CH, Chang HW, Choi

SH, et al. The prognostic value of hypoxia markers in

T2-staged oral tongue cancer. Oral Oncol. 2009;45:

63–8.

167. Feron O. Pyruvate into lactate and back: from the

Warburg effect to symbiotic energy fuel exchange in

cancer cells. Radiother Oncol. 2009;92:329–33.

168. Liang X, Yang D, Hu J, Hao X, Gao J, Mao Z. Hypoxia

inducible factor-alpha expression correlates with

vascular endothelial growth factor-C expression and

lymphangiogenesis/angiogenesis in oral squamous

cell carcinoma. Anticancer Res. 2008;28:1659–66.

169. Mohamed KM, Le A, Duong H, Wu Y, Zhang Q,

Messadi DV. Correlation between VEGF and HIF-

Page 18: Tumor Microenvironment in Head and Neck Squamous Cell ... · Col15-binding integrins, PGE2 Glycolytic: (MCT4 ...

J.M. Curry et al234

1alpha expression in human oral squamous cell

carcinoma. Exp Mol Pathol. 2004;76:143–52.

170. Teppo S, Sundquist E, Vered M, Holappa H, Parkki-

senniemi J, Rinaldi T, et al. The hypoxic tumor

microenvironment regulates invasion of aggressive

oral carcinoma cells. Exp Cell Res. 2013;319:376–89.

171. Shih YH, Chang KW, Chen MY, Yu CC, Lin DJ, Hsia

SM, et al. Lysyl oxidase and enhancement of cell

proliferation and angiogenesis in oral squamous cell

carcinoma. Head Neck. 2013;35:250–6.

172. Erler JT, Bennewith KL, Nicolau M, Dornhofer N,

Kong C, Le QT, et al. Lysyl oxidase is essential for

hypoxia-induced metastasis. Nature. 2006;440:1222–6.

173. Zhang Q, Tang X, Lu Q, Zhang Z, Rao J, Le AD. Green

tea extract and (-)-epigallocatechin-3-gallate inhibit

hypoxia- and serum-induced HIF-1alpha protein accu-

mulation and VEGF expression in human cervical

carcinoma and hepatoma cells. Mol Cancer Ther.

2006;5:1227–38.

174. Fath DM, Kong X, Liang D, Lin Z, Chou A, Jiang Y,

et al. Histone deacetylase inhibitors repress the

transactivation potential of hypoxia-inducible factors

independently of direct acetylation of HIF-alpha. J

Biol Chem. 2006;281:13612–9.

175. Kong X, Lin Z, Liang D, Fath D, Sang N, Caro J.

Histone deacetylase inhibitors induce VHL and

ubiquitin-independent proteasomal degradation of

hypoxia-inducible factor 1alpha. Mol Cell Biol.

2006;26:2019–28.

176. Zhang Q, Tang X, Lu QY, Zhang ZF, Brown J, Le AD.

Resveratrol inhibits hypoxia-induced accumulation of

hypoxia-inducible factor-1alpha and VEGF expres-

sion in human tongue squamous cell carcinoma and

hepatoma cells. Mol Cancer Ther. 2005;4:1465–74.

177. Warburg O. On the origin of cancer cells. Science.

1956;123:309–14.

178. Lunt SY, Vander Heiden MG. Aerobic glycolysis:

meeting the metabolic requirements of cell prolifer-

ation. Annu Rev Cell Dev Biol. 2011;27:441–64.

179. Sonveaux P, Vegran F, Schroeder T, Wergin MC,

Verrax J, Rabbani ZN, et al. Targeting lactate-fueled

respiration selectively kills hypoxic tumor cells in

mice. J Clin Invest. 2008;118:3930–42.

180. Zu XL, Guppy M. Cancer metabolism: facts, fantasy,

and fiction. Biochem Biophys Res Commun. 2004;

313:459–65.

181. Moreno-Sanchez R, Rodriguez-Enriquez S, Marin-

Hernandez A, Saavedra E. Energy metabolism in

tumor cells. Febs J. 2007;274:1393–418.

182. Guppy M, Greiner E, Brand K. The role of the

Crabtree effect and an endogenous fuel in the energy

metabolism of resting and proliferating thymocytes.

Eur J Biochem. 1993;212:95–9.

183. Pfeiffer T, Schuster S, Bonhoeffer S. Cooperation and

competition in the evolution of ATP-producing path-

ways. Science. 2001;292:504–7.

184. Busk M, Walenta S, Mueller-Klieser W, Steiniche T,

Jakobsen S, Horsman MR, et al. Inhibition of tumor

lactate oxidation: consequences for the tumor micro-

environment. Radiother Oncol. 2011;99:404–11.

185. Pavlides S, Whitaker-Menezes D, Castello-Cros R,

Flomenberg N, Witkiewicz AK, Frank PG, et al. The

reverse Warburg effect: aerobic glycolysis in cancer

associated fibroblasts and the tumor stroma. Cell

Cycle. 2009;8:3984–4001.

186. Tripathi P, Kamarajan P, Somashekar BS, MacKinnon

N, Chinnaiyan AM, Kapila YL, et al. Delineating

metabolic signatures of head and neck squamous

cell carcinoma: phospholipase A2, a potential ther-

apeutic target. Int J Biochem Cell Biol. 2012;44:

1852–61.

187. Sandulache VC, Ow TJ, Pickering CR, Frederick MJ,

Zhou G, Fokt I, et al. Glucose, not glutamine, is the

dominant energy source required for proliferation

and survival of head and neck squamous carcinoma

cells. Cancer. 2011;117:2926–38.

188. Walenta S, Salameh A, Lyng H, Evensen JF, Mitze M,

Rofstad EK, et al. Correlation of high lactate levels in

head and neck tumors with incidence of metastasis.

Am J Pathol. 1997;150:409–15.

189. Brizel DM, Schroeder T, Scher RL, Walenta S, Clough

RW, Dewhirst MW, et al. Elevated tumor lactate

concentrations predict for an increased risk of meta-

stases in head-and-neck cancer. Int J Radiat Oncol

Biol Phys. 2001;51:349–53.

190. Brennan PA, Mackenzie N, Quintero M. Hypoxia-

inducible factor 1alpha in oral cancer. J Oral Pathol

Med. 2005;34:385–9.

191. Zhang Q, Zhang ZF, Rao JY, Sato JD, Brown J,

Messadi DV, et al. Treatment with siRNA and anti-

sense oligonucleotides targeted to HIF-1alpha

induced apoptosis in human tongue squamous cell

carcinomas. Int J Cancer. 2004;111:849–57.

192. Li X, Fan Z. The epidermal growth factor receptor

antibody cetuximab induces autophagy in cancer

cells by downregulating HIF-1alpha and Bcl-2 and

activating the beclin 1/hVps34 complex. Cancer Res.

2010;70:5942–52.

193. Kato K, Gong J, Iwama H, Kitanaka A, Tani J, Miyoshi

H, et al. The antidiabetic drug metformin inhibits

gastric cancer cell proliferation in vitro and in vivo.

Mol Cancer Ther. 2012;11:549–60.

194. Bao B, Wang Z, Ali S, Ahmad A, Azmi AS, Sarkar SH,

et al. Metformin inhibits cell proliferation, migration

and invasion by attenuating CSC function mediated

by deregulating miRNAs in pancreatic cancer cells.

Cancer Prev Res (Phila). 2012;5:355–64.

195. Klubo-Gwiezdzinska J, Jensen K, Costello J, Patel A,

Hoperia V, Bauer A, et al. Metformin inhibits growth

and decreases resistance to anoikis in medullary

thyroid cancer cells. Endocr Relat Cancer. 2012;19:

447–56.

196. Zhuang Y, Miskimins WK. Metformin induces both

caspase-dependent and poly(ADP-ribose) polymerase-

dependent cell death in breast cancer cells. Mol

Cancer Res. 2010;9:603–15.

197. Noto H, Goto A, Tsujimoto T, Noda M. Cancer risk in

diabetic patients treated with metformin: a system-

atic review and meta-analysis. PLoS One. 2012;

7:e33411.

198. Luo Q, Hu D, Hu S, Yan M, Sun Z, Chen F. In vitro

and in vivo anti-tumor effect of metformin as a novel

therapeutic agent in human oral squamous cell

carcinoma. BMC Cancer. 2012;12:517.