role of the microenvironment in liver metastasis: from pre ......review role of the microenvironment...

13
Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver metastases remain a major barrier to successful manage- ment of malignant disease, particularly for cancers of the gastro- intestinal tract but also for other malignancies, such as breast carcinoma and melanoma. The ability of metastatic cells to survive and proliferate in the liver is determined by the outcome of complex, reciprocal interactions between tumor cells and different local resident subpopulations, including the sinusoidal endothelium, stellate, Kupffer, and inammatory cells that are mediated through cellcell and cellextracellular matrix adhesion and the release of soluble factors. Cross-communication between different hepatic resident cells in response to local tissue damage and inammation and the recruitment of bone marrow cells further enhance this intercellular communication network. Both resident and recruited cells can play opposing roles in the pro- gression of metastasis, and the balance of these divergent effects determines whether the tumor cells will die, proliferate, and colonize the new site or enter a state of dormancy. Moreover, this delicate balance can be tilted in favor of metastasis, if factors produced by the primary tumor precondition the microenviron- ment to form niches of activated resident cells that promote tumor expansion. This review aims to summarize current knowledge on these diverse interactions and the impact they can have on the clinical management of hepatic metastases. Clin Cancer Res; 22(24); 597182. Ó2016 AACR. Introduction Cancer metastasis remains the major challenge to successful management of malignant disease. The liver is the main site of metastatic disease and a major cause of death from gastrointes- tinal malignancies, such as colon, gastric, and pancreatic carci- nomas as well as melanoma, breast cancer, and sarcomas (1). Circulating metastatic cells that enter the liver encounter unique cellular populations. These include the parenchymal hepatocytes and the nonparenchymal hepatocytes, including liver sinusoidal endothelial cells (LSEC), hepatic stellate cells (HSC), Kupffer cells (KC), dendritic cells, liver-associated lymphocytes, and portal broblasts. Circulating and bone marrowderived immune cells are also recruited to the liver in response to, and possibly in preparation for, invading tumor cells and can affect the nal outcome (reviewed in refs. 2, 3; see Table 1). This review summarizes our current understanding of the role of the liver microenvironment in the growth of hepatic metastases and the reciprocal interactions between metastatic tumor cells and different liver cell populations that regulate the process. Pre- and Prometastatic Niches Drive the Progression of Liver Metastasis The process of liver metastasis has previously been divided into four major phases (detailed in Table 1) that follow tumor cell entry into the liver (4, 5). Emerging evidence suggests that in addition, a premetastatic phase could set the stage for liver colonization by disseminating tumor cells. Evidence for premetastatic niche formation in the liver The term "premetastatic niche" was coined to describe a microenvironment in a secondary organ site that has been rendered permissive to metastatic outgrowth in advance of cancer cell entry through the activity of circulating factors released by the primary tumor (68). Although the dependency of metastasis on premetastatic niches remains controversial and difcult to verify in the clinical setting (9, 10), it appears that a consensus is emerging on two scores: (i) that their existence may have important implications for the clinical management of metastatic disease, and (ii) that much can be learned from their interrogation about the cellular/molecular changes required to facilitate tumor cell growth in a distant site, such as the liver. Two recent studies have conrmed their potential relevance to liver metastasis. In a mouse model of metastatic pancreatic ductal adenocarcinoma (PDAC), Costa-Silva and colleagues showed that PDAC-derived exosomes taken up by hepatic KCs, upregulate TGFb production, leading to increased bronectin production by HSCs and recruitment of bone marrowderived macrophages. They identied macrophage migration inhibitory factor (MIF) as essential for premetastatic niche formation and metastasis. Moreover, in exosomes derived from patients with stage I PDAC who later developed liver metastasis, MIF levels were found to be higher than in patients whose tumors did not progress, thus identifying MIF as a potential predictor of PDAC liver metastasis (11). This group also showed that exosomes from human breast and pancreatic cancer cell lines that metas- tasize selectively to the lung (MDA-MB-231) or liver (BxPC-3 and HPAF-II) fused preferentially with broblasts and epithelial cells in the lung and KCs in the liver, and this was mediated by Departments of Surgery, Medicine, and Oncology, McGill University and the McGill University Health Centre, Montreal, Quebec, Canada. Corresponding Author: Pnina Brodt, McGill University Health Centre, 1001 Boulevard D ecarie, Montr eal, Quebec H4A 3J1, Canada. Phone: 514-934-1934, ext. 36692; Fax: 514-938-7033; E-mail: [email protected] doi: 10.1158/1078-0432.CCR-16-0460 Ó2016 American Association for Cancer Research. Clinical Cancer Research www.aacrjournals.org 5971 on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Upload: others

Post on 22-Mar-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

Review

Role of the Microenvironment in Liver Metastasis:From Pre- to Prometastatic NichesPnina Brodt

Abstract

Liver metastases remain a major barrier to successful manage-ment of malignant disease, particularly for cancers of the gastro-intestinal tract but also for other malignancies, such as breastcarcinoma and melanoma. The ability of metastatic cells tosurvive and proliferate in the liver is determined by the outcomeof complex, reciprocal interactions between tumor cells anddifferent local resident subpopulations, including the sinusoidalendothelium, stellate, Kupffer, and inflammatory cells that aremediated through cell–cell and cell–extracellularmatrix adhesionand the release of soluble factors. Cross-communication betweendifferent hepatic resident cells in response to local tissue damageand inflammation and the recruitment of bone marrow cells

further enhance this intercellular communication network. Bothresident and recruited cells can play opposing roles in the pro-gression of metastasis, and the balance of these divergent effectsdetermines whether the tumor cells will die, proliferate, andcolonize the new site or enter a state of dormancy. Moreover,this delicate balance can be tilted in favor of metastasis, if factorsproduced by the primary tumor precondition the microenviron-ment to formniches of activated resident cells that promote tumorexpansion. This review aims to summarize current knowledge onthese diverse interactions and the impact they can have on theclinical management of hepatic metastases. Clin Cancer Res; 22(24);5971–82. �2016 AACR.

IntroductionCancer metastasis remains the major challenge to successful

management of malignant disease. The liver is the main site ofmetastatic disease and a major cause of death from gastrointes-tinal malignancies, such as colon, gastric, and pancreatic carci-nomas as well as melanoma, breast cancer, and sarcomas (1).

Circulating metastatic cells that enter the liver encounterunique cellular populations. These include the parenchymalhepatocytes and the nonparenchymal hepatocytes, including liversinusoidal endothelial cells (LSEC), hepatic stellate cells (HSC),Kupffer cells (KC), dendritic cells, liver-associated lymphocytes,and portal fibroblasts. Circulating and bone marrow–derivedimmune cells are also recruited to the liver in response to, andpossibly in preparation for, invading tumor cells and can affect thefinal outcome (reviewed in refs. 2, 3; see Table 1).

This review summarizes our current understanding of the roleof the livermicroenvironment in the growth of hepaticmetastasesand the reciprocal interactions betweenmetastatic tumor cells anddifferent liver cell populations that regulate the process.

Pre- and Prometastatic Niches Drive theProgression of Liver Metastasis

The process of livermetastasis has previously been divided intofour major phases (detailed in Table 1) that follow tumor cell

entry into the liver (4, 5). Emerging evidence suggests that inaddition, a premetastatic phase could set the stage for livercolonization by disseminating tumor cells.

Evidence for premetastatic niche formation in the liverThe term "premetastatic niche" was coined to describe a

microenvironment in a secondary organ site that has beenrendered permissive to metastatic outgrowth in advance ofcancer cell entry through the activity of circulating factorsreleased by the primary tumor (6–8). Although the dependencyof metastasis on premetastatic niches remains controversial anddifficult to verify in the clinical setting (9, 10), it appears that aconsensus is emerging on two scores: (i) that their existence mayhave important implications for the clinical management ofmetastatic disease, and (ii) that much can be learned from theirinterrogation about the cellular/molecular changes required tofacilitate tumor cell growth in a distant site, such as the liver.Two recent studies have confirmed their potential relevance toliver metastasis. In a mouse model of metastatic pancreaticductal adenocarcinoma (PDAC), Costa-Silva and colleaguesshowed that PDAC-derived exosomes taken up by hepatic KCs,upregulate TGFb production, leading to increased fibronectinproduction by HSCs and recruitment of bone marrow–derivedmacrophages. They identified macrophage migration inhibitoryfactor (MIF) as essential for premetastatic niche formation andmetastasis. Moreover, in exosomes derived from patients withstage I PDAC who later developed liver metastasis, MIF levelswere found to be higher than in patients whose tumors did notprogress, thus identifying MIF as a potential predictor of PDACliver metastasis (11). This group also showed that exosomesfrom human breast and pancreatic cancer cell lines that metas-tasize selectively to the lung (MDA-MB-231) or liver (BxPC-3and HPAF-II) fused preferentially with fibroblasts and epithelialcells in the lung and KCs in the liver, and this was mediated by

Departments of Surgery, Medicine, and Oncology, McGill University and theMcGill University Health Centre, Montreal, Quebec, Canada.

Corresponding Author: Pnina Brodt, McGill University Health Centre, 1001Boulevard D�ecarie, Montr�eal, Quebec H4A 3J1, Canada. Phone: 514-934-1934,ext. 36692; Fax: 514-938-7033; E-mail: [email protected]

doi: 10.1158/1078-0432.CCR-16-0460

�2016 American Association for Cancer Research.

ClinicalCancerResearch

www.aacrjournals.org 5971

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 2: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

the exosomal integrin laminin receptors a6b1 and a6b4 and thefibronectin receptor avb5, respectively. In exosome-fused KCs,the proinflammatory factors S100P and S100A8 were upregu-lated (Table 1; Fig. 1). Significantly, in PDAC patients, a corre-lation was documented between the levels of circulating, av-bearing exosomes and disease stage, suggesting that exosomesmay play a role clinically and their levels and integrin cargo maypredict metastases in specific organs (12). Kowanetz and col-leagues reported that tumor-derived GM-CSF could mobilizeLy6CþLy6Gþ myeloid cells into metastases, enabling niches inthe lung and liver, and also identified S100A8 as a driving factor(13). Tumor-derived TIMP-1 was identified as another potentialinducer of increased liver susceptibility to metastasis, acting viahepatic SDF-1 and neutrophil recruitment (14). Interestingly,S100A8 was identified as a driver of liver premetastatic nicheformation in several studies (12–15), suggesting that it mayhave clinical utility as a predictor of liver metastasis.

The stage of primary tumor development at which premeta-static niches can be formed is difficult to assess in the clinicalsetting. Costa-Silva and colleagues found exosomal MIF upre-gulation in mice with pretumoral pancreatic lesions, and highplasma exosomal MIF levels were also detected in patientswith stage I PDAC (11), suggesting that they could be formedat very early stages of cancer development. The potentialcontribution of circulating cancer cells (CTC) that may bepresent early in tumor development and even after resectionof the primary tumor (16, 17) is also unknown.

Resident and Immune Cells Can PlayDiverse and Opposing Roles in theDevelopment of Liver MetastasesHost innate resistance mechanism can destroy tumor cellsprior to extravasation

Blood-borne cancer cells entering the liver first encounter theLSECs, KCs, and hepatic natural killer (NK) cells (pit cells) thattogether mount the first line of defense (18). Cancer cellsentrapped in the sinusoids may undergo destruction due tomechanical stress and deformation-associated trauma or may diedue to phagocytosis by KCs or cytolysis by NK cell–releasedperforin/granzyme (18; reviewed in ref. 4). NO, ROS, and toxicradicals released by LSECs in response to local ischemia/reperfu-sion can contribute to tumor cell death (19, 20). Release of NOand IFNg by LSECs and NK cells can result in upregulation oftumor cell Fas and apoptosis (18) that can be augmented by NKand LSEC-derived TNFa (21). TNFa is one of the numerouscytokines and chemokines unleashed by KCs in response toinflammation (Table 1; ref. 22). In addition to causing cell deathdirectly, some of these factors can mobilize and activate addi-tional innate immune cells, such as neutrophils, thereby adding tothe local tumoricidal capacity (reviewed in ref. 23). In severalanimal tumor models, including colon cancer, loss of NK cellsincreased cancer cell growth in the liver, whereas enhanced NKactivity reduced liver metastasis (24–26). Indirect evidence alsosuggests that susceptibility to NK-mediated immune attack canaffect the ability of human cancer cells to generate livermetastases(27). Circulating neutrophils can also release various factorsabundant in their granules to kill tumor cells (see Table 1), andtheir cytokines and chemokines can activate the tumoricidalpotential of resident KCs and recruit host immune T cells withantitumorigenic activities (reviewed in ref. 28).

Cancer cells can escape these cytotoxic effects by formingclusters with blood or other cancer cells that shield them fromthe lethal effects of shear stress or NK-mediated cytotoxicity(reviewed in ref. 29). The release of inflammatory mediators,such as IL1b, TNFa, and IL18, can also initiate a cascade thatfacilitates their rapid exit from the vasculature to a less "toxic"microenvironment (see below and Fig. 2).

The sinusoidal endothelium actively participates in differentphases of metastasis

Although the rapid inflammatory response initiated by cancercell entry can lead to cell death, it may also have a tumor-protective effect by upregulating the expression of LSEC celladhesion molecules (CAM) and in this way, enhance cancercell adhesion and transendothelial migration into the space ofDisse, where they can escape the cytotoxic effects of KCs andNK cells (30). Cancer cells can adhere to inflammation-inducedE-selectin, VCAM-1, and ICAM-1, either as single cells or inassociation with host cells, such as KCs or neutrophils (reviewedin ref. 31). Blockade of this inflammatory cascade was shownto reduce liver metastasis (32–34), and TNFR1 signaling wasidentified as a major driver of this process (35).

Ultimately, the outcomeof these opposing effects on tumor cellsurvival and progression to the next phase depends on multiplefactors, including the expression on the cancer cells of the respec-tive counter receptors, namely the E-selectin ligands sLewa andsLewx and CD44 isoforms (36; reviewed in refs. 29, 37), andVCAM-1 and ICAM-1 counter receptors integrins a4b1 and LFA-1and Mac-1 (38), respectively. E-selectin binding triggers a signal-ing cascade in both tumor cells and LSECs, leading to diapedesisand transendothelial migration (39) and altering gene expressionin a tumor type–specific manner (40). Clinical studies havedocumented increased expression of E-selectin in and aroundcolorectal cancer liver metastases and elevated soluble E-selectin,ICAM-1, and VCAM-1 levels that correlated with disease outcomein patients with colorectal cancer (reviewed in ref. 23).

LSECs may have other metastasis-promoting functions. LSEC-secreted fibronectin can induce epithelial–mesenchymal transi-tion (EMT) in colorectal cancer cells via integrin a9b1, enhancingERK signaling and tumor invasion (41). Human LSECs wereshown to induce colorectal cancer migration and EMT via MIF(42), thereby increasing their metastatic potential.

LSECs contribute to tumor-induced angiogenesis. A recentstudy identified Notch1 as a negative regulator of sinusoidalendothelial cells sprouting into micrometastases and thereby ofangiogenesis and liver metastasis (43), raising questions aboutthe advisability of Notch targeting for cancer therapy. Moreover,liver metastasis of different carcinomas, including colorectalcancer, can co-opt the sinusoidal endothelium at the tumor–liverinterface, giving rise to a histologic growth pattern termed the"replacement" or "sinusoidal" growth pattern that seems to resultfrom tumor cell invasion between LSECs and the matrix in thespace of Disse and is characterized by the formation of intrameta-static vessels that appear continuous with the sinusoidal vessels(reviewed in ref. 23). The factors that drive this co-opting mech-anism remain to be identified, but a recent study by Im andcolleagues suggests that Ang2 may play a regulatory role (44).

Collectively, the evidence shows that LSECs are not a passivebarrier to tumor cell extravasation but participate actively in themetastatic process. Cancer cell interaction with LSECs can recip-rocally alter thephenotypes of both cell types, and thismay lead to

Brodt

Clin Cancer Res; 22(24) December 15, 2016 Clinical Cancer Research5972

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 3: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

intravascular tumor cell destruction but can also promote metas-tasis through enhanced tumor cell migration and increasedangiogenesis (Table 1; Figs. 2 and 3).

The role of macrophages is context dependentKCs constitute approximately 10% of all liver cells. Recent

studies suggest that they may be established prenatally andmaintained into adulthood, independently of replenishmentfrom bloodmonocytes but in anM-CSF andGM-CSF–dependentmanner (45). They reside mainly in the hepatic sinusoids andare anchored to the endothelium by long cytoplasmic processes(46; reviewed in ref. 47). In their interaction with invading can-cer cells, KCs can play diverse and opposing roles, dependingon factors such as the stage of the metastatic process, tumorload, and interactions with other immune cells. As discussed,KCs can promote metastasis by orchestrating the premetastaticniche. However, tumor–KC interactions can have opposingeffects. Several studies have documented a rapid adhesion oftumor cells to KCs in the sinusoidal lumen, resulting mainly in

tumor cell phagocytosis or apoptosis, within hours of tumor cellentry (47). This tumoricidal effect may require cooperation withlocal or recruited NK cells (48), and its efficiency in removingcancer cells likely depends on the tumor load (49). An intravitalmicroscopy study (50) recently revealed that 74% of intra-arte-rially injected rat colon cancer cells adhered to the KCs within6 hours, and this was associated with increased liver TNFa andIL1b levels, consistent with our own findings (51, 52) andindicative of KC activation. Extensive phagocytosis of tumorcells was observed within 2 hours postinjection. Interestingly,although elimination of KCs 2 days prior to tumor injectionincreased liver metastasis, it had no significant effect up to 1 weekafter tumor injection, suggesting that KCs exerted their mostpotent antitumor effect within the first 24 hours of tumor cellentry into the liver (50). A bimodal effect of KCdepletionwas alsodocumented in another study of murine colorectal cancer, and itwas attributed to decreased VEGF and increased iNOS levels inlivers subjected to late-stage KC depletion (53). Indeed, tumorcells that survive the initial tumoricidal assault by KCs can benefit

Table 1. Cells and mediators involved in the different phases of liver metastasis

Soluble factors involvedPhase of the metastaticprocess Cell type involved Cell surface markers Recruiting or activating Produced References

Premetastatic nicheKCs CD11b, F4/80 M-CSF, GM-CSF, MIF TGFb, S100P, S100A8 (11–14)HSCs a-SMA, desmin, GFAP KC-derived TGFb and TNFa Fibronectin (11, 13)MDSCs CD11b, Ly6C, Ly6G Tumor-derived GM-CSF S100A8 (12, 15)Neutrophils CD11b, Ly6G TIMP-1, SDF-1 S100A8 (7, 14, 67)

Post-tumor invasion(1) Microvascular phase:

Tumor cells trapped inthe vasculature

LSECs CD31, TNFR1, TNFR2, avb3,E-selectin, VCAM-1, ICAM-1

TNFa, IL1b, IL18 ROS, NO, IFNg , TNFa (5, 18–20, 23, 31)

KCs CD11b, F4/80 M-CSF, GM-CSF, IFNg TNFa, IL1, IL6, IL8, IFNg ,MIP-1a, MIP-2, IP-10,MCP-1, CCL5, VEGF

(5, 23, 31, 45–47, 51)

Neutrophils CD11b, Ly6G S100A8, S100A9, IL8, CCL2 TNFa, ROS, defensins,perforins, elastase

(28, 31, 64)

Hepatic NK cells NK1.1, NK1.2, FcgRIII CXCL9, CXCL10, IL1, IL12,IL15, IL18

Grz, Perforin, IFNg ,TNFa

(5, 31, 48)

(2) Extravascular,preangiogenic phase:Tumor cells transmigrateinto the space of Disse,activating a local stromalresponse

HSCs a-SMA, desmin, GFAP KC-derived TGFb, PDGF,bFGF, SDF-1, IGF-I, CCL2

ECM deposition, MCP-1,CCL5, CCL21, TGFb

(74–77, 84)

Neutrophils CD11b, Ly6G; N1: ICAM-1;N2: CXCR4

S100A8, S100A9, CXCL1,CXCL2, CXCL5, TNFa,IL8, G-CSF, IFNg , CCL2,CCL5, TGFb

TNFa, ROS, NOS,MMP-8, MMP-9,elastase, cathepsin G,VEGF

(28, 60, 64, 71)

KCs CD11b, F4/80 M-CSF, GM-CSF, IFNg IL6, VEGF, HGF,MMP-9, MMP-14

(5, 50, 53)

(3) Angiogenic phase:Micrometastases arevascularized

Blood-derivedmonocytes

CD11b, F4/80, CD68, CD163,CD206

TNFa, MCP-1 VEGF, MMPs, bFGF (57, 60)

HSCs a-SMA, desmin, GFAP KC-derived TGFb, PDGF,bFGF, SDF-1, IGF-I, CCL2

VEGF, angiopoietin-1,MMP-2, -9, and -13

(74, 79–82, 88)

LSECs CD31, TNFR1, ICAM-1 VEGF, Ang1, inhibited byNotch1 and regulated byAng2

Fibronectin, MIF (23, 41–44)

Macrophages (M2) CD11b, F4/80, Ly6C, CD163 TGFb, IL10 NO, TNFa, IFN, VEGF,EGF, bFGF, TGFb,IL10

(57–61)

(4) Growth phase:Metastases expansion

Hepatocytes Albumin, tyrosineaminotransferase

Adhesion to tight junctionintegral proteins

AREG, EGF, HB-EGF,ErB2, IGF-I, HGFL,ErB3, bFGF

(92–98, 101–105,107)

Tregs CD4, CD25, Foxp3 TGFb, IL10, CCL5, CCL22 VEGF, TGFb, IL10 (35, 111, 124)

NOTE: Shown are the four phases of the metastatic process that follow tumor cell invasion into the liver. Listed are the hepatic cells known to be involvedat each phase, the mediators regulating their recruitment and activation, and the factors they produce to block or promote the metastatic process. Theprocess has been simplified and divided into distinct phases in the interest of clarity. However, it should be viewed as dynamic with overlapping cellular andmolecular drivers at different phases. Moreover, as the metastatic process advances, the cancer and hepatic cells evolve and their phenotypes change as aconsequence of their interactions. These newly acquired properties help propel the process forward (see also Figs. 1–3).Abbreviations: ECM, extracellular matrix; MDSC, myeloid-derived suppressor cell; MMP, matrix metalloproteinase, NK, natural killer; Treg, regulatory T cell.

The Microenvironment in Liver Metastasis

www.aacrjournals.org Clin Cancer Res; 22(24) December 15, 2016 5973

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 4: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

from their protumorigenic functions. Adhesion to KC can facil-itate tumor cell extravasation, among others, by sequential acti-vation of endothelial CAMs (29, 54). In addition, KCs canproduce cytokines and growth factors, including IL6, hepatocytegrowth factor (HGF), and VEGF, and matrix metalloproteinases(MMP), such as MMP-9 and MMP-14, that can accelerate tumorcell invasion into and within the parenchymal space as well aspromote tumor cell proliferation and angiogenesis, therebyenhancing liver metastasis (see Table 1; Figs. 2 and 3).

Taken together, the evidence suggests that KC targeting couldbecome an effective antimetastatic strategy only within a verynarrow window. Limiting KC activity may be beneficial if it couldprevent premetastatic niche formation,whereas enhancing theKCtumoricidal potential using agents such as IFNg , GM-CSF, or

muramyl dipeptide may be most effective if achieved beforetumor cells enter the liver in large numbers (47, 55). The contri-bution of CTCs that may be present before liver micrometastasesare established or after surgical resection of the primary tumor(16, 17) to the tumor load and KC activation is unknown. Thismultifaceted role of KCs in liver metastasis may explain thelimited success to date of KC-targeting interventions (47).

Importantly, in addition to KCs, blood-derived monocytescan also be recruited to the tumor site in response to localinjury and inflammation and differentiate locally into matureCD11bloF4/80hi macrophages in a CCR2-dependent manner(56). These macrophages can also trigger HSC activation andfibrogenesis (57), a process important in the early stages ofextravascular tumor expansion (see below).

© 2016 American Association for Cancer Research

A B

ExosomesG-CSF

PDAC

LSEC

BM-derivedmyeloid cells

Vessel lumen

Space of Disse

KC

HSC

Hepatocyte

Fibronectin

S100A8 S100A8

Potential therapeutic strategies:1. Enhancing KC tumoricidal activities2. Inducing HSC apoptosis3. Inhibition of inflammatory mediators

Ly6C+Ly6G+

myeloid cells

Neutrophil32

1

aHSC

MIF

?

TGFβ

TGFβ

TGFβ

Tumor-derivedexosomes

Tumor-derivedexosomes

S100A8

Tumor-derived G-CSF

aKC

αβ

Figure 1.

The premetastatic niche in the liver. Shown is a diagrammatic representation of the multistep process involved in the formation of premetastatic niches inthe liver based on data described in refs. 11, 12. The symbols used for depiction of different hepatic cells are listed in Fig. 3. A and B, A diagrammaticrepresentation of a PDAC is shown in inset A, and an enlarged image of a Kupffer cell activated by PDAC-derived exosomes is shown in inset B.Encircled numbers denote cellular and molecular interactions that can potentially be targeted for liver metastasis prevention, as detailed in the right-hand corner box. BM, bone marrow.

Brodt

Clin Cancer Res; 22(24) December 15, 2016 Clinical Cancer Research5974

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 5: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

Macrophages have inherent plasticity, and their phenotypescan change within a spectrum of activation states between thepolar M1 and M2 phenotypes (58, 59). M1 macrophages aretumoricidal due to high NO and TNFa production levels andproduce Th1, Th17, and the NK-attracting chemokines CXCL9and CXCL10 (60). In contrast, M2 macrophages can promotetumor growth through production of growth factors, such asVEGF, EGF, bFGF, and TGFb (61). In addition, M1 macrophagesactivate Th1-type immune responses that can further amplifyM1/killer-type activity through the production of IFNg , whereasM2 macrophages induce regulatory T cells (Treg) and thereby animmunotolerant microenvironment through release of TGFb

and IL10 (Table 1; ref. 62). Evidence regarding the role ofmacrophage polarization in liver metastasis is scant, becausemacrophages within or around hepatic metastases have, with fewexceptions, not been subtyped. A recent clinical study identifiedthe M2/M1 ratio in hepatic resections as a correlate of colorectalcancer metastasis (63), implicating M2 macrophages in the clin-ical disease. Furthermore, F4/80, the cell surface marker frequent-ly used to identify KCs, is also expressed on recruited monocytes,and strategies used to eliminate macrophages in vivo are not KCspecific. The source and identities of macrophages in manypublished studies remain therefore to be verified. As immuneediting of the tumor microenvironment and the conversion of

III. Angiogenic phase

Fas/FasL

Fas/FasL

VEGF

VEGF-R

MMP-9

MMP-9VEGF

LSEC

aKC

I. Microvascular phase

II. Preangiogenic phase

Potential therapeutic strategies:3. TGFbRII blockade4. MDSC depletion5. TNFR2 blockade6. PD-1/PD-L1 blockade

Potential therapeutic strategies:10. IGF-I neutralization11. Claudin-2 blockade

IV. Growth phase

Potential therapeutic strategies:1. TNF inhibition2. E-selectin blockade

Potential therapeutic strategies:7. Blocking M2 polarization8. VEGF inhibition9. N1 induction

sLew×/E-Sel.

TNFα

TNFα

TNFα

TGFβ

TGFβ

TGFβ

IL10TGFβ

NOTNFα

TNFR1

HSC

aHSC

aHSC

M1

N1

N2

IGF-I Hepatocyte

M2

CoIIVMDSC

ArgROS

Treg MMPsVEGFTGFβIGF

CD8+

T cell

Tumor cell

Vessel lumen

Space of Disse

Parenchyma

12

3

4

56

7

8

9

11

10

© 2016 American Association for Cancer Research

Figure 2.

The multistep process of liver colonization by disseminated cancer cells. Shown is a diagrammatic representation of the major cell types, soluble factors, andextracellular matrix (ECM) components in the liver microenvironment that impact the progression of liver metastasis. The four major phases in the process aredelineated by dashed lines. Top, major intercellular interactions that occur upon entry of the tumor cells into the sinusoidal vessels (the microvascular phase)and precede tumor extravasation; middle, events following tumor cell extravasation into the space of Disse, divided into the preangiogenic phase (middle left)that is orchestrated by stellate cell activation, ECM deposition, cytokine production, and the recruitment of innate and adaptive immune cells and culminatesin neovascularization (the vascular phase; middle right); bottom, these interactions enable tumor expansion (the growth phase). The symbols used fordepiction of different hepatic cells are listed in Fig. 3 or indicated in the diagram. The encircled numbers denote potential targets for therapeuticintervention, as detailed in the boxes inserted within each of the depicted phases. MDSC, myeloid-derived suppressor cell; MMP, matrix metalloproteinase;Treg, regulatory T cell.

The Microenvironment in Liver Metastasis

www.aacrjournals.org Clin Cancer Res; 22(24) December 15, 2016 5975

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 6: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

Produce ROS, NO, IFNγ,TNFα

Express CAM, protect tumorcells from apoptosis, facilitatetransmigration, participate inangiogenesis, induce EMT

Participate in premetastaticniche formation, drive fibrogenicresponse, ECM production, mediate angiogenesis,recruit inflammatory cells,produce MMPs

Mediate tumor adhesion, producegrowth factors (e.g., IGF-I, HGFL),alter tumor transcriptome, undergoEMT to provide stromal support,enhance angiogenesis via bFGF

(5, 11, 12, 47–53)

(59–63)

(113, 117, 123, 124)

Produce perforin and TNFα,induce Fas/FasL interaction,kill tumor cells

Induce Tregs, block T-cellexpansion, suppress cytotoxicT-cell function, produce ROS,produce arginase

M2 macrophages produceVEGF, EGF, and bFGF; induce Tregs via IL10 and TGFβ

M1 macrophages produce NO,TNFα, and T-cell– and NK-recruiting chemokines

Participate in premetastaticniche formation; activate LSECCAM; produce VEGF, IL6,HGF, and MMP-9; activate HSC

Phagocytose tumor cells, causeapoptosis (early in the process),produce TNFα, mobilizeneutrophils and NK cells

(5, 18–20, 31, 33, 41)

Hepatocytes

Antimetastatic effects ReferencesPrometastatic effectsCell type

Liver sinusoidendothelial cells(LSEC)

Hepatic stellate cells (HSC)

Kupffer cells(KC)

Macrophages

Myeloid-derivedsuppressor cells(MDSC)

Natural killer (NK) cells

T lymphocytes

Neutrophils

(11, 23, 77, 82–86)

Resting Activated

(93–96, 101, 103, 105)

(4)

(28, 64, 77)

(109, 111)Tregs inhibit effector T-cellexpansion, eliminate cytotoxicT cells, subvert immuneactivation, produce VEGF

Cytotoxic T cells kill tumorcells, release perforin andTNFα, recruit macrophages

Participate in premetastaticniche formation, enhancetransendothelial migration,entrap tumor cells; N2 neutrophilsproduce VEGF, MMP-9, and CCL5

N1 neutrophils produce ROS,NO, TNFα, and defensins; recruitCD8+ T cells

© 2016 American Association for Cancer Research

Figure 3.

Parenchymal, nonparenchymal, and immune cells of the liver and their role in metastasis. Listed are the parenchymal and nonparenchymal cells of the liver andtheir anti- and prometastatic functions. Symbols shown for each cell type were used to depict intercellular communications in Figs. 1 and 2. EMT, epithelial-mesenchymal transition; Treg, regulatory T cell.

Brodt

Clin Cancer Res; 22(24) December 15, 2016 Clinical Cancer Research5976

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 7: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

tumor-associated (TAM) M2 macrophages to M1 killer macro-phages are emerging as potentially relevant anticancer strategies(62), further characterization of both the source and the statusof macrophage populations involved in liver metastasis willbecome critical.

The neutrophils: A double-edged swordNeutrophils are part of the innate immune response to patho-

gens and are also rapidly activated in response to invading cancercells (28). Bone marrow–derived neutrophils are mobilized tosites of inflammation or cancer via their cell-surface receptorCXCR2 and in response to chemokines and cytokines (Table 1)secreted by activatedmacrophages, endothelial cells, or the tumorcells (reviewed in ref. 28). At the tumor site, the neutrophils canexert opposing effects that depend on the inflammatory/immunecontext (reviewed in ref. 64). They can release cytolytic factors(detailed in Table 1) or inhibit tumor growth indirectly viarecruitment of CD8þ cytotoxic T cells and macrophages. Thismay require direct contact with the tumor cells (64), such as canoccur within the liver sinusoidal lumen (65, 66).

On the other hand, neutrophils can be mobilized into pre-metastatic niches in the liver in response to S100A8 and S100A9(7) and were shown to contribute to niche formation in anorthotropic colon carcinoma model (67). Moreover, in the vas-cular lumen, the physical interaction with tumor cells thatcould lead to tumor cell kill may, alternatively, anchor circulatingtumor cells to the vascular endothelium and enhance their migra-tion into the extravascular space. Tumor-derived cytokines, suchas IL8, induce expression of integrins, such as CD11b/CD18, onthe neutrophils, increasing their adhesion to tumor cells viathe counter receptor ICAM-1 and facilitating transendothelialmigration, as shown in murine melanoma and lung carcinomamodels (66, 68). Tumor cell entrapment in the sinusoids canalso be mediated by neutrophil-produced extracellular (DNA)traps, resulting in increased tumor retention in the sinusoids,enhanced tumor cell adhesion, proliferation, migration and inva-sion, and increased livermetastasis (69, 70). Neutrophils can alsorelease extracellular matrix (ECM)–degrading proteinases, suchas MMP-8, MMP-9, elastase, and cathepsin G, thereby increasingtumor invasion (64), and can promote angiogenesis through therelease of VEGF (Table 1; Figs. 2 and 3).

Moreover, a TGFb-mediated neutrophil polarization thatalters the phenotype of tumor-associated neutrophils fromtumor inhibitory (N1) to tumor promoting (N2) was recentlydescribed (71). TGFb produced by M2 macrophages may con-tribute to this process (reviewed in ref. 59). However, the contri-bution of polarized neutrophils to liver metastasis remains tobe confirmed.

Clinical studies implicate neutrophils in tumor promotion.With few exceptions, elevated circulating neutrophil counts orneutrophil-to-lymphocyte ratios were associated with pooreroutcomes and distant metastases in patients with variousepithelial malignancies (72), including carcinomas of the gas-trointestinal tract. High neutrophil counts may also be associ-ated with increased resistance to therapy (64, 72). The 5-yearsurvival for patients with colorectal cancer undergoing hepaticresections that had neutrophil-to-lymphocyte ratios greaterthan 5 was worse than for those with normal ratios (73).However, high neutrophil infiltration into the tumor site maybe a consequence of advanced tumor stage (and, therefore,poorer prognosis) rather than its cause.

HSCs orchestrate a prometastatic microenvironment that isrequired for transition from the avascular to the vascularstage of metastasis

HSCs orchestrate the characteristic fibrogenic response of theliver to injury. Normally quiescent in the space of Disse (74),they are activated (aHSC) in response to liver damage andinflammatory stimuli, acquire a myofibroblast-like phenotype(a-SMAþ), and produce ECM rich in collagen I and IV (74, 75).Chemokines and cytokines released by aHSCs (Table 1) alsorecruit inflammatory/immune cells (76, 77), thereby shapingthe immune microenvironment.

In addition to their role in premetastatic niche formation(above), HSCs can orchestrate a prometastatic niche followingtumor extravasation. In response to growth factors and inflam-matory mediators (detailed in Table 1), HSCs are activated (74)and can trigger a process akin to the early events in liver repair.This was observed in animal models (4) and is also evidencedby the increased production of collagen IV in and aroundhepatic metastases in clinical specimens (78). Macrophages,hepatocytes, and LSECs contribute to this process by releasingTGFb and/or TNFa (57). In turn, aHSC-derived proangiogenicfactors, such as VEGF and angiopoietin-1 (79, 80), initiateangiogenesis (3), and this is enhanced by HSC-derived MMPsthat facilitate endothelial cell migration and tumor invasion(23, 74; reviewed in refs. 5, 81; Table 1; Fig. 2).

Several lines of evidence confirm the essential role of HSCs inliver metastasis. Olaso and colleagues showed that myofibro-blasts that infiltrated avascular micrometastases of B16 mela-noma cells induced a stromal response favorable to angio-genesis, which preceded endothelial cell recruitment and wasfollowed by colocalization of myofibroblasts and endothelialcells within angiogenic structures (82, 83). HSC and macro-phage recruitment into metastatic sites and subsequent angio-genesis were shown to be CCR2/CCL2 dependent (84). Morerecently, Eveno and colleagues (85), using immunofluores-cence, showed that 9 days postinjection of colorectal cancerLS174 cells into SCID mice, the hepatic micrometastases con-sisted of proliferating cancer cells, a well-organized network ofaHSC and laminin deposits, but no vascular network. As theliver metastases grew, an organized vascular network appearedand laminin colocalized with CD31þ endothelial cells. Co-injection of tumor cells and aHSCs enhanced metastasis. More-over, analysis of liver metastasis from patients with colorectalcancer revealed a surface-marker expression pattern similar tothat observed in the coinjection studies.

Recently, suppression of TGFb receptor II signaling byQGAP1 was shown to prevent HSC activation, and IQGAP1downregulation was observed in myofibroblasts associatedwith human colorectal cancer liver metastases, consistent withtheir activation in this context (86). Furthermore, a-SMAþ cells,whose presence correlated with the degree of fibrous stroma,were identified in liver metastases of colon, gastric, and pan-creatic adenocarcinomas (87).

The importance of HSCs to metastatic niche formation andtheir role in primary liver cancer (88) have raised interest inHSC targeting as an anticancer/antimetastatic strategy. There islittle direct evidence that such an approach can succeed, duepartially to the present lack of agents that specifically targetHSCs without toxicity. Indirect evidence from a rat model ofcholangiocarcinoma (89) and pancreatic stellate cell targetingin a PDAC model (81) suggests a potential therapeutic benefit,

The Microenvironment in Liver Metastasis

www.aacrjournals.org Clin Cancer Res; 22(24) December 15, 2016 5977

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 8: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

but this remains to be verified. Of note, portal fibroblasts andbone marrow–derived fibrocytes (90) can also contribute to thestromal response. Liver-invading cancer cells located in portaltracts and unable to activate HSCs may engage portal tractfibroblasts that produce IL8, a chemokine involved in invasionand angiogenesis (91). Specific targeting of HSCs may, there-fore, not be sufficient to deprive metastatic cells of a growth-promoting stroma.

Hepatocytes can promote metastasis directly and indirectlyThe role of the parenchymal hepatocytes in liver metastasis is

not well understood. Tumor cell adhesion to hepatocytes wasidentified as one of the earliest events in liver metastasis for-mation and a correlate of the metastatic potential (92, 93).Desmosomes (94), integrins av, a6, and b1 that bind tohepatocyte ECM (93), osteopontin binding via counter recep-tors CD44 and integrin av (95), and claudins (96, 97) were allimplicated. Adhesion of human colorectal cancer cells to hepa-tocyte-derived ECM was shown to upregulate the expression ofgenes involved in tumor cell survival, motility, and prolifera-tion, in particular EGF family members implicated in livermetastasis (98) such as AREG, EGFR, HBEGF, and erbB2 andstem cell markers CD133 (PROM1) and LGR5 (99), suggestingthat colorectal cancer adhesion to hepatocyte ECM inducesautocrine growth-promoting mechanisms for tumor expansion.Tumor cell adhesion to hepatocytes can have other conse-quences. The FasL/FLICE-like inhibitory protein on murineMC38 cells was shown to induce apoptosis in hepatocytes byactivating Fas signaling, and this destructive process created aniche for tumor expansion (100). Furthermore, hepatocytesproduce several growth factors, including IGF-I (101). As wehave shown, blockade of IGF-I signaling in metastatic tumorcells could inhibit liver metastasis (102–104). Other factorsinclude the HGF-like protein/macrophage-stimulating protein(HGFL) that can enhance tumor growth, motility, and inva-sion via the Ron receptor (105); heregulin, a ligand of ErbB3shown to enhance integrin avb5-mediated cell migration anderbB3/erbB2 signaling in human colorectal cancer cells (106);and the proangiogenic factor bFGF (107).

The clinical relevance of tumor–hepatocyte interactions isnot clear. Three different histologic growth patterns have beenidentified in liver metastases specimens, namely the desmo-plastic, pushing, and replacement growth patterns. The lattertwo are characterized by close tumor–hepatocyte proximity(23). Whether or not direct adhesion between metastaticcancer cells and hepatocytes influences the eventual growthpattern of colorectal cancer liver metastases is not known.Blocking tumor–hepatocyte interactions could have beneficialantimetastatic effects, as recently demonstrated by Tabariesand colleagues (108).

An immunosuppressive microenvironment facilitatesmetastatic expansion

Tumor cells can activate specific T-cell–mediated immuneresponses that curtail tumor expansion through various mechan-isms (reviewed in ref. 109). However, tumor cells can evadeT-cell–mediated killing, among others, by giving rise to and/orrecruiting immunosuppressive cells, such as myeloid-derivedsuppressor cells (MDSC) and regulatory T cells (Treg) that alterthe immune landscape, inducing a state of immune tolerance thatis permissive to tumor expansion (reviewed in ref. 110).

The role of Tregs. Treg cells may be thymically derived (natural,nTregs) or induced from na€�ve CD4þ T-cell precursors (iTreg)in response to IL10 and TGFb. They exert an immunosup-pressive effect by different mechanisms that result in theinhibition of an effective, antitumorigenic T-cell response(reviewed in ref. 111). Although high Treg levels were docu-mented in the blood and local tumor sites of patients withvarious epithelial cancers (e.g., ref. 112), relatively little isknown about their role in liver metastasis. Connolly andcolleagues, using models of early, preinvasive pancreaticneoplasia and advanced colorectal cancer showed that Tregsaccelerated the development of liver metastases (113). Werecently documented the accumulation of Tregs around coloncarcinoma MC38 liver micrometastases and showed that thiswas TNFR2 dependent (35).

Intriguingly, in 57 colon cancer specimens frompatients under-going chemotherapy or chemoimmunotherapy, higher Treg infil-tration scores were associated with a better prognosis and a betteroutcome (114). However, the relationship between Treg accu-mulation in the primary tumors and in the corresponding livermetastases has not been examined. Although Treg-targetingstrategies are in development (115, 116; reviewed in ref. 110),their potential therapeutic benefit for liver metastases has notyet been examined.

The role of MDSCs. MDSCs are a heterogeneous population ofmyeloid cells that can be of the monocytic (Mo-MDSC,CD11bþLy6Cþ) or granulocytic (G-MDSC, CD11bþLy6Gþ)lineages (117). In humans, MDSCs are CD33þ and/or CD11bþ

andHLA-DR� (118). Under normal physiologic conditions, bonemarrow–derived immature myeloid cells differentiate intomature granulocytes or monocytes, able to mediate host innateimmune responses in the target tissue. However, in the tumormicroenvironment, these precursors do not mature and exertimmunosuppressive and tumor-promoting effects instead(119). Although they express granulocyte and monocyte surfacemarkers, these cells have been characterized on the basis offunctional criteria, namely their immunosuppressive capabilities(117, 120). An increase in circulating MDSCs in cancer patientsand their recruitment into tumor sites have been documented(117, 121, 122).

In the liver, MDSCs can be recruited to the metastasesthrough chemokines, such as CXCL1 and CXCL2, producedby LSECs and KCs or by activated HSCs (77). There, they canpromote tumor growth by inducing immunosuppression(123), producing arginase (117), and increasing Treg numbers(Table 1; Fig. 2; ref. 124). Clinical and experimental evidenceconfirms their role in metastasis (125). Recruitment of tumor-promoting myeloid cells into colorectal cancer liver metastaseshas recently been documented, and their depletion was shownto result in a marked reduction in liver metastasis (126). Werecently reported that MDSCs accumulate in hepatic microme-tastases of MC38 cells within days of tumor cell injectionand documented a relative preponderance of Ly6Chigh cells(the more suppressive, arginase-high subtype; ref. 127). Wealso identified CD33þHLA-DR�TNFR2þ myeloid cells in theperiphery of hepatic metastases from patients with colorectalcancer, implicating MDSCs in the clinical disease (35).

Several strategies are being developed to selectively elim-inate MDSCs, including pharmacologic induction of MDSC dif-ferentiation, inhibition of MDSC expansion from bone marrow

Brodt

Clin Cancer Res; 22(24) December 15, 2016 Clinical Cancer Research5978

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 9: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

precursors, or blockade of their function (128). However, some ofthe drugs developed are not MDSC specific, or their long-termadministration is not possible due to toxicity (128). Their trans-lational potential remains therefore to be verified.

ConclusionsCancer cells entering the liver encounter a unique and complex

microenvironment. Parenchymal and nonparenchymal livercells, as well as recruited inflammatory and immune cells, par-ticipate in the response to invading tumor cells andmay inhibit orfavor the progression ofmetastasis. The factors that determine theoutcomeof these opposing effects and the pattern of growth of themetastases are not yet well understood. This duality of functionand the shifts in the types of cells and mediators involved atdifferent stages of the metastatic process render the attempts totarget specific cellular interactions in the microenvironmentextremely challenging and may explain the slow progress to datein identifying agents that can successfully and specificallyinhibit metastatic disease. With the advent of genomic profilingfor personalized cancer treatment and the increased availabilityof surgically resected liver metastases for cellular/molecular

interrogation, our understanding of the relationship betweenthe genetic background and clinical history of the patient, thegenomic/transcriptomic profile of the cancer cells, and the typeof response mounted by the microenvironment may improve,opening new avenues for prevention and/or treatment of livermetastatic disease.

Disclosure of Potential Conflicts of InterestNo potential conflicts of interest were disclosed.

AcknowledgmentsThe author is indebted to Dr. Janusz Rak (McGill University Health Centre)

for insightful editorial comments and to Simon Milette for the superb artisticrendering of the metastatic niches of the liver and for his help with the table.

Grant SupportThis work was made possible by support from the Canadian Institute for

Health Research and by a PSR-SIIRI-843grant from the Qu�ebec Minist�ere de

l'�Economie, de l'Innovation et des Exportations.

Received February 21, 2016; revised September 19, 2016; accepted September22, 2016; published OnlineFirst October 19, 2016.

References1. Hess KR, Varadhachary GR, Taylor SH, Wei W, Raber MN, Lenzi R,

et al. Metastatic patterns in adenocarcinoma. Cancer 2006;106:1624–33.

2. Kmiec Z. Cooperation of liver cells in health and disease. Adv AnatEmbryol Cell Biol 2001;161:III–XIII.

3. Smedsrod B, Le Couteur D, Ikejima K, Jaeschke H, Kawada N, Naito M,et al. Hepatic sinusoidal cells in health and disease: update from the 14thInternational Symposium. Liver Int 2009;29:490–501.

4. Vidal-Vanaclocha F. The prometastatic microenvironment of the liver.Cancer Microenviron 2008;1:113–29.

5. Vidal-Vanaclocha F. The tumor microenvironment at different stages ofhepatic metastasis. In: Brodt P, editor. Liver metastasis: biology andclinical management. Dordrecht, the Netherlands: Springer Scien-ceþBusiness Media B.V.; 2011. p. 1.

6. Kaplan RN, Riba RD, Zacharoulis S, Bramley AH, Vincent L, Costa C, et al.VEGFR1-positive haematopoietic bone marrow progenitors initiate thepre-metastatic niche. Nature 2005;438:820–7.

7. Hiratsuka S, Watanabe A, Aburatani H, Maru Y. Tumour-medi-ated upregulation of chemoattractants and recruitment of mye-loid cells predetermines lung metastasis. Nat Cell Biol 2006;8:1369–75.

8. Hiratsuka S, Watanabe A, Sakurai Y, Akashi-Takamura S, Ishibashi S,Miyake K, et al. The S100A8-serum amyloid A3-TLR4 paracrinecascade establishes a pre-metastatic phase. Nat Cell Biol 2008;10:1349–55.

9. Psaila B, Lyden D. Themetastatic niche: adapting the foreign soil. Nat RevCancer 2009;9:285–93.

10. Sceneay J, SmythMJ,Moller A. The pre-metastatic niche: finding commonground. Cancer Metastasis Rev 2013;32:449–64.

11. Costa-Silva B, Aiello NM, Ocean AJ, Singh S, Zhang H, Thakur BK, et al.Pancreatic cancer exosomes initiate pre-metastatic niche formation inthe liver. Nat Cell Biol 2015;17:816–26.

12. Hoshino A, Costa-Silva B, Shen TL, Rodrigues G, Hashimoto A, TesicMark M, et al. Tumour exosome integrins determine organotropicmetastasis. Nature 2015;527:329–35.

13. Kowanetz M, Wu X, Lee J, Tan M, Hagenbeek T, Qu X, et al. Granu-locyte-colony stimulating factor promotes lung metastasis throughmobilization of Ly6GþLy6Cþ granulocytes. Proc Natl Acad SciU S A 2010;107:21248–55.

14. Seubert B, Grunwald B, Kobuch J, Cui H, Schelter F, Schaten S, et al. Tissueinhibitor of metalloproteinases (TIMP)-1 creates a premetastatic niche in

the liver through SDF-1/CXCR4-dependent neutrophil recruitment inmice. Hepatology 2015;61:238–48.

15. Zhang Y, Davis C, Ryan J, Janney C, Pe~na MMO. Development andcharacterization of a reliablemousemodel of colorectal cancermetastasisto the liver. Clin Exp Metastasis 2013;30:903–18.

16. Bork U, Rahbari NN, Scholch S, Reissfelder C, Kahlert C, BuchlerMW, et al. Circulating tumour cells and outcome in non-met-astatic colorectal cancer: a prospective study. Br J Cancer 2015;112:1306–13.

17. Tsai WS, Chen JS, Shao HJ, Wu JC, Lai JM, Lu SH, et al. Circulating tumorcell count correlates with colorectal neoplasm progression and is aprognostic marker for distant metastasis in non-metastatic patients.Sci Rep 2016;6:24517.

18. Braet F, Nagatsuma K, Saito M, Soon L, Wisse E, Matsuura T. The hepaticsinusoidal endothelial lining and colorectal liver metastases. WorldJ Gastroenterol 2007;13:821–5.

19. Wang HH, McIntosh AR, Hasinoff BB, Rector ES, Ahmed N, Nance DM,et al. B16 melanoma cell arrest in the mouse liver induces nitric oxiderelease and sinusoidal cytotoxicity: a natural hepatic defense againstmetastasis. Cancer Res 2000;60:5862–9.

20. Yanagida H, Kaibori M, Yoshida H, Habara K, Yamada M, Kamiyama Y,et al. Hepatic ischemia/reperfusion upregulates the susceptibility ofhepatocytes to confer the induction of inducible nitric oxide synthasegene expression. Shock 2006;26:162–8.

21. Hehlgans T, Pfeffer K. The intriguing biology of the tumour necrosisfactor/tumour necrosis factor receptor superfamily: players, rules and thegames. Immunology 2005;115:1–20.

22. Ramadori G, Moriconi F, Malik I, Dudas J. Physiology and pathophys-iology of liver inflammation, damage and repair. J Physiol Pharmacol2008;59 Suppl 1:107–17.

23. Van den Eynden GG, Majeed AW, Illemann M, Vermeulen PB, Bird NC,Hoyer-Hansen G, et al. Themultifaceted role of themicroenvironment inliver metastasis: biology and clinical implications. Cancer Res 2013;73:2031–43.

24. Bertin S, Neves S, Gavelli A, Baque P, Brossette N, Simoes S, et al.Cellular and molecular events associated with the antitumor responseinduced by the cytosine deaminase/5-fluorocytosine suicide genetherapy system in a rat liver metastasis model. Cancer Gene Ther2007;14:858–66.

25. Takehara T, Uemura A, Tatsumi T, Suzuki T, Kimura R, Shiotani A, et al.Natural killer cell-mediated ablation of metastatic liver tumors by

The Microenvironment in Liver Metastasis

www.aacrjournals.org Clin Cancer Res; 22(24) December 15, 2016 5979

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 10: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

hydrodynamic injection of IFNalpha gene to mice. Int J Cancer 2007;120:1252–60.

26. Tatsumi T, Takehara T, Yamaguchi S, Sasakawa A, Miyagi T, Jinushi M,et al. Injection of IL-12 gene-transduced dendritic cells into mouse livertumor lesions activates both innate and acquired immunity. Gene Ther2007;14:863–71.

27. Pfister K, Radons J, Busch R, Tidball JG, Pfeifer M, Freitag L, et al. Patientsurvival by Hsp70 membrane phenotype: association with differentroutes of metastasis. Cancer 2007;110:926–35.

28. Spicer J, Brodt P, Ferri LE. Role of inflammation in the early stages of livermetastasis. In: Brodt P, editor. Liver metastasis: biology and clinicalmanagement. New York: Springer; 2011. p. 155–85.

29. Paschos KA, Majeed AW, Bird NC. Natural history of hepatic metastasesfrom colorectal cancer - pathobiological pathways with clinical signifi-cance. World J Gastroenterol 2014;20:3719–37.

30. Glinskii OV, Huxley VH, Glinsky GV, Pienta KJ, Raz A, Glinsky VV.Mechanical entrapment is insufficient and intercellular adhesion is essen-tial for metastatic cell arrest in distant organs. Neoplasia 2005;7:522–7.

31. Brodt P. Role of the host inflammatory response in colon carcinomainitiation, progression and liver metastasis. In: Beauchemin N, Huot J,editors. Metastasis of colorectal cancer. New York, NY: Springer; 2010.p. 289–319.

32. Jiao S-F, Sun K, Chen X-J, Zhao X, Cai N, Liu Y-J, et al. Inhibition of tumornecrosis factor alpha reduces the outgrowth of hepatic micrometastasisof colorectal tumors in a mouse model of liver ischemia-reperfusioninjury. J Biomed Sci 2014;21:1.

33. Khatib AM, Fallavollita L,Wancewicz EV,Monia BP, Brodt P. Inhibition ofhepatic endothelial E-selectin expression by C-raf antisense oligonucleo-tides blocks colorectal carcinoma liver metastasis. Cancer Res 2002;62:5393–8.

34. Yoshimoto K, Tajima H, Ohta T, Okamoto K, Sakai S, Kinoshita J, et al.Increased E-selectin in hepatic ischemia-reperfusion injury mediates livermetastasis of pancreatic cancer. Oncol Rep 2012;28:791–6.

35. Ham B, Wang N, D'Costa Z, Fernandez MC, Bourdeau F, Auguste P, et al.TNF receptor-2 facilitates an immunosuppressive microenvironment inthe liver to promote the colonization and growth of hepatic metastases.Cancer Res 2015;75:5235–47.

36. Elliott VA, Rychahou P, Zaytseva YY, Evers BM. Activation of c-Met andupregulation of CD44 expression are associated with the metastaticphenotype in the colorectal cancer liver metastasis model. PLoS One2014;9:e97432.

37. Witz IP. The selectin-selectin ligand axis in tumor progression. CancerMetastasis Rev 2008;27:19–30.

38. Greenwald E, Yuki K. A translational consideration of intercellular adhe-sionmolecule-1 biology in the perioperative setting. Transl Perioper PainMed 2016;1:17–23.

39. Tremblay PL, Huot J, Auger FA.Mechanisms by which E-selectin regulatesdiapedesis of colon cancer cells under flow conditions. Cancer Res2008;68:5167–76.

40. Aychek T, Miller K, Sagi-Assif O, Levy-Nissenbaum O, Israeli-Amit M,Pasmanik-ChorM, et al. E-selectin regulates gene expression inmetastaticcolorectal carcinoma cells and enhances HMGB1 release. Int J Cancer2008;123:1741–50.

41. Ou J, Peng Y, Deng J, MiaoH, Zhou J, Zha L, et al. Endothelial cell-derivedfibronectin extra domain A promotes colorectal cancer metastasis viainducing epithelial-mesenchymal transition. Carcinogenesis 2014;35:1661–70.

42. Hu CT, Guo LL, Feng N, Zhang L, Zhou N, Ma LL, et al. MIF, secreted byhuman hepatic sinusoidal endothelial cells, promotes chemotaxis andoutgrowth of colorectal cancer in liver prometastasis. Oncotarget2015;6:22410–23.

43. Banerjee D, Hernandez SL, Garcia A, Kangsamaksin T, Sbiroli E,Andrews J, et al. Notch suppresses angiogenesis and progression ofhepatic metastases. Cancer Res 2015;75:1592–602.

44. Im JH, Tapmeier T, Balathasan L, Gal A, Yameen S, Hill S, et al. G-CSFrescues tumor growth and neo-angiogenesis during livermetastasis underhost angiopoietin-2 deficiency. Int J Cancer 2013;132:315–26.

45. Yona S, KimKW,Wolf Y,Mildner A, VarolD, BrekerM, et al. Fatemappingreveals origins and dynamics ofmonocytes and tissuemacrophages underhomeostasis. Immunity 2013;38:79–91.

46. Bouwens L, Baekeland M, De Zanger R, Wisse E. Quantitation, tissuedistribution and proliferation kinetics of Kupffer cells in normal rat liver.Hepatology 1986;6:718–22.

47. van der Bij GJ, Oosterling SJ, Meijer S, Beelen RH, van Egmond M.Therapeutic potential of Kupffer cells in prevention of liver metastasesoutgrowth. Immunobiology 2005;210:259–65.

48. TimmersM, VekemansK, VermijlenD, AsosinghK, Kuppen P, Bouwens L,et al. Interactions between rat colon carcinoma cells and Kupffer cellsduring the onset of hepatic metastasis. Int J Cancer 2004;112:793–802.

49. Bayon LG, Izquierdo MA, Sirovich I, van Rooijen N, Beelen RH, Meijer S.Role of Kupffer cells in arresting circulating tumor cells and controllingmetastatic growth in the liver. Hepatology 1996;23:1224–31.

50. Matsumura H, Kondo T, Ogawa K, Tamura T, Fukunaga K,Murata S, et al.Kupffer cells decrease metastasis of colon cancer cells to the liver in theearly stage. Int J Oncol 2014;45:2303–10.

51. KhatibAM,Auguste P, Fallavollita L,WangN, Samani A,KontogianneaM,et al. Characterization of the host proinflammatory response to tumorcells during the initial stages of liver metastasis. Am J Pathol 2005;167:749–59.

52. Khatib AM, Kontogiannea M, Fallavollita L, Jamison B, Meterissian S,Brodt P. Rapid induction of cytokine and E-selectin expression in the liverin response to metastatic tumor cells. Cancer Res 1999;59:1356–61.

53. Wen SW, Ager EI, Christophi C. Bimodal role of Kupffer cells duringcolorectal cancer liver metastasis. Cancer Biol Ther 2013;14:606–13.

54. Jessup JM, Samara R, Battle P, Laguinge LM. Carcinoembryonic antigenpromotes tumor cell survival in liver through an IL-10-dependent path-way. Clin Exp Metastasis 2004;21:709–17.

55. Asao T, Shibata HR, Batist G, Brodt P. Eradication of hepaticmetastases ofcarcinoma H-59 by combination chemoimmunotherapy with liposomalmuramyl tripeptide, 5-fluorouracil, and leucovorin. Cancer Res 1992;52:6254–7.

56. Tosello-Trampont A-C, Landes SG, Nguyen V, Novobrantseva TI, HahnYS. Kuppfer cells trigger nonalcoholic steatohepatitis development indiet-induced mouse model through tumor necrosis factor-a production.J Biol Chem 2012;287:40161–72.

57. Dey A, Allen J, Hankey-Giblin PA. Ontogeny and polarization of macro-phages in inflammation: blood monocytes versus tissue macrophages.Front Immunol 2014;5:683.

58. Murray PJ, Allen JE, Biswas SK, Fisher EA, Gilroy DW, Goerdt S, et al.Macrophage activation and polarization: nomenclature and experimentalguidelines. Immunity 2014;41:14–20.

59. SchouppeE,DeBaetselier P, VanGinderachter JA, SarukhanA. Instructionof myeloid cells by the tumor microenvironment: open questions on thedynamics and plasticity of different tumor-associated myeloid cell popu-lations. Oncoimmunology 2012;1:1135–45.

60. Galdiero MR, Bonavita E, Barajon I, Garlanda C, Mantovani A, Jaillon S.Tumor associated macrophages and neutrophils in cancer. Immunobiol-ogy 2013;218:1402–10.

61. Mills CD. Anatomy of a discovery: m1 and m2 macrophages. FrontImmunol 2015;6:212.

62. Mills CD, Lenz LL, Harris RA. A breakthrough: macrophage-directedcancer immunotherapy. Cancer Res 2016;76:513–6.

63. Cui YL, Li HK, Zhou HY, Zhang T, Li Q. Correlations of tumor-associatedmacrophage subtypes with livermetastases of colorectal cancer. Asian PacJ Cancer Prev 2013;14:1003–7.

64. Sionov RV, Fridlender ZG, Granot Z. The multifaceted roles neutro-phils play in the tumor microenvironment. Cancer Microenviron2015;8:125–58.

65. McDonald B, Spicer J,Giannais B, Fallavollita L, Brodt P, Ferri LE. Systemicinflammation increases cancer cell adhesion to hepatic sinusoids byneutrophil mediated mechanisms. Int J Cancer 2009;125:1298–305.

66. Spicer JD, McDonald B, Cools-Lartigue JJ, Chow SC, Giannias B, Kubes P,et al. Neutrophils promote liver metastasis via Mac-1-mediated interac-tions with circulating tumor cells. Cancer Res 2012;72:3919–27.

67. Yamamoto M, Kikuchi H, Ohta M, Kawabata T, Hiramatsu Y, Kondo K,et al. TSU68 prevents liver metastasis of colon cancer xenografts bymodulating the premetastatic niche. Cancer Res 2008;68:9754–62.

68. Huh SJ, Liang S, Sharma A, Dong C, Robertson GP. Transiently entrappedcirculating tumor cells interact with neutrophils to facilitate lung metas-tasis development. Cancer Res 2010;70:6071–82.

Clin Cancer Res; 22(24) December 15, 2016 Clinical Cancer Research5980

Brodt

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 11: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

69. Cools-Lartigue J, Spicer J, McDonald B, Gowing S, Chow S, Giannias B,et al. Neutrophil extracellular traps sequester circulating tumor cells andpromote metastasis. J Clin Invest 2013;123:3446–58.

70. Tohme S, Yazdani HO, Al-Khafaji AB, Chidi AP, Loughran P, Mowen K,et al. Neutrophil extracellular traps promote the development and pro-gression of liver metastases after surgical stress. Cancer Res 2016;76:1367–80.

71. Fridlender ZG, Sun J, Kim S, Kapoor V, Cheng G, Ling L, et al. Polarizationof tumor-associated neutrophil phenotype by TGF-beta: "N1" versus"N2" TAN. Cancer Cell 2009;16:183–94.

72. Donskov F. Immunomonitoring and prognostic relevance of neutrophilsin clinical trials. Semin Cancer Biol 2013;23:200–7.

73. Halazun KJ, Aldoori A, Malik HZ, Al-Mukhtar A, Prasad KR, Toogood GJ,et al. Elevated preoperative neutrophil to lymphocyte ratio predictssurvival following hepatic resection for colorectal liver metastases. EurJ Surg Oncol 2008;34:55–60.

74. Friedman SL. Hepatic stellate cells: protean, multifunctional, and enig-matic cells of the liver. Physiol Rev 2008;88:125–72.

75. Gressner AM, Bachem MG. Molecular mechanisms of liver fibrogen-esis–a homage to the role of activated fat-storing cells. Digestion1995;56:335–46.

76. MuhannaN, Doron S,WaldO,Horani A, Eid A, PappoO, et al. Activationof hepatic stellate cells after phagocytosis of lymphocytes: a novel path-way of fibrogenesis. Hepatology 2008;48:963–77.

77. Zhao W, Zhang L, Xu Y, Zhang Z, Ren G, Tang K, et al. Hepatic stellatecells promote tumor progression by enhancement of immunosuppres-sive cells in an orthotopic liver tumor mouse model. Lab Invest2014;94:182–91.

78. Burnier JV, Wang N, Michel RP, Hassanain M, Li S, Lu Y, et al. Type IVcollagen-initiated signals provide survival and growth cues required forliver metastasis. Oncogene 2011;30:3766–83.

79. Copple BL, Bai S, Burgoon LD, Moon J-OK. Hypoxia-inducible factor-1a regulates expression of genes in hypoxic hepatic stellate cellsimportant for collagen deposition and angiogenesis. Liver Int 2011;31:230–44.

80. Taura K,DeMinicis S, Seki E, Hatano E, Iwaisako K,Osterreicher CH, et al.Hepatic stellate cells secrete angiopoietin 1 that induces angiogenesis inliver fibrosis. Gastroenterology 2008;135:1729–38.

81. KangN, Shah VH,Urrutia R.Membrane-to-nucleus signals and epigeneticmechanisms for myofibroblastic activation and desmoplastic stroma:potential therapeutic targets for liver metastasis? Mol Cancer Res2015;13:604–12.

82. Olaso E, Salado C, Egilegor E, Gutierrez V, Santisteban A, Sancho-Bru P,et al. Proangiogenic role of tumor-activated hepatic stellate cells inexperimental melanoma metastasis. Hepatology 2003;37:674–85.

83. Badiola I,Olaso E, CrendeO, Friedman SL, Vidal-Vanaclocha F.Discoidindomain receptor 2 deficiency predisposes hepatic tissue to colon carci-noma metastasis. Gut 2012;61:1465–72.

84. Yang X, Lu P, Ishida Y, Kuziel WA, Fujii C, Mukaida N. Attenuated livertumor formation in the absence of CCR2with a concomitant reduction inthe accumulation of hepatic stellate cells, macrophages and neovascular-ization. Int J Cancer 2006;118:335–45.

85. Eveno C, Hainaud P, Rampanou A, Bonnin P, Bakhouche S, Dupuy E,et al. Proof of prometastatic niche inductionby hepatic stellate cells. J SurgRes 2015;194:496–504.

86. Liu C, BilladeauDD, AbdelhakimH, Leof E, Kaibuchi K, BernabeuC, et al.IQGAP1 suppresses TbetaRII-mediated myofibroblastic activation andmetastatic growth in liver. J Clin Invest 2013;123:1138–56.

87. Terada T, Makimoto K, Terayama N, Suzuki Y, Nakanuma Y. Alpha-smooth muscle actin-positive stromal cells in cholangiocarcinomas,hepatocellular carcinomas and metastatic liver carcinomas. J Hepatol1996;24:706–12.

88. Ju MJ, Qiu SJ, Fan J, Xiao YS, Gao Q, Zhou J, et al. Peritumoral activatedhepatic stellate cells predict poor clinical outcome in hepatocellularcarcinoma after curative resection. Am J Clin Pathol 2009;131:498–510.

89. Sirica AE, Gores GJ. Desmoplastic stroma and cholangiocarcinoma:clinical implications and therapeutic targeting. Hepatology 2014;59:2397–402.

90. Brenner DA, Kisseleva T, Scholten D, Paik YH, Iwaisako K, Inokuchi S,et al. Origin of myofibroblasts in liver fibrosis. Fibrogenesis Tissue Repair2012;5(Suppl 1):S17.

91. Mueller L, Goumas FA, AffeldtM, Sandtner S, GehlingUM, Brilloff S, et al.Stromal fibroblasts in colorectal liver metastases originate from residentfibroblasts and generate an inflammatorymicroenvironment. AmJPathol2007;171:1608–18.

92. Wang J, Fallavollita L, Brodt P. Inhibition of experimental hepatic metas-tasis by amonoclonal antibody that blocks tumor-hepatocyte interaction.J Immunother Emphasis Tumor Immunol 1994;16:294–302.

93. Mook ORF, van Marie J, Jonges R, Vreeling-Sindelarova H, Frederiks WM,Van Noorden CJF. Interactions between colon cancer cells and hepato-cytes in rats in relation to metastasis. J Cell Mol Med 2008;12:2052–61.

94. Shimizu S, Yamada N, Sawada T, Ikeda K, Nakatani K, Seki S, et al.Ultrastructure of early phase hepatic metastasis of human colon carci-noma cells with special reference to desmosomal junctions with hepa-tocytes. Pathol Int 2000;50:953–9.

95. Huang J, Pan C, Hu H, Zheng S, Ding L. Osteopontin-enhanced hepaticmetastasis of colorectal cancer cells. PLoS One 2012;7:e47901.

96. Tabaries S, Dupuy F,Dong Z,Monast A, AnnisMG, Spicer J, et al. Claudin-2 promotes breast cancer liver metastasis by facilitating tumor cell inter-actions with hepatocytes. Mol Cell Biol 2012;32:2979–91.

97. Georges R, Bergmann F, Hamdi H, Zepp M, Eyol E, Hielscher T, et al.Sequential biphasic changes in claudin1 and claudin4 expression arecorrelated to colorectal cancer progression and liver metastasis. J Cell MolMed 2012;16:260–72.

98. Radinsky R, Risin S, Fan D, Dong Z, Bielenberg D, Bucana CD, et al. Leveland function of epidermal growth factor receptor predict the metastaticpotential of human colon carcinoma cells. ClinCancer Res 1995;1:19–31.

99. Zvibel I, Wagner A, Pasmanik-Chor M, Varol C, Oron-Karni V, Santo EM,et al. Transcriptional profiling identifies genes induced by hepatocyte-derived extracellular matrix in metastatic human colorectal cancer celllines. Clin Exp Metastasis 2013;30:189–200.

100. Li H, Fan X, Stoicov C, Liu JH, Zubair S, Tsai E, et al. Human and mousecolon cancer utilizes CD95 signaling for local growth and metastaticspread to liver. Gastroenterology 2009;137:934–44.

101. Long L, Nip J, Brodt P. Paracrine growth stimulation by hepatocyte-derived insulin-like growth factor-1: a regulatory mechanism for carci-noma cells metastatic to the liver. Cancer Res 1994;54:3732–7.

102. Long L, Rubin R, Baserga R, Brodt P. Loss of the metastatic phenotype inmurine carcinoma cells expressing an antisense RNA to the insulin-likegrowth factor receptor. Cancer Res 1995;55:1006–9.

103. Samani AA, Chevet E, Fallavollita L, Galipeau J, Brodt P. Loss of tumor-igenicity and metastatic potential in carcinoma cells expressing theextracellular domain of the type 1 insulin-like growth factor receptor.Cancer Res 2004;64:3380–5.

104. Wang N, Rayes RF, Elahi SM, Lu Y, Hancock MA, Massie B, et al. The IGF-Trap: novel inhibitor of carcinoma growth and metastasis. Mol CancerTher 2015;14:982–93.

105. Wagh P, Peace BE,Waltz SE. Themet-related receptor tyrosine kinase Ronin tumor growth and metastasis. Adv Cancer Res 2008;100:1–33.

106. Yoshioka T, Nishikawa Y, Ito R, Kawamata M, Doi Y, Yamamoto Y, et al.Significance of integrin alphavbeta5 and erbB3 in enhanced cell migra-tion and liver metastasis of colon carcinomas stimulated by hepatocyte-derived heregulin. Cancer Sci 2010;101:2011–8.

107. Dome B, HendrixMJ, Paku S, Tovari J, Timar J. Alternative vascularizationmechanisms in cancer: Pathology and therapeutic implications. Am JPathol 2007;170:1–15.

108. Tabaries S, Annis MG, Hsu BE, Tam CE, Savage P, Park M, et al. Lynmodulates Claudin-2 expression and is a therapeutic target for breastcancer liver metastasis. Oncotarget 2015;6:9476–87.

109. Hadrup S, Donia M, Thor Straten P. Effector CD4 and CD8 T cells andtheir role in the tumor microenvironment. Cancer Microenviron2013;6:123–33.

110. Butt AQ, Mills KH. Immunosuppressive networks and checkpoints con-trolling antitumor immunity and their blockade in the development ofcancer immunotherapeutics and vaccines. Oncogene 2014;33:4623–31.

111. Facciabene A,MotzGT, CoukosG. T-regulatory cells: key players in tumorimmune escape and angiogenesis. Cancer Res 2012;72:2162–71.

112. Wolf AM, Wolf D, Steurer M, Gastl G, Gunsilius E, Grubeck-LoebensteinB. Increase of regulatory T cells in the peripheral blood of cancer patients.Clin Cancer Res 2003;9:606–12.

113. Connolly MK, Mallen-St Clair J, Bedrosian AS, Malhotra A, Vera V,Ibrahim J, et al. Distinct populations of metastases-enabling myeloid

www.aacrjournals.org Clin Cancer Res; 22(24) December 15, 2016 5981

The Microenvironment in Liver Metastasis

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 12: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

cells expand in the liver of mice harboring invasive and preinvasive intra-abdominal tumor. J Leuk Biol 2010;87:713–25.

114. Correale P, Rotundo MS, Del Vecchio MT, Remondo C, Migali C,Ginanneschi C, et al. Regulatory (FoxP3þ) T-cell tumor infiltration is afavorable prognostic factor in advanced colon cancer patients undergoingchemo or chemoimmunotherapy. J Immunother 2010;33:435–41.

115. Rech AJ, Vonderheide RH. Clinical use of anti-CD25 antibody daclizu-mab to enhance immune responses to tumor antigen vaccination bytargeting regulatory T cells. Ann N Y Acad Sci 2009;1174:99–106.

116. Rech AJ, Mick R, Martin S, Recio A, Aqui NA, Powell DJ Jr, et al. CD25blockade depletes and selectively reprograms regulatory T cells in concertwith immunotherapy in cancer patients. Sci Transl Med 2012;4:134ra62.

117. KeskinovAA, ShurinMR.Myeloid regulatory cells in tumor spreading andmetastasis. Immunobiology 2015;220:236–42.

118. Almand B, Clark JI, Nikitina E, van Beynen J, English NR, Knight SC,et al. Increased production of immature myeloid cells in cancerpatients: a mechanism of immunosuppression in cancer. J Immunol2001;166:678–89.

119. Gabrilovich DI, Ostrand-Rosenberg S, Bronte V. Coordinated regulationof myeloid cells by tumours. Nat Rev Immunol 2012;12:253–68.

120. Haverkamp JM, Smith AM, Weinlich R, Dillon CP, Qualls JE, Neale G,et al. Myeloid-derived suppressor activity is mediated by monocyticlineages maintained by continuous inhibition of extrinsic and intrinsicdeath pathways. Immunity 2014;41:947–59.

121. KhaledYS, Ammori BJ, Elkord E. Increased levels of granulocyticmyeloid-derived suppressor cells in peripheral blood and tumour tissue of pan-creatic cancer patients. J Immunol Res 2014;2014:879897.

122. Meyer C, Cagnon L, Costa-Nunes CM, Baumgaertner P, Montandon N,Leyvraz L, et al. Frequencies of circulating MDSC correlate with clinicaloutcome of melanoma patients treated with ipilimumab. Cancer Immu-nol Immunother 2014;63:247–57.

123. Kusmartsev S, Nefedova Y, Yoder D, Gabrilovich DI. Antigen-specificinhibition of CD8þ T cell response by immaturemyeloid cells in cancer ismediated by reactive oxygen species. J Immunol 2004;172:989–99.

124. Huang B, Pan PY, LiQ, Sato AI, LevyDE, Bromberg J, et al. Gr-1þCD115þimmature myeloid suppressor cells mediate the development of tumor-induced T regulatory cells and T-cell anergy in tumor-bearing host.Cancer Res 2006;66:1123–31.

125. Diaz-Montero CM, Salem ML, Nishimura MI, Garrett-Mayer E, ColeDJ, Montero AJ. Increased circulating myeloid-derived suppressor cellscorrelate with clinical cancer stage, metastatic tumor burden, anddoxorubicin-cyclophosphamide chemotherapy. Cancer ImmunolImmunother 2009;58:49–59.

126. Zhao L, Lim SY, Gordon-Weeks AN, Tapmeier TT, Im JH, Cao Y, et al.Recruitment of a myeloid cell subset (CD11b/Gr1 mid) via CCL2/CCR2 promotes the development of colorectal cancer liver metastasis.Hepatology 2013;57:829–39.

127. Rodriguez PC, Ernstoff MS, Hernandez C, Atkins M, Zabaleta J, Sierra R,et al. Arginase I-producing myeloid-derived suppressor cells in renal cellcarcinoma are a subpopulation of activated granulocytes. Cancer Res2009;69:1553–60.

128. Ugel S, Delpozzo F, Desantis G, Papalini F, Simonato F, Sonda N, et al.Therapeutic targeting of myeloid-derived suppressor cells. Curr OpinPharmacol 2009;9:470–81.

Clin Cancer Res; 22(24) December 15, 2016 Clinical Cancer Research5982

Brodt

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460

Page 13: Role of the Microenvironment in Liver Metastasis: From Pre ......Review Role of the Microenvironment in Liver Metastasis: From Pre- to Prometastatic Niches Pnina Brodt Abstract Liver

2016;22:5971-5982. Published OnlineFirst October 19, 2016.Clin Cancer Res   Pnina Brodt  Prometastatic NichesRole of the Microenvironment in Liver Metastasis: From Pre- to

  Updated version

  10.1158/1078-0432.CCR-16-0460doi:

Access the most recent version of this article at:

   

   

  Cited articles

  http://clincancerres.aacrjournals.org/content/22/24/5971.full#ref-list-1

This article cites 125 articles, 29 of which you can access for free at:

  Citing articles

  http://clincancerres.aacrjournals.org/content/22/24/5971.full#related-urls

This article has been cited by 9 HighWire-hosted articles. Access the articles at:

   

  E-mail alerts related to this article or journal.Sign up to receive free email-alerts

  Subscriptions

Reprints and

  [email protected]

To order reprints of this article or to subscribe to the journal, contact the AACR Publications Department at

  Permissions

  Rightslink site. Click on "Request Permissions" which will take you to the Copyright Clearance Center's (CCC)

.http://clincancerres.aacrjournals.org/content/22/24/5971To request permission to re-use all or part of this article, use this link

on August 18, 2021. © 2016 American Association for Cancer Research. clincancerres.aacrjournals.org Downloaded from

Published OnlineFirst October 19, 2016; DOI: 10.1158/1078-0432.CCR-16-0460