Next Article in Journal
EML4-ALK Variants: Biological and Molecular Properties, and the Implications for Patients
Next Article in Special Issue
Reduced Cytokine Release in Ex Vivo Response to Cilengitide and Cetuximab Is a Marker for Improved Survival of Head and Neck Cancer Patients
Previous Article in Journal
Chimeric Antigen Receptor (CAR) T Cell Therapy for Malignant Pleural Mesothelioma (MPM)
Previous Article in Special Issue
Integrins as Therapeutic Targets: Successes and Cancers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exploring the Role of RGD-Recognizing Integrins in Cancer

1
Department of Oral and Maxillofacial Surgery, University Hospital Rechts der Isar, Technische Universität München, Ismaninger Straße 22, 81679 Munich, Germany
2
Clinical Research Unit, Department of Obstetrics & Gynecology, Technische Universität München, Ismaninger Straße 22, 81675 Munich, Germany
3
Institute for Advanced Study and Center for Integrated Protein Science (CIPSM), Department Chemie, Technische Universität München, Lichtenbergstraße 4, 85747 Garching, Germany
4
Pharmaceutical Radiochemistry, Technische Universität München, Walther-Meißner-Straße 3, 85748 Garching, Germany
5
Department of Nuclear Medicine, Technische Universität München, Ismaninger Straße 22, 81679 Munich, Germany
6
Institute of Pathology, Technische Universität München, Trogerstraße 18, 81675 Munich, Germany
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Cancers 2017, 9(9), 116; https://doi.org/10.3390/cancers9090116
Submission received: 6 July 2017 / Revised: 28 August 2017 / Accepted: 31 August 2017 / Published: 4 September 2017
(This article belongs to the Special Issue Integrins in Cancer)

Abstract

:
Integrins are key regulators of communication between cells and with their microenvironment. Eight members of the integrin superfamily recognize the tripeptide motif Arg-Gly-Asp (RGD) within extracelluar matrix (ECM) proteins. These integrins constitute an important subfamily and play a major role in cancer progression and metastasis via their tumor biological functions. Such transmembrane adhesion and signaling receptors are thus recognized as promising and well accessible targets for novel diagnostic and therapeutic applications for directly attacking cancer cells and their fatal microenvironment. Recently, specific small peptidic and peptidomimetic ligands as well as antibodies binding to distinct integrin subtypes have been developed and synthesized as new drug candidates for cancer treatment. Understanding the distinct functions and interplay of integrin subtypes is a prerequisite for selective intervention in integrin-mediated diseases. Integrin subtype-specific ligands labelled with radioisotopes or fluorescent molecules allows the characterization of the integrin patterns in vivo and later the medical intervention via subtype specific drugs. The coating of nanoparticles, larger proteins, or encapsulating agents by integrin ligands are being explored to guide cytotoxic reagents directly to the cancer cell surface. These ligands are currently under investigation in clinical studies for their efficacy in interference with tumor cell adhesion, migration/invasion, proliferation, signaling, and survival, opening new treatment approaches in personalized medicine.

1. Introduction

Cellular recognition of the interstitial neighborhood and microenvironment is required within all organisms, which depends on cell-extracellular matrix (ECM) and cell-cell communications for the correct functioning of individual cells as essential features of life [1]. In this respect, the cell adhesion and signaling receptors of the integrin superfamily are thought to play a major role. Integrins are heterodimeric transmembrane glycoproteins, consisting of one α- and one β-subunit. They execute crucial regulatory functions during cell adhesion, migration/invasion, proliferation, survival, and apoptosis. Moreover, integrins are capable of bidirectional signaling across cell membranes, referred to as “outside-in” and “inside-out” signaling, which results in information exchange between the ECM proteins and intracellular molecules [2,3,4,5]. Since integrins are linked to cytoskeletal components within the cell and with ECM proteins within the extracellular space, they function as mechanotransducers exhibiting force-sensing capabilities. These functional features of integrins have gained increasing attention in recent years [6,7].
In 1984, Pierschbacher and Ruoslahti described the Arg-Gly-Asp (RGD) peptide motif as a highly conserved minimal integrin recognition sequence within fibronectin [8]. Subsequently, the RGD motif was identified in many other ECM proteins, including vitronectin [9], von Willebrand factor [10], osteopontin [11], and laminin [12]. Although many integrins recognize ECM proteins via the RGD motif, the specificity of integrins governing this interaction turned out to be more complex. Meanwhile, it became clear that ligand specificity for integrins—besides the RGD motif—also depends on the distinct conformational and spatial presentation of this motif in various ECM proteins as well as on other synergistically functioning adjacent molecular regions [13,14,15]. Based on these findings, synthetic peptides and peptidomimetics, displaying the RGD motif have been developed as potent integrin ligands and antagonists, inhibiting the adhesion of anchorage-dependent cells to ECM proteins and thereby controlling integrin-mediated (tumor) biological functions.
Among the 24 human integrin subtypes known to date, eight integrin dimers, i.e., αvβ1, αvβ3, αvβ5, αvβ6, αvβ8, α5β1, α8β1, and αIIbβ3, recognize the tripeptide RGD motif within ECM proteins. Those members constitute a most important integrin receptor subfamily instrumental in cancer and their metastasis [16]. The exploration of functional differences between distinct members of the integrin subfamily is challenging, because the integrin expression profiles strongly depend on the cell type and the microenvironment within a given temporal and spatial context. To add even more complexity, integrins act in concert with other receptor proteins that are functionally and/or spatially organized within multiprotein networks within so-called focal adhesions (FA) [17,18]. Thus, the elucidation of the specific properties of distinct integrin subtypes governing their in vivo activities requires highly active and selective ligands.
The rapidly growing field of integrins as targets for cancer diagnosis and therapy has stimulated us to provide a thorough update on the impact of RGD-recognizing integrins in cancer, their distinct tumor-cell-specific functions and their crosstalk with other membrane receptors in various cancer entities. For previous reviews of integrins in cancer please see [19,20,21,22,23,24,25,26,27,28].

2. Integrin Activation upon Conformational Rearrangements

Over the many years following their discovery, integrins have been characterized in great detail. Nevertheless, the exact mechanisms of their molecular activation and bidirectional signaling remained unclear for a long time, mostly because of the missing structural information regarding these non-covalently linked heterodimeric receptors. This changed in 2001, when the structure of the extracellular domain of αvβ3 in the absence or presence of Cilengitide, a small cyclic peptide ligand, was solved [29,30]. These structural data unraveled that the extracellular domains of αvβ3 exist in a resting state in a conformation in which the integrin head groups are bent towards the cell membrane. Ligands may even bind to the kinked integrin in the resting state in which the transmembrane domains of each subunit are closely associated [31,32]. Upon ligand binding, however, structural rearrangements in the hinge region of each integrin subunit occur via the so-called “switchblade” mechanism. Meanwhile, several structural studies have elucidated the details of the conformational alterations during integrin activation [2]. Key elements in the outside-in activation are a shift of the α7-helix in the βI-domain and a dissociation of the transmembrane helices of the α- and the β-subunit [31,33]. Homo-oligomerization of both subunits then activates the intracellular part by the binding of intracellular proteins, such as talin and kindlin, to the cytoplasmic domain of the β3-subunit, which is crucial for outside-in signaling [34,35] (Figure 1). In turn, the binding of talin results in changes within the integrin conformation to an activated and erected state, accompanied by the separation of the integrin transmembrane and cytoplasmic domains (inside-out). The headpieces either remain in the closed and bent state or in the open conformation, which then results in high ligand binding affinity.
The elucidation of the molecular mechanisms of integrin activation reveals that integrins can harbor “on” and “off” states upon conformational rearrangements by switching from a low to a high affinity state. The binding of ECM ligands with high affinity then triggers increased adhesiveness and signal transduction to the cytoplasm in the outside-in direction. To this end, phosphorylation events occur to activate downstream signaling kinases, such as the mitogen-activated protein kinases (MAPK), phosphoinositide kinase (PI3K)/Akt, and extracellular signal regulated kinase (ERK) [31,36,37,38,39], which in turn have an impact on cell proliferation, migration/invasion, and cell survival. Modulations of those integrin-mediated cell biological events are instrumental in malignant cellular transformation and tumor metastasis [40,41,42]. Figure 2 summarizes the various functions of specific integrins in the tumor biological contexts.

3. Integrins in Cancer Progression and Metastasis

In a series of various cancer entities, the expression of αvβ3, αvβ5, α5β1, α6β4, α4β1, αvβ6 and αvβ8 has been shown to correlate well with metastasis and poor patient prognosis. Here, αvβ3, αvβ5, and α5β1 are among the most prominent and well-studied integrin subtypes [21]. Table 1 provides an overview of RGD-binding integrins, their expression in various human cancer entities and their prognostic impact.
As such, the enhanced expression of αvβ3 favors tumor growth and invasion and metastasis to the bones, especially when ligated by its major ECM ligands vitronectin and/or osteopontin [43,44]. In addition to αv-integrins, α5β1 expression levels are strongly induced upon hypoxia and promote tumor metastasis in breast cancer. Thus, high α5 expression in clinical biopsies is associated with an increased risk of mortality [45].
Moreover, integrin αvβ6 has been proposed to serve as a marker for tumor cell invasiveness of gastric, pancreatic, cholangiocellular, breast, ovarian, colon, and head and neck cancer. Clinical data have demonstrated that αvβ6 expression levels correlated with poor patient outcome [46,47,48,49,50,51,52,53,54,55]. In accordance, high αvβ6 expression seems to be associated with a more aggressive disease and, consequently, poor outcome and prognosis for patients afflicted with colon cancer [56,57]. The clinical outcome is mainly attributable to the ability of colon cancer cells to metastasize to the liver. In gastric cancer, αvβ6 expression was present in 37% of cases and was predictive of the reduced survival of cancer patients. The analysis of the relationship between αvβ6 expression and the progression of cancer revealed that induced expression of matrix metalloproteases (MMP) and activation of the TGF-β1 pathway are reiteratively activated [53,58,59]. In oral squamous cell carcinoma (OSCC) cells, αvβ6 is highly expressed at the invasive tumor front. In gastroenteropancreatic adenocarcinomas, however, only weak αvβ6 expression was noticed at the cancer cell membranes [51]. In gastroenteropancreatic malignancies, αvβ6 was predominantly expressed in pancreatic ductal adenocarcinomas, followed by gastric carcinomas of the intestinal type, and intestinal adenocarcinomas, without a correlation of αvβ6 expression and its ligands, such as fibronectin and tenascin. Most interestingly, αvβ6 expression was elevated in well to moderately differentiated as compared to poorly differentiated tumors. This observation indicates a possible additional role of αvβ6 during cancer differentiation of gastroenteropancreatic adenocarcinomas [51].
In addition, we wish to stress here, that the integrins αvβ3 and α5β1 work in concert with growth factors and their receptors in order to influence differentiation processes in cancer stem cells in several cancer entities [75,76,77]. Cancer stem cells display increased expression of αvβ3 on their membranes. The ligation by vitronectin results in the loss of nuclear β-catenin and the downregulation of genes that contribute to cancer stem cell maintenance [78]. To this end, strategies to target αvβ3 and downstream signaling molecules, such as K-Ras are currently under investigation with the aim of interfering with cancer stemness [79]. Moreover, αvβ3 represents an indicator for chemosensitivity predicting the response of tumor cells to chemotherapeutics [80,81].
As some malignancies are associated with viral infection, αvβ6 and αvβ8 interestingly plays a crucial role during the entry of the foot-and-mouth-disease virus, the Herpes simplex virus (HSV), and the Epstein-Barr viruses. Virus entry into the cells depends on activated viral glycoproteins. HSV infect cells via species-specific glycoproteins and the conserved apparatus gH/gL and gB. Here, HSV uses αvβ6 or αvβ8 as gH/gL receptors [82]. Furthermore, the fusion of epithelial cells with Epstein-Barr virus proteins can be triggered by the binding of viral glycoproteins gH/gL to αvβ8 [83,84].
In the subsequent sections, we will address the characteristics and tumor biologically relevant functions of specific integrin subtypes and their interactions with distinct signaling networks.

3.1. Integrin-Mediated Cell Adhesion, Migration, and Invasion

Epithelial-type, anchorage-dependent tumor cells need to establish adhesive cell-cell and cell-ECM contacts with their microenvironment in order to evade anoikis, a special form of apoptosis attributable to the loss of cell-cell and cell-ECM contacts.
Normal epithelial cells acquire migratory capacities exclusively during embryonic development, tissue renewal and during tissue remodeling, e.g., during wound healing. In contrast, during transformation, tumor cells gain the ability to cross tissue boundaries and to invade into the surrounding ECM and the adjacent vasculature by loosening cell-cell and cell-ECM contacts. This is a prerequisite for the ability of cells to disseminate to distant body compartments and form metastases [85,86]. Hereby, effective cell migration of individual cancer cell types is influenced by the adhesive strength of their cell-cell and cell-ECM contacts [87]. As such, by altering the adhesive strength, the fine-tuned balance between increased and decreased cell motility might be tipped in order to move a cell´s body forward [88]. Hereby, integrins are also involved in cell shape alterations, which, at least in part, depend on integrin clustering and actin filament polymerization at the leading cell front of the migrating tumor cell. This is achieved by integrin cooperation with small guanosine-5′-triphosphate (GTP)-binding proteins, such as Rho and Rac [89].
The acquirement of persistent adhesive and migratory properties involves the epithelial-mesenchymal transition (EMT), which results in the loss of epithelial cell polarity and the formation of an elongated fibroblastoid cell morphology. During EMT, the expression of the cell-cell adhesive protein E-cadherin is decreased concomitant with the elevation of vimentin, desmin, and other mesenchymal markers [90,91].
Breast cancer cells utilize EMT to facilitate their invasion into the vasculature and to establish tumor metastases at distant sites [92]. In the following sections, the most important RGD-recognizing integrin subtypes and their roles in communicating cellular adhesion, migration and invasion, will be considered.

3.1.1. Integrin αvβ3 and α5β1

Integrin αvβ3 was first identified by Ruoslahti and coworkers [93]. Because of its predominant ECM ligand, it was originally named vitronectin receptor. In later studies, however, αvβ3 turned out to be a highly promiscuous integrin that binds to a plethora of different ECM proteins, among them fibronectin, osteopontin, and laminin, via the RGD motif, thereby triggering integrin signaling. The need for αvβ3-provoked signaling in order to induce cell migration has been reported for many different cell types, including smooth muscle cells [94], endothelial cells [95], and various tumor cells [96]. In human ovarian and breast cancer cells, enhancement of αvβ3/vitronectin-mediated cell adhesion was shown to be indispensable for the provision of a sufficient grip of cells to the underlying ECM and the necessary cytoskeletal rearrangements [97,98,99]. Indeed, tumor cells exhibiting high αvβ3 expression levels are capable of protruding broad lamellipodia, associated with decreased RhoA activity [100,101]. With regard to the regulation of cell contractility, cell/ECM stiffness, and cell movements, αvβ3 needs to cooperate with α5β1 in response to applied tension [87,102]. In prostate cancer cells, αvβ3 expression also provoked increased chemokine receptor expression, leading to enhanced migration/invasion during tumor cell metastasis to the bone [103].
Similar observations were made for integrin α5β1. Breast cancer cells with high α5β1 levels have revealed a 3-fold increased cell invasiveness, compared with cells exhibiting low α5β1 expression. However, invasiveness can be reduced by the myosin light chain kinase inhibitor ML-7, only in cells with a high α5β1 expression. This suggests a crucial role for α5β1 in the enhancement of cell migration and invasion by transmission and generation of contractile forces [104]. Moreover, the endocytic recycling of α5β1 also enhances cancer cell invasion, driven by RhoA and filopodial spike-based protrusions [105].

3.1.2. Integrin αvβ6

The expression of αvβ6 is initiated during embryonic development with high levels being exclusively restricted to epithelial cells, the developing lung tissue, and the kidney epithelia [106,107]. In a physiological context, αvβ6 is not constitutively expressed in differentiated epithelial cells, however, it becomes upregulated in the context of tissue remodeling, including wound healing and carcinogenesis [108]. The prevalence of αvβ6 expression has been described in several kinds of malignancies [46].
In addition to RGD-containing ECM ligands, such as fibronectin, tenascin-C, osteopontin, and vitronectin, αvβ6 binds to the RGD motif contained within the latency associated peptide (LAP) of TGF-β to activate the inactive, latent form of the TGF-β molecule. TGF-β signaling has a strong impact on EMT during invasive growth and cancer progression and is a crucial physiological process during embryonic development. In differentiated cells, however, EMT-related cellular alterations have also been observed during fibrotic tissue remodeling, wound healing, invasive cancer growth, and tumor metastasis. One hallmark of EMT is the downregulation of E-cadherin, which provokes loss of epithelial cell-cell adherence, followed by the detachment from the epithelial tissue and increased cell motility [109]. Because αvβ6 functions as a potent activator of TGF-β, a crucial role for αvβ6 expression during the EMT process has been suggested [46]. The activation of TGF-β1 by αvβ6 at the epithelial-mesenchymal interface seems to be a highly conserved EMT mechanism, first established during embryonic development, and then reactivated in the context of carcinogenesis. Cell alterations associated with EMT include a switch from αvβ1 and/or αvβ5 to αvβ6 and render αvβ6 an important EMT-marker for invasive cancer cells [42,110,111]. This switch enables epithelial cells to bind to ECM proteins, to cross tissue boundaries, and to invade mesenchymal tissue. Thereby, cancer cells evade apoptosis (anoikis), which under normal conditions would be provoked because of the loss of cell-cell and cell-ECM contacts [112,113]. Increased expression of αvβ6 in OSCCs triggers EMT, indicated by a decreased E-cadherin expression and the gain of vimentin expression [58]. Among the most common skin cancers, basal cell carcinomas (BCC) are caused by the deregulation of the Sonic hedgehog (Shh) signaling pathway. Although αvβ6 is expressed in low risk BCC, it is markedly upregulated in infiltrative BCC and promotes invasion by modulating the tumor stroma and activating the TGF-β1 pathway [67].
The induction of tumor cell-associated proteolytic activity and the downregulation of E-cadherin, as hallmarks of EMT, promote cancer cell migration, invasion, and tumor metastasis. Hence, the high invasive capacity of αvβ6-positive cancer cells is shown by the observation that αvβ6 expression is correlated with an elevated MMP-9 expression at the invasive cancer front [114]. MMPs are zinc-dependent metallo-endopeptidases involved in the degradation of ECM components [59]. Under physiological conditions, the basement membrane functions as a natural barrier for migrating/invasive cells. Elevated expression of MMP-9 degrades collagen type IV, one of the major components of the basement membrane. In ovarian cancer tissues and cultivated ovarian cancer cells, elevated αvβ6 expression is associated with the secretion of urokinase-type plasminogen activator (uPA), MMP-2, and MMP-9 [49]. Immunoprecipitation studies have revealed that αvβ6 directly interacts with the uPA receptor (uPAR) in ovarian cancer cells to promote cell migration and ERK activation. The interaction of αvβ6 with uPAR is restricted to ovarian cancer cells accompanied with increased TGF-β1 [115]. In OSCCs, ligand binding to αvβ6 leads to the recruitment of the focal adhesion kinase (FAK) and the activation of the Raf-ERK/MAPK pathway, thereby inducing MMPs and promoting cell proliferation and experimental metastasis [116]. Integrin αvβ6 expression in OSCC also activates MMP-3, a crucial factor for ECM degradation and remodeling, the subsequent activation of further MMPs, cell migration, and invasion upon binding to fibronectin [58]. Thus, αvβ6 contributes to the remodeling of the ECM as a prerequisite for cell invasion, migration, and tumor metastasis. These cell-cell and cell-ECM interactions in OSCCs also involve EMT-associated events, such as vimentin expression and E-cadherin downregulation [58]. As in OSCC, invasive endometrial carcinomas have been characterized by elevated αvβ6 at the leading invasive front, concurrently with MMP expression and metastasis formation [70]. Several studies have described the influence of αvβ6 activation/inhibition on the MMP pathway together with its binding of RGD-containing ligands, such as fibronectin, tenascin-C, vitronectin, and the LAP of TGF-β1 or TGF-β3, respectively. Deregulated ECM degradation further promotes invasive growth and malignant progression, as abnormal ECM degradation products bind to integrins, leading to the initiation of integrin signal transduction. For example, the cleavage of fibronectin by MMP-9 results in a 120-kDa fragment that promotes the migration of αvβ6-positive tumor cells. In turn, the increased expression of MMP-9 and MMP-2 leads to a positive feedback loop that might further promote cancer development [117]. At the invasive front of OSCC, MMP-7, -9, -12, and high levels of αvβ6 have been described as prognostic markers [72]. The last 11 amino acids (EKQKVDLSTDC) of the cytoplasmic tail of the β6-subunit seem to be sufficient to activate MMP-9 and promote invasiveness. However, this process depends on specific integrin subunits, because the β3-subunit activates MMP-2 [118]. In OSCC cell lines the migration of αvβ6-positive cells attributable to MMP-9 activation depends on integrin binding to insoluble LAP, whereas soluble LAP downregulates the αvβ6-induced adhesion and migration of OSSC cells [119]. Accordingly, the overexpression of αvβ6 enables cells to migrate onto fibronectin with increased invasiveness because of MMP-9 upregulation [50,120]. The main signaling event, associated with αvβ6-dependent MMP activation involves the ERK/MAPK pathway [121]. Ligand binding of fibronectin to αvβ6 recruits the Fyn tyrosin kinase to the β6-subunit. This activates MMP-3, which, in turn, degrades the ECM. Thus, the interrelation of MMP and integrin expression provides a putative molecular mechanism for αvβ6 enhancement to promote metastasis of OSCC cells. Although the exact molecular mechanisms of the integrin signaling network still remain to be fully elucidated, evidence points to a major role of the ERK/MAPK and TGF-β1 pathway in MMP activation by αvβ6 in various cancer types [116].
A different aspect of integrin function during cancer development is provided by the role of αvβ6 during endocytosis. Its interaction with the hematopoietic lineage cell-specific protein-1 (HS1) associated protein (HAX-1) is responsible for the clathrin-mediated endocytosis of αvβ6, an event that promotes the invasive behavior of OSCCs [122]. A recent study further indicates that the clathrin or non-clathrin-mediated endocytosis of integrins affects cell-signaling pathways that control cancer progression [123]. Consequently, the interaction between HAX-1 and αvβ6 might be the first evidence that integrin-triggered endocytosis plays a crucial role during cancer development and progression.

3.1.3. Integrin αvβ8

A further RGD-recognizing integrin is represented by αvβ8. In accordance with its β6 counterpart, β8-integrin is a 100 kDa protein that exclusively heterodimerizes with the 130-kDa αv-subunit [124]. Integrin αvβ8 is far less studied than other members of the integrin αv-subfamily; however, it is not only structurally, but also functionally related to αvβ6. Nevertheless, its role in adhesion, invasion and migration remains to be clarified.
Its biological functions are mainly associated with its action as a potent activator of TGF-β1 [125]. Loss-of-function experiments have revealed that mice lacking αvβ6 and αvβ8 activity reproduce the abnormalities observed in TGF-β1 and TGF-β3 null mice [126]. Site-directed mutagenesis of latent TGF-β1 has demonstrated that the high affinity binding of αvβ8 to latent TGF-β1 is defined by Leu-218 following the RGD motif contained within LAP [127]. In contrast to αvβ6, αvβ8 has a shorter cytoplasmic tail and does not bind to actin. Therefore, the αvβ8-dependent activation of latent TGF-β1 from the large latent complex (LLC) have been suggested not to work by force, but to require proteolytic cleavage by co-expressed MMP-14 (or MT1-MMP) [128]. The control of the subsequent intracellular transduction of αvβ6 signaling is gradually being elucidated. So far, the Band 4.1B and the Rho GDP Dissociation Inhibitor 1 (RhoGDI1) have been identified as β8-binding proteins [129] and the cytoskeletal adapter protein spinophilin has been described to regulate tumor cell invasion in glioblastomas [130]. Interestingly, during wound healing, TGF-β is activated via αvβ8 and αvβ6. In contrast to αvβ6, the inhibition of αvβ8 enhances the degree of cell motility [131].
Moreover, αvβ8 plays a dominant role in promoting the migration of astrocytes onto vitronectin [132]. Astrocytic αvβ8 expression has been proposed to act as a central regulator of brain vessel homeostasis because of the regulation of TGF-β activation and downstream genes that promote vessel differentiation and stabilization, such as the plasminogen activator inhibitor-1 (PAI-1) and thrombospondin-1 [133].

3.2. Impact of Integrins on Cellular Proliferation

In normal cells, distinct mechanisms regulate and terminate integrin-triggered signaling pathways to keep cell biological events under a balanced control and to maintain cells anchored onto the ECM in a quiescent state. Under normal conditions, integrins are essential players for the maintenance of tissue integrity. In contrast, tumor cells are characterized by uncontrolled behavior and acquire self-sufficiency regarding the expression and action of growth factors and their receptors, cytokines, and the activation of oncogenes, leading to the loss of control over cell proliferative activity in their pericellular microenvironenment. Hanahan and Weinberg have defined the hallmarks of cancer as (i) continuous proliferation, (ii) self-sufficiency, (iii) invasion and metastasis, (iv) limitless replicative potential, (v) promotion of angiogenesis, and (vi) evasion of apoptosis [134].

3.2.1. Integrin αvβ3

The growth-promoting effects of αvβ3 on breast and prostate cancer cells and on malignant melanoma have been well characterized [20]. In ovarian cancer, αvβ3 is instrumental in increasing cell proliferation by means of signaling that involves integrin-linked kinase [99]. Consequently, the blocking of αv results in drastic cell cycle arrest [135].
During their functions in cell proliferation, integrins crosstalk with growth factors and their respective receptors [136]. In breast and pancreatic cancer as well as glioma cells, αvβ3 physically and functionally cooperates with, for example, the epidermal growth factor receptor (EGF-R), Erb-B2, and the platelet-derived growth factor (PDGF-R) whose activation drives cell proliferation [137,138,139]. In ovarian cancer cells, αvβ3 expression correlates with enhanced expression and activity of the EGF-R [140]. In this way, even in the absence of respective ligands, growth factor receptor and integrins synergistically induce cellular signaling cascades. In cancer cells, constitutively activated integrin signaling not only leads to dysregulated tumor cell growth, but also avoids apoptosis. The effects of induced cell proliferation by αvβ3 is even further enhanced by the action of anti-apoptotic proteins Bcl-2 and FLIP [90].
TGF-β signaling is well known for its anti-proliferative effects. However, its role in tumor biology is heterogeneous and depends on the oncological context. TGF-β function can switch from inhibition of cell proliferation as a tumor suppressor to activation of a pro-oncogenic EMT-program. The effect of TGF-β signaling is influenced by the abundance of active TGF-β molecules. Here, RGD-binding integrins play a crucial role by binding to LAP and activating the latent form of the TGF-β molecule. The consequences will be explained and exemplified in the following based on the prevailing data for αvβ6 [125].

3.2.2. Integrin αvβ6

The binding of LAP to αvβ6 releases and activates the inactive pro/LAP-TGF-β [141,142]. Hence, αvβ6-positive cancer cells become self-sufficient in growth signals. In a study of cervical cancer patients, those with positive αvβ6 expression had different expression levels of p53, PCNA, Ki-67, and TIPE2 as proliferation markers. Thus, αvβ6 was described to cause active proliferation but to inhibit apoptosis, at least in cervical cancers [143].

3.2.3. Integrin αvβ8

In contrast to the effects of αvβ6 on pro/LAP-TGF-β, the activation of TGF-β by αvβ8 has interestingly been described to inhibit the proliferation of the airway epithelium in intact bronchial tissue [144]. In this manner, the β8-subunit seems to regulate the growth of epithelial cells and has been suggested to inhibit tumor growth in nude mice [145]. Interestingly, integrins seem to play heterogenous roles in regulating cell proliferation: growth-promoting integrin signals may be counterbalanced via inhibitory integrin signaling pathways, which can be activated by cytoplasmatic integrins domains as the results of either alternative splicing or evolutionary divergence [146].
Integrin αvβ8 has been suggested to regulate proliferation negatively for three main reasons. First, the cytoplasmic part of the β8-subunit differs in sequence, compared with other strongly related cytoplasmic domains of αv-associated integrin subunits, such as β1, β3, β5, and β6. Second, the β8 cytoplasmic domain cannot support stable, high-affinity adhesion to vitronectin, and third, β8 expression is restricted and mainly upregulated in non-proliferating cell types [145]. However, the overexpression of αvβ8 facilitates the proliferation and invasion of OSCC cells via the MEK/ERK signaling pathway upon the binding of αvβ8 to collagen type-1 [147].
In the context of chronic obstructive pulmonary disease (COPD), TGF-β activation via αvβ8 is known to enhance interleukin (IL)-1β-dependent fibroblast expression of the ccr6 chemokine ligand ccl20, suggesting that αvβ8, ccl20, and ccr6 interaction leads to the accumulation of dendritic cells (DC) around airways [148]. The convergence of the TGF-β and the IL-1β signaling pathways on the ccl20 promoter has been defined as a mechanism by which the αvβ8-mediated activation of TGF-β regulates IL-1β-dependent ccl20 expression in COPD patients [149]. The αvβ8-mediated activation of TGF-β regulates the chemokine secretion of lung fibroblasts, which directs DC and regulates fibrotic and immune responses in the lung [150]. Interestingly, T regulatory cells (tTregs) lacking αvβ8 are unable to suppress pathogenic T-cell responses during active inflammation; however, the deletion of αvβ8 does not result in a spontaneous inflammatory phenotype [151]. For thymus-derived tTregs, αvβ8 is referred to as a marker protein that mediates the processing of latent TGF-β1 from the latent TGF-β1/glycoprotein-A repetitions predominant protein (GARP) complex on the surface of tTregs [152].

3.3. Integrin Effects on Cell Survival and Apoptosis

The integrin-mediated anchorage of epithelial-type cells to the ECM is instrumental for maintenance of cell viability via the activation of FAK-, the PI3K/Akt, and the FAK-MAPK-pathways. In contrast, unligated integrins may eventually lead to cellular apoptosis (integrin-mediated death) [153,154]. Tumor cells derived from epithelia, even losing their ECM contacts, have the ability to overcome controlled cell death (anoikis), e.g., by increasing the expression of receptor tyrosine kinases or small guanosine-5′-triphosphate (GTP)ases in concert with downregulated caspase 8 expression [155,156]. Moreover, tumor cells may circumvent anoikis by modulating the expression pattern and cell surface density of their integrin subtypes and by internalizing activated integrins with sustained signaling in endosomes [111,157,158]. Cancer cells also manage to escape from anoikis by integrin crosstalk with growth factor receptors, resulting in effective survival signaling [159,160]. Interestingly, αvβ3-expressing tumor cells, which are detached from the ECM, nevertheless exhibit enhanced tumor growth in vitro in an anchorage-independent fashion. The growth-promoting effect depends on the recruitment of c-Src to the β3-integrin cytoplasmic tail. This, in turn, activates c-Src, leads to the phosphorylation of the Crk-associated substrate (CAS), and promotes tumor cell survival in a mechanism independent of FAK activation. These findings have unraveled an interesting role of αvβ3 in anchorage-independent tumor cell growth and aggressiveness [21]. Loss of normal integrin action may also arise from abnormal integrin internalization. To this end, β3-integrin has been reported to associate with caveolin-1, indicative of raft/caveolar endocytosis. Following cell detachment, lipid rafts are rapidly internalized. This, in turn, inhibits key integrin signaling molecules such as Erk, PI3K, and Rac. In cancer cells, however, this process appears to be hindered by β3-integrin whereby integrin signaling is maintained [21]. In glioblastoma cells, which overexpress αvβ3 at the invasive tumor front, fibronectin expression is elevated accompanied by increased cell motility and resistance to apoptosis. Subsequently, it has turned out that apoptosis in glioblastoma cells is not regulated by integrins alone but by their interconnections to the p53 pathways [161,162]. It is observed that integrin α5β1 is more important for inducing apoptosis in glioblastomas than αvβ3 [161,162]. Loss of anchorage-dependence and consequently, resistance to anoikis, are extremely important features during malignant transformation of tumor cells, enhancing tumor development and metastasis [155,163].

3.4. Integrins αvβ3, αvβ5, and α5β1 in Tumor Angiogenesis

Tumor cells, like normal cells, strongly depend on the continuous supply of nutrients and oxygen in order to maintain their vital features. Further outgrowth of tumors above a tumor size of approximately 2 mm in diameter is largely restrained by the absence of functional blood vessels in the tumor vicinity. Therefore, tumor cells are forced to induce the formation of new vessels, an event involving the three endothelial integrins, αvβ3, αvβ5, and α5β1 [164,165]. This is achieved by the secretion of pro-angiogenic molecules, such as the vascular endothelial growth factor (VEGF) and the fibroblast growth factor-2 (FGF2), which upon engagement of their respective receptors leads to the so-called “angiogenic switch” [166]. VEGF-mediated angiogenesis occurs via αvβ3, whereas FGF-2-mediated angiogenesis is mainly triggered by αvβ3 and α5β1 [167]. Signaling by these receptor/ligand systems contributes to the formation of new vessels, derived from the surrounding pre-existing vasculature. Hereby, per se quiescent endothelial cells are activated to proliferate and migrate towards the tumor site. This motile activity depends on the remodeling of the ECM, predominantly provoked via proteolytic ECM breakdown by MMPs and uPA. Integrin αvβ3 co-localizes and assists during the activation of MMPs on the surface of angiogenic blood vessels. This provokes the remodeling of collagen and the basement membrane. The consecutive ECM breakdown then facilitates endothelial cell migration [168,169]. In accordance, pancreatic tumor cells have been shown to exhibit increased αvβ3 levels that correlated with enhanced MMP-2 activation and lymph node metastasis [65].
One predominant trigger for tumor angiogenesis becomes relevant as soon as tumor growth exceeds blood supply and causes a hypoxic gradient. Hypoxia induces the translocation of the hypoxia-inducible factors-1 α (HIF-1α), HIF-2α, and HIF-3α into the nucleus. Here, they recognize hypoxia response elements (HREs) and induce gene transcription upon binding to the respective promoter regions of their many target genes, including αv, β3, β1, and β6, and the integrin-linked kinase (ILK) and VEGF [170,171]. In turn, these integrins enhance a migratory and invasive tumor cell phenotype that establishes a positive feedback loop and furthers EMT and tumor metastasis [172,173,174,175,176,177,178,179].
In animal models, the blockade of αvβ3 and αvβ5 by specific peptide antagonists, such as Cilengitide, or by blocking antibodies, such as Vitaxin, results in the drastic reduction of angiogenesis. This blockade also inhibits downstream signaling via the FAK/Src/Akt-pathway, leading to endothelial and tumor cell apoptosis [180]. Cilengitide is effective in combination with chemotherapy [181]. However, studies in αv-, as well as β3/β5-knock-out mice indicate that αvβ3 is important but not indispensable for tumor angiogenesis. In addition, β5 does not represent an absolute requirement for this process, because extensive angiogenesis and tumor growth were still observed [182,183,184]. Thus, α5β1 has been proposed also to take over as yet not fully resolved functions during angiogenesis, especially since Hynes and co-workers have shown that αv and α5 cooperate and are capable of functionally backing each other up [185]. Interestingly, low dose administration of the αvβ3 ligand activated this integrin [182,186,187] and opens a new way in stabilization of the vascular system in the treatment with cytotoxic drugs [188].
As regards αvβ8 expression, it activates a TGF-β gradient in the brain, a gradient that is angio-suppressive and inhibits endothelial cell sprouting. Consistently, the loss of αvβ8 in the brain or downstream TGF-β1 signaling via ALK5-Smad3 in endothelial cells increases vascular sprouting, branching, and proliferation, resulting in vascular dysplasia and hemorrhage [189]. However, in a developmental context, αvβ8, βPix, and GIT1 have been suggested to regulate vascular stability, cerebral angiogenesis, and endothelial cell proliferation in the developing embryo [190]. In the developing retina and on astrocytes αvβ8 expression is essential for neovascularization and regulates blood vessel sprouting [191]. Astrocytoma cells selectively suppress αvβ8 expression to manipulate their antigenic balance in order to exploit the signaling pathways of developmental brain angiogenesis in adult brain tumors [192]. The reduced expression of the β8-chain had been considered to be instrumental for the pathogenesis of sporadic brain arteriovenous malformations [193].

3.5. TGF-β1 and Its Integration with Integrin Signaling

One major inducer of EMT is TGF-β, which is produced at high levels in breast, prostate, and other cancer cells [194]. The TGF-β family comprises three isoforms: TGF-β1, TGF-β2, and TGF-β3. Each member is processed as a homodimeric pro-TGF-β. This cytokine is deposited in the ECM in an inactive form linked to LAP and one of four latent TGF-β-binding proteins (LTBP) [195]. Upon the binding of αvβ3, αvβ6, or αvβ8 to the RGD motif contained within LAP, a conformational switch is provoked, leading to the exposure of TGF-β1 to the neighboring cells where it ligates its respective receptors. This provokes the downregulation of E-cadherin and increases the expression of mesenchymal proteins, such as N-cadherin [196].
Pro-TGF-β1 monomers contain an aminoterminal pro-domain of 249 amino acid residues, separated by a pro-protein convertase cleavage site from the 112-amino acid residue carboxy-terminus of the TGF-β1 domain. It is cleaved intracellulary by furin-like enzymes and remains non-covalently linked to LAP in order to form the small inactive latent complex (SLC) [197]. The binding of LAP to the mature TGF-β molecule prevents signaling through the TGF-β-receptors [198]. In order to enable receptor binding, TGF-β1 has to be activated and released from this complex. However, TGF-β1 is mostly secreted as a large latent complex (LLC) formed by the dimerization and covalent disulfide linkage of the LAP-TGF-β to LTBPs or GARP [199]. Each pro-domain forms a ring around TGF-β to keep it inactive/latent [200]. LTBPs belong to the LTBP/fibrillin protein family, comprising fibrillin-1-3 and LTBP1-4 [201]. Except for LTBP3, the LTBP-isoforms LTBP1, LTBP2, and LTBP4 can bind to all TGF-β isoforms. LTBPs have a binding capacity towards fibronectin and vitronectin, thus in the LLC, TGF-β is bound to ECM proteins. This complex building provides in the ECM a pool of latent TGF-β that is activated in a context-dependent manner. In contrast to other mechanisms, such as proteolysis, this way of TGF-β1 activation requires cell traction forces and the association of the LLC with the ECM [202]. Thereby, the binding of αvβ6 to LAP directly influences the migration of cancer cells. A recent crystal structure analysis of the αvβ6 head group bound to pro/LAP-TGF-β1 has provided insights into the interaction between integrins and their macromolecular ligands. Furthermore, it elucidated the way in which integrin binding to ECM molecules transmits activating forces with biological consequences, such as TGF-β1 activation [7]. Interestingly, the resulting effects on the integrin domains by small molecules and RGD-containing peptide binding depend on the force vector and the physiological orientation of the integrin dimer and their ligands.
The activation mechanism of TGF-β1 by αvβ6 has been thoroughly studied [119,141,203]. Upon binding of αvβ6, αvβ3, or αvβ8 to the RGDLXX(I/L) motif within LAP of pro-TGF-β1, these integrins associate with the actin cytoskeleton. This triggers conformational changes in the LLC complex (LAP-TGF-β-LTBP1), releasing TGF-β1 from the LLC and enabling TGF-β1 binding to its receptor [198,199]. Interestingly, TGF-β1 activation by integrins is LTBP-specific, because αvβ6 can only activate TGF-β molecules that are bound to LTBP1 but not to LTBP3. The various isoforms of the LTBPs exhibit significant differences in the sequence of the hinge region, suggesting a functional role in the specific activation of TGF-β1 by the various integrin subtypes [203]. Although LAP-TGF-β1 also binds to other integrins, such as αvβ1 and αvβ3, only αvβ6 can release and activate TGF-β1 from the LLC [204]. However, the mere binding of integrin to LAP is not sufficient to release and activate TGF-β1 in vivo [200]. This is explainable by the different cytoskeleton-binding properties that influence the binding of the cytoplasmic integrin tail to actin and exert traction forces. Recent data have revealed that traction forces by αvβ6 on pro/LAP-TGF-β1 is required for TGF-β1 activation: truncation of the cytoplasmic domain of the actin-binding β6-subunit or deletion of the binding site of the pro-domain of LAP-TGF-β1 to the ECM inhibits the exertion of tensile force across the pro-domain and thus TGF-β1 activation [205].
Upon TGF-β1 activation, the successive signaling events in turn lead to the additional expression of αvβ3 and αvβ6 and of ECM protein ligands, followed by the activation of the PI3K-, Akt, and NF-κB pathways. Moreover, growth factor receptors, such as the EGF-R, are upregulated upon TGF-β signaling; this is of note here, since the EGF-R physically and functionally crosstalks with integrins in a synergistic fashion [127,204,206,207,208,209,210,211,212].

4. Challenges for the Design of Novel Integrin Ligands and Their Translation into Clinical Applications

The great discovery that the small tripeptide sequence RGD is sufficient to inhibit the interaction of ECM proteins with integrins [8] has stimulated the search for small peptidic or peptidomimetic molecules to selectively address various integrin subtypes and thus their pathophysiological features with high affinity [27,28,213,214]. This culminated in the development of the peptidic compound Cilengitide (4a) (Figure 3) [215,216], which was tested in phase III clinical studies for the treatment of glioblastomas [217,218]. The failure of these studies was a backlash for the development of integrin ligands [219]. However, recent investigations into the mechanism and details of the activities of the various integrin subtypes and their role in apoptosis have shed light on the reasons for this failure and have opened new exciting applications for this molecule [182]. The binding of small ligands can occur to the resting state of an integrin receptor leading to its activation. However, a stronger binding to the focal adhesion complex requires higher concentrations of specific ligands [186].

5. Improving the Activity and Selectivity of Integrin Ligands

Although the above-mentioned integrins all recognize RGD, they differ in their binding activity towards different ECM proteins. Integrins of the αv-family are well known for their strong binding to vitronectin [225]. Very early, the activity and selectivity of peptide ligands to the different integrin subtypes was observed to be controllable by cyclization [226,227], RGD flanking residues [228], the chirality of the amino acids, and the N-methylation of peptide bonds [215,226,227,228,229]. The substitution of arginine by lysine, which removes ligand binding to all αv-integrins, retains, however, the activity for the platelet receptor αIIbβ3 [27,230]. This knowledge and the structural data concerning integrin head groups with and without a ligand have enabled the development of many derivatives with improved integrin binding properties, including peptide mimetics and finally have also led to non-peptidic ligands. X-ray structures of αvβ3 [29,30], αIIbβ1/αIIbβ3 [231], α5β1 [232], and αvβ6 [7] have facilitated the rational design and optimization of ligands. For this aspect of integrin ligand design, please refer to previously published literature [27,213,214]. These selective ligands for distinct integrin subtypes bear the potential of serving as promising new drug candidates for personalized medicine in various cancer entities. Figure 3 depicts a few selected integrin ligands resulting from an extensive effort to develop RGD-peptides or low-molecular RGD-mimicking peptidomimetics, featuring high stability for diagnostic and therapeutic in vivo applications. To this end, the ligands have to be functionalized which is frequently accompanied by a decline in integrin-binding affinity. Hence, after functionalization, the binding affinity of a newly developed integrin ligands has to be carefully controlled and may be minimized by a well-considered selection of the labeling site and an appropriate spacer moiety being placed in between [214,215,220,221,222,223,224,233].
Moreover, during ligand development, we must avoid making the specificity of integrin ligands to be used for therapeutic purposes towards the platelet integrin αIIbvβ3 too narrow, since their systemic administration might lead to major hemorrhagic disorders [26,27,28].

6. In Vivo Targeting of Integrins for Cancer Imaging and Therapy

Because of their different functions in cancer biology and the availability of small molecule ligands, RGD-binding integrins have been identified as attractive in vivo targets for the molecular imaging of tumors [107]. The respective contrast agents are usually based on the peptidic or non-peptidic integrin ligands, which are conjugated to suitable labels, e.g., fluorophores, radionuclides, or MRI contrast agents, such as gadolinium complexes. Moreover, integrin ligands have been incorporated into larger structures, e.g., by grafting them onto the surface of nanomaterials, or by their inclusion into liposomes [234,235,236].

6.1. Integrin αvβ3 and α5β1

A glance at the currently available literature reveals that the overwhelming majority of pertinent studies is focused on αvβ3. Targeted probes comprising αvβ3-selective cyclic pentapeptides of the RGDxK (x = f, y) type are popular, most likely because the c(RGDxK) motif can be relatively easily synthesized. Thus, they quickly entered the portfolios of producers of fine chemicals as an off-the-shelf building block. Mice models for αvβ3-expressing tumors, such as subcutaneous U87MG glioma or M21 melanoma xenografts, are routinely generated. The imaging of αvβ3 in rodents by using modified RGD peptides has evolved into a standard benchmark scheme for in vivo feasibility studies and the validation of novel bioconjugation and labeling approaches. Hence, a vast number of αvβ3-targeting probes has been described, whereas only a small fraction has actually been pursued beyond the proof-of-principle stage. Notwithstanding this, radiolabeled probes for mapping of αvβ3 expression are widely considered to possess significant clinical potential for tumor imaging. The encouraging results of clinical applications of the first c(RGDfK)-based positron emission tomography (PET) radiotracer 18F-galacto-RGD in cancer patients have triggered the development of numerous αvβ3-targeted radiopharmaceuticals for imaging of tumor angiogenesis [237,238,239]. A recent review by Chen et al. [240] summarizes human PET images and additional data for eight different compounds [241,242,243,244,245,246,247,248], some of which are currently in clinical trials, highlighting their capability to visualize solid tumors. However, as has been repeatedly pointed out, the initially envisaged purpose of imaging αvβ3, i.e., the in vivo quantification of the tumor angiogenesis and/or the patient stratification for antiangiogenic therapies, is somewhat thwarted by the finding that αvβ3 expression does not necessarily correspond to angiogenic activity in (tumor) tissues [165,240,249]. Many tumor cell types display membraneous or cytoplasmic αvβ3 expression, but also substantial physiological expression is observed in many organs. As a result, the clinical value of the in vivo mapping of αvβ3 remains to be defined [249].
In contrast to αvβ3, which is not strictly required for tumor angiogenesis [183,184,185], α5β1 suggests itself as a more attractive target, because of the evidence for a stricter link between angiogenesis and β1 integrin expression [250]. Although α5β1 is only weakly expressed in quiescent murine and human endothelial cells [251], it is upregulated on endothelial cells during vessel sprouting during the tumor angiogenic process [167]. Despite this promising perspective, only few imaging probes for α5β1 have been described so far. The first in vivo imaging of α5β1 expression was performed by using 68Ga-labeled derivatives of a α5β1-selective peptidomimetic [252,253], followed by radiolabeled derivatives of a linear peptide derived by phage display [254] and of a cyclic peptide comprising the HisoDGR structural motif [255]. The best sensitivity and contrast for α5β1 during PET imaging was hitherto achieved by using 68Ga-Aquibeprin [256,257], a trimer of the aforementioned peptidomimetic [252] with an α5β1 affinity (IC50) of 80 pM. However, to the best of our knowledge, no α5β1 imaging in humans has been reported, as yet.
Although in the context of molecular imaging, the ubiquity of c(RGDxK)-based probes has led to the common perception of an equivalence of integrin expression and angiogenesis, this distorted notion might soon be adjusted by the recent advent of imaging agents, targeting αvβ6 [258].

6.2. Integrin αvβ6

Several characteristics qualify αvβ6 as a potential diagnostic and therapeutic target. First, it is not constitutively expressed under physiological conditions in differentiated adult epithelia but within a specific pathological context upon cellular alterations going along with EMT, i.e., wound healing, carcinogenesis, and tumor metastasis. Second, the use of targeting diagnostic and therapeutic tools should allow a clear demarcation between healthy and diseased tissue. As an EMT marker, αvβ6 is specifically expressed at the epithelial-mesenchymal boundary and thus enables the specific targeting of invasive and migrating cells. Consequently, because of its high affinity and its de novo expression in cancer tissue, its involvement in cellular invasion and metastasis, and its expression in fibrotic tissue, αvβ6 represents a promising cancer cell target [47,58,259]. To this end, increasing efforts are currently being undertaken to synthesize—besides ligands for targeting other integrins—αvβ6-specific peptidic and peptidomimetic integrin ligands/antagonists [260,261,262,263,264]. For the diagnostic imaging of human cancer cells, the peptide A20FMDV2, derived from the food-and-mouth disease virus sequence [262], the peptide H2009.1 [263], and the cyclic peptide S02 [264] are currently being explored for their use as radiolabeled integrin ligands. The idea of targeting RGD-binding integrins for tumor imaging has been followed extensively, however, mainly for αvβ3 [265]. Nevertheless, αvβ6 has been imaged in vivo by single-photon emission computed tomography (SPECT) [258,259,260,261,262,263,264,265,266,267,268] and PET [262,269,270,271,272,273].
To date, a series of various αvβ6-targeting tracers have been developed for diagnostic purposes, including linear 10- to 20-mer peptides and “stapled” cystine peptides [266,267,268]. Our group have developed and tested enzymatically stable cyclic peptides as novel αvβ6 ligands, revealing so far the lowest molecular weight of all αvβ6-binding molecules with sub-nanomolar binding affinity [223]. Recent work has demonstrated that these novel ligands can serve as the basis for the synthesis of promising new PET-tracers [273]. 68Ga-Avebehexin [271], a 68Ga-labeled derivative of the metabolically stable cyclic nonapeptide cyclo-(FRGDLAFp(NMe)K) [223], has shown the most advantageous tumor-to-background contrast because of excellent renal clearance (Figure 4). Especially for cancer entities, which exhibit high αvβ6 expression in more than 95% of cases, like in OSCC, its imaging encourages a novel and most promising tool for cancer diagnosis [47,262].
Until now, only a single study on αvβ6 targeting carried out in living human subjects has been reported. Using the 68Ga- and 177Lu-labeled compound SFITGv6, which comprises the binding sequence FRGDLMQL, Altmann et al. performed PET/CT scans of head and neck squamous cell carcinoma (HNSCC) and non-small-cell lung carcinoma (NSCLC) patients and found that the tracer accumulated specifically in tumors, but not in inflammatory lesions [274]. Finally, αvβ6-targeted imaging and therapy has been repeatedly pointed out to hold greatest promise for pancreatic carcinoma (PDAC) [51,73,267,269,275], evidence in humans is still lacking.
As examples, Figure 4 shows µPET images (maximal intensity projections) of SCID mice bearing subcutaneous tumor xenografts on the right shoulders (positions indicated by arrows). Left: M21 human melanoma with high α5β1 expression, imaged using 68Ga-Aquibeprin [256,257]. Right: H2009 human lung adenocarcinoma with high αvβ6 expression, imaged using 68Ga-Avebehexin [73,275].

7. Integrins in Cancer Therapy

Binding molecules with high affinity and selectivity against specific integrin subtypes can serve as key pharmacological tools for studying the biological functions of integrins. We suggest that a detailed analysis of the correlation of integrin subtype expression with cancer progression and the understanding of the underlying molecular mechanisms will help to generate molecular diagnostic and therapeutic tools to improve cancer patient care.
This knowledge will provide the basis for the clinical translation to diagnostic and therapeutic applications. Therefore, increasing efforts are currently being made to synthesize also meanwhile αvβ6-specific non-peptidic and peptidic integrin ligands with antagonizing/inhibiting effects [260,261,262,263,264].
The interruption of integrin-specific functions and signaling by specific integrin ligands has been considered as a promising potential therapeutic approach. This aspect will be exemplified here for αvβ6: the US patent 7150871 describes the successful reduction of metastasis in αvβ6 overexpressing lung cancer cells upon the specific binding of the αvβ6-specific function blocking monoclonal antibody 10D5 [47]. Moreover, in colon cancer cells, this antibody provokes increased cellular apoptosis, accompanied by enhanced ERK phosphorylation [276]. The binding of 10D5 to β6 has been suggested to disrupt binding to ERK and thus reduce metastasis. The major capsid protein VP1 of the O1 strain of the foot-and-mouth-disease virus is well known as a high affinity ligand for αvβ6. A 17-mer peptide derived from the viral coat protein-1 VP1 has been used to create the humanized single chain antibody scFv B63 that can be used to interfere with αvβ6-mediated cancer cell invasion [277]. As described above, small molecule peptidic and peptidomimetic integrin ligands/antagonists are also being explored for their efficacy in inhibiting specific integrin functions, tumor progression, and metastasis. Moreover, the inhibition of the binding of LAP-TGF-β1 to αvβ6 may represent a specific and context-dependent therapeutic approach for αvβ6-positive neoplasia.

8. Conclusions

The context-specific expression and selective functions of RGD-recognizing integrins open the possibility for targeting distinct physiological and pathological processes, including cell proliferation, differentiation, apoptosis, adhesion, and migration. All these processes are involved in cell invasion and angiogenesis. Selective ligands for the various integrin subtypes are the key for targeted diagnostic approaches, such as molecular imaging and drug-based therapeutic approaches. Potential clinical applications range from oncologic, to fibrotic and to inflammatory diseases management. Future studies in translationally orientated pre-clinical models and clinical trials will provide insights as to which application might be of clinical relevance. Current clinical trials are addressing the detection of integrin αvβ6 in pancreatic cancer to evaluate the feasibility of [18]FP-R01-MG-F2 PET/CT imaging in patients afflicted with pancreatic cancer (NCT02683824) [278]. A further study will provide data considering the side effects of 18F-αvβ6-binding-peptide and its feasibility for imaging in patients with primary tumors or cancer that has spread to the breast, colon, lung, or pancreas, in order to improve the detection of the cancer location within the body (NCT03164486) [279].

Acknowledgments

The authors declare no conflict of interest. The work had no specific funding. There are no financial disclosures from any authors.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

RGDArg-Gly-Asp
BMBasement membrane
BMPBone morphogenetic protein
CAMCell adhesion molecule
CDCrohn’s disease
COPDChronic obstructive pulmonary disease
DCDendritic cells
ECMExtracellular matrix
ERKExtracellular signal regulated kinase
FAFocal adhesion
TGF-βTransforming growth factors-β
MAPMitogen-activated protein kinases
PI3KPhosphoinositide kinase
GTPGuanosine-5′-triphosphate
EGF-REpidermal growth factor receptor
VEGFVascular endothelial growth factor
PDGF-RPlatelet-derived growth factor
FAKFocal adhesion kinase
CASCrk-associated substrate
LAPLatency-associated peptide
MMPMatrix metalloproteases
uPAUrokinase-type plasminogen activator
HREsHypoxia response elements
HAX-1Hematopoietic lineage cell-specific protein-1 (HS1) associated protein
HIFHypoxia-inducible factor
OSCCOral squamous cell carcinoma
BCCBasal cell carcinomas
ShhSonic hedgehog
SPECTSingle-photon emission computed tomography
tTregsT regulatory cells
SLCSmall latent complex
LLCLarge latent complex
LTBPLatent TGF- β binding proteins
FMDVFoot-and-mouth disease virus
HNSCCHead and neck squamous cell carcinoma
NSCLCNon-small-cell lung carcinoma
PDACPancreatic carcinoma
SPECTSingle photon emission computed tomography
PETPositron emission tomography

References

  1. Horwitz, A.R. The origins of the molecular era of adhesion research. Nat. Rev. Mol. Cell Biol. 2012, 13, 805–811. [Google Scholar] [CrossRef] [PubMed]
  2. Zhu, J.; Zhu, J.; Springer, T.A. Complete integrin headpiece opening in eight steps. J. Cell Biol. 2013, 201, 1053–1068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Calderwood, D.A. Integrin activation. J. Cell Sci. 2004, 117, 657–666. [Google Scholar] [CrossRef] [PubMed]
  4. Mould, A.P.; Barton, S.J.; Askari, J.A.; Craig, S.E.; Humphries, M.J. Role of ADMIDAS cation-binding site in ligand recognition by integrin α5β1. J. Biol. Chem. 2003, 278, 51622–51629. [Google Scholar] [CrossRef] [PubMed]
  5. Van Agthoven, J.F.; Xiong, J.-P.; Alonso, J.L.; Rui, X.; Adair, B.D.; Goodman, S.L.; Arnaout, M.A. Structural basis for pure antagonism of integrin αvβ3 by a high-affinity form of fibronectin. Nat. Struct. Mol. Biol. 2014, 21, 384–388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Rahmouni, S.; Lindner, A.; Rechenmacher, F.; Neubauer, S.; Sobahi, T.R.; Kessler, H.; Cavalcanti-Adam, E.A.; Spatz, J.P. Hydrogel micropillars with integrin selective peptidomimetic functionalized nanopatterned tops for the separate measurement of cell traction forces transmitted through αvβ3- or α5β1-integrins. Adv. Mater. 2013, 25, 5869–5874. [Google Scholar] [CrossRef] [PubMed]
  7. Dong, X.; Zhao, B.; Iacob, R.E.; Zhu, J.; Koksal, A.C.; Lu, C.; Engen, J.R.; Springer, T.A. Force interacts with macromolecular structure in activation of TGF-β. Nature 2017, 542, 55–59. [Google Scholar] [CrossRef] [PubMed]
  8. Pierschbacher, M.D.; Ruoslahti, E. Cell attachment activity of fibronectin can be duplicated by small synthetic fragments of the molecule. Nature 1984, 309, 30–33. [Google Scholar] [CrossRef] [PubMed]
  9. Suzuki, S.; Oldberg, A.; Hayman, E.G.; Pierschbacher, M.D.; Ruoslahti, E. Complete amino acid sequence of human vitronectin deduced from cDNA. Similarity of cell attachment sites in vitronectin and fibronectin. EMBO J. 1985, 4, 2519–2524. [Google Scholar] [PubMed]
  10. Plow, E.F.; Pierschbacher, M.D.; Ruoslahti, E.; Marguerie, G.A.; Ginsberg, M.H. The effect of Arg-Gly-Asp-containing peptides on fibrinogen and von Willebrand factor binding to platelets. Proc. Natl. Acad. Sci. USA 1985, 82, 8057–8061. [Google Scholar] [CrossRef] [PubMed]
  11. Oldberg, A.; Franzén, D.; Heinegård, D. Cloning and sequence analysis of rat bone sialoprotein (osteopontin) cDNA reveals an Arg-Gly-Asp cell-binding sequence. Proc. Natl. Acad. Sci. USA 1986, 83, 8819–8823. [Google Scholar] [CrossRef] [PubMed]
  12. Grant, D.S.; Tashiro, K.-I.; Segui-Real, B.; Yamada, Y.; Martin, G.R.; Kleinman, H.K. Two different laminin domains mediate the differentiation of human endothelial cells into capillary-like structures in vitro. Cell 1989, 58, 933–943. [Google Scholar] [CrossRef]
  13. Aota, S.; Nomizu, M.; Yamada, K.M. The short amino acid sequence Pro-His-Ser-Arg-Asn in human fibronectin enhances cell-adhesive function. J. Biol. Chem. 1994, 269, 24756–24761. [Google Scholar] [PubMed]
  14. Pierschbacher, M.D.; Ruoslahti, E. Arg-Gly-Asp: A versatile cell recognition signal. Cell 1986, 44, 517–518. [Google Scholar]
  15. Ruoslahti, E.; Pierschbacher, M.D. New perspectives in cell adhesion: RGD and integrins. Science 1987, 238, 491–497. [Google Scholar] [CrossRef] [PubMed]
  16. Schittenhelm, J.; Klein, A.; Tatagiba, M.S.; Meyermann, R.; Fend, F.; Goodman, S.L.; Sipos, B. Comparing the expression of integrins αvβ3, αvβ5, αvβ6, αvβ8, fibronectin and fibrinogen in human brain metastases and their corresponding primary tumors. Int. J. Clin. Exp. Pathol. 2013, 6, 2719–2732. [Google Scholar] [PubMed]
  17. Huveneers, S.; Danen, E.H.J. Adhesion signaling-crosstalk between integrins, Src and Rho. J. Cell Sci. 2009, 122, 1059–1069. [Google Scholar] [CrossRef] [PubMed]
  18. Geiger, B.; Spatz, J.; Bershadsky, A.D. Environmental sensing through focal adhesions. Nat. Rev. Mol. Cell Biol. 2009, 10, 21–33. [Google Scholar] [CrossRef] [PubMed]
  19. Niu, J.; Li, Z. The roles of integrin αvβ6 in cancer. Cancer Lett. 2017, 403, 128–137. [Google Scholar] [CrossRef] [PubMed]
  20. Jin, H.; Varner, J. Integrins: Roles in cancer development and as treatment targets. Br. J. Cancer 2004, 90, 561–565. [Google Scholar] [CrossRef] [PubMed]
  21. Desgrosellier, J.S.; Cheresh, D.A. Integrins in cancer: Biological implications and therapeutic opportunities. Nat. Rev. Cancer 2010, 10, 9–22. [Google Scholar] [CrossRef] [PubMed]
  22. Sheldrake, H.M.; Patterson, L.H. Strategies to inhibit tumor associated integrin receptors: Rationale for dual and multi-antagonists. J. Med. Chem. 2014, 57, 6310–6315. [Google Scholar] [CrossRef] [PubMed]
  23. Niland, S.; Eble, J.A. Integrin-mediated cell-matrix interaction in physiological and pathological blood vessel formation. J. Oncol. 2012, 125278. [Google Scholar] [CrossRef] [PubMed]
  24. Goodman, S.L.; Picard, M. Integrins as therapeutic targets. Trends Pharmacol. Sci. 2012, 33, 405–412. [Google Scholar] [CrossRef] [PubMed]
  25. Goswami, S. Importance of integrin receptors in the field of pharmaceutical and medical science. Adv. Biol. Chem. 2013, 3, 224–252. [Google Scholar] [CrossRef]
  26. Kiran Marelli, U.; Rechenmacher, F.; Ali Sobahi, T.R.; Mas-Moruno, C.; Kessler, H. Tumor targeting via integrin ligands. Front. Pharmacol. Anti-Cancer Drugs 2013, 3. [Google Scholar] [CrossRef] [Green Version]
  27. Meyer, A.; Auernheimer, J.; Modlinger, A.; Kessler, H. Targeting RGD recognizing integrins: Drug development, biomaterial research, tumor imaging and targeting. Curr. Pharm. Des. 2006, 12, 2723–2747. [Google Scholar] [CrossRef] [PubMed]
  28. Arndt, T.; Arndt, U.; Reuning, U.; Kessler, H. Integrins in angiogenesis: Implications for tumor therapy. In Cancer Therapy. Molecular Targets in Tumor-Host Interactions; Weber, G.F., Ed.; Horizon Bioscience: Norfolk, UK, 2005; pp. 93–141. [Google Scholar]
  29. Xiong, J.P.; Stehle, T.; Diefenbach, B.; Zhang, R.; Dunker, R.; Scott, D.L.; Joachimiak, A.; Goodman, S.L.; Arnaout, M.A. Crystal structure of the extracellular segment of integrin αvβ3. Science 2001, 294, 339–345. [Google Scholar] [CrossRef] [PubMed]
  30. Xiong, J.P.; Stehle, T.; Zhang, R.; Joachimiak, A.; Frech, M.; Goodman, S.L.; Arnaout, M.A. Crystal structure of the extracellular segment of integrin αvβ3 in complex with an Arg-Gly-Asp ligand. Science 2002, 296, 151–155. [Google Scholar] [CrossRef] [PubMed]
  31. Müller, M.A.; Opfer, J.; Volkhardt, L.A.; Brunie, L.; Sinner, E.-K.; Boettiger, D.; Bochen, A.; Kessler, H.; Gottschalk, K.-E.; Reuning, U. The glycophorin A transmembrane sequence within integrin αvβ3 creates a non-signalling integrin with low basal affinity that is strongly adhesive under force. J. Mol. Biol. 2013, 425, 2988–3006. [Google Scholar] [CrossRef] [PubMed]
  32. Gottschalk, K.-E.; Kessler, H. The structures of integrins and integrin-ligand complexes: Implications for drug design and signal transduction. Angew. Chem. Int. Ed. 2002, 41, 3767–3774. [Google Scholar] [CrossRef]
  33. Gottschalk, K.-E.; Adams, P.D.; Brunger, A.T.; Kessler, H. Transmembrane signal transduction of the αIIbβ3 integrin. Protein Sci. 2002, 11, 1800–1812. [Google Scholar] [CrossRef] [PubMed]
  34. Moser, M.; Legate, K.R.; Zent, R.; Faessler, R. The tail of integrins, talin, and kindlins. Science 2009, 324, 895–899. [Google Scholar] [CrossRef] [PubMed]
  35. Mas-Moruno, C.; Fraioli, R.; Rechenmacher, F.; Neubauer, S.; Kapp, T.G.; Kessler, H. αvβ3- or α5β1-integrin-selective peptidomimetics for surface coating. Angew. Chem. Int. Ed. 2016, 55, 7048–7067. [Google Scholar] [CrossRef] [PubMed]
  36. Askari, J.A.; Buckley, P.A.; Mould, A.P.; Humphries, M.J. Linking integrin conformation to function. J. Cell Sci. 2009, 122, 165–170. [Google Scholar] [CrossRef] [PubMed]
  37. Hoefling, M.; Kessler, H.; Gottschalk, K.-E. The transmembrane structure of integrin αIIbβ3—Significance to signal transduction. Angew. Chem. Int. Ed. 2009, 48, 6590–6593. [Google Scholar] [CrossRef] [PubMed]
  38. Gottschalk, K.-E.; Kessler, H. A computational model of transmembrane integrin clustering. Structure 2004, 12, 1109–1116. [Google Scholar] [CrossRef] [PubMed]
  39. Müller, M.A.; Brunie, L.; Bächer, A.S.; Kessler, H.; Gottschalk, K.-E.; Reuning, U. Cytoplasmic salt bridge formation in integrin αvβ3 stabilizes its inactive state affecting integrin-mediated cell biological effects. Cell Signal. 2014, 11, 2493–2503. [Google Scholar] [CrossRef] [PubMed]
  40. Schiller, H.B.; Hermann, M.R.; Polleux, J.; Vignaud, T.; Zanivan, S.; Friedel, C.C.; Sun, Z.Q.; Raducanu, A.; Gottschalk, K.E.; Thery, M.; et al. β1- and αv-class integrins cooperate to regulate myosin II during rigidity sensing of fibronectin-based microenvironments. Nat. Cell Biol. 2013, 15, 625–636. [Google Scholar] [CrossRef] [PubMed]
  41. Stucci, S.; Tucci, M.; Passarelli, A.; Silvestris, F. αvβ3 integrin: Pathogenetic role in osteotropic tumors. Crit. Rev. Oncol. Hematol. 2015, 96, 183–193. [Google Scholar] [CrossRef] [PubMed]
  42. Janes, S.M.; Watt, F.M. Switch from αvβ1 to αvβ6 integrin expression protects squamous cell carcinomas from anoikis. J. Cell Biol. 2004, 166, 419–431. [Google Scholar] [CrossRef] [PubMed]
  43. Takayama, S.; Ishii, S.; Ikeda, T.; Masamura, S.; Doi, M.; Kitajima, M. The relationship between bone metastasis from human breast cancer and integrin αvβ3 expression. Anticancer Res. 2005, 25, 79–83. [Google Scholar] [PubMed]
  44. Furger, K.A.; Allan, A.L.; Wilson, S.M.; Hota, C.; Vantyghem, S.A.; Postenka, C.O.; Al-Katib, W.; Chambers, A.F.; Tuck, A.B. β3 integrin expression increases breast carcinoma cell responsiveness to the malignancy-enhancing effects of osteopontin. Mol. Cancer Res. 2003, 1, 810–819. [Google Scholar] [PubMed]
  45. Ju, J.A.; Godet, I.; Ye, I.C.; Byun, J.; Jayatilaka, H.; Lee, S.J.; Xiang, L.; Samanta, D.; Lee, M.H.; Wu, P.H.; et al. Hypoxia selectively enhances integrin α5β1 receptor expression in breast cancer to promote metastasis. Mol. Cancer Res. 2017, 15, 723–734. [Google Scholar] [CrossRef] [PubMed]
  46. Bates, R.C.; Bellovin, D.I.; Brown, C.; Maynard, E.; Wu, B.; Kawakatsu, H.; Sheppard, D.; Oettgen, P.; Mercurio, A.M. Transcriptional activation of integrin beta6 during the epithelial-mesenchymal transition defines a novel prognostic indicator of aggressive colon carcinoma. J. Clin. Investig. 2005, 115, 339–347. [Google Scholar] [CrossRef] [PubMed]
  47. Bandyopadhyay, A.; Raghavan, S. Defining the role of integrin αvβ6 in cancer. Curr. Drug Targets 2009, 10, 645–652. [Google Scholar] [CrossRef] [PubMed]
  48. Kawashima, A.; Tsugawa, S.; Boku, A.; Kobayashi, M.; Minamoto, T.; Nakanishi, I.; Oda, Y. Expression of αv integrin family in gastric carcinomas: Increased αvβ6 is associated with lymph node metastasis. Pathol. Res. Pract. 2003, 199, 57–64. [Google Scholar] [CrossRef] [PubMed]
  49. Ahmed, N.; Pansino, F.; Clyde, R.; Murthi, P.; Quinn, M.A.; Rice, G.E.; Agrez, M.V.; Mok, S.; Baker, M.S. Overexpression of alpha(v)beta6 integrin in serous epithelial ovarian cancer regulates extracellular matrix degradation via the plasminogen activation cascade. Carcinogenesis 2002, 23, 237–244. [Google Scholar] [CrossRef] [PubMed]
  50. Ramos, D.M.; But, M.; Regezi, J.; Schmidt, B.L.; Atakilit, A.; Dang, D.; Ellis, D.; Jordan, R.; Li, X. Expression of integrin beta 6 enhances invasive behavior in oral squamous cell carcinoma. Matrix Biol. J. Int. Soc. Matrix Biol. 2002, 21, 297–307. [Google Scholar] [CrossRef]
  51. Sipos, B.; Hahn, D.; Carceller, A.; Piulats, J.; Hedderich, J.; Kalthoff, H.; Goodman, S.L.; Kosmahl, M.; Kloppel, G. Immunohistochemical screening for beta6-integrin subunit expression in adenocarcinomas using a novel monoclonal antibody reveals strong up-regulation in pancreatic ductal adenocarcinomas in vivo and in vitro. Histopathology 2004, 45, 226–236. [Google Scholar] [CrossRef] [PubMed]
  52. Patsenker, E.; Wilkens, L.; Banz, V.; Osterreicher, C.H.; Weimann, R.; Eisele, S.; Keogh, A.; Stroka, D.; Zimmermann, A.; Stickel, F. The αvβ6 integrin is a highly specific immunohistochemical marker for cholangiocarcinoma. J. Hepatol. 2010, 52, 362–369. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, Z.Y.; Xu, K.S.; Wang, J.S.; Yang, G.Y.; Wang, W.; Wang, J.Y.; Niu, W.B.; Liu, E.Y.; Mi, Y.T.; Niu, J. Integrin αvβ6 acts as a prognostic indicator in gastric carcinoma. Clin. Oncol. R. Coll. Radiol. 2008, 20, 61–66. [Google Scholar] [CrossRef] [PubMed]
  54. Arihiro, K.; Kaneko, M.; Fujii, S.; Inai, K.; Yokosaki, Y. Significance of α9β1 and αvβ6 integrin expression in breast carcinoma. Breast Cancer 2000, 7, 19–26. [Google Scholar] [CrossRef] [PubMed]
  55. Ahmed, N.; Riley, C.; Rice, G.E.; Quinn, M.A.; Baker, M.S. αvβ6 integrin—A marker for the malignant potential of epithelial ovarian cancer. J. Histochem. Cytochem. 2002, 50, 1371–1380. [Google Scholar] [CrossRef] [PubMed]
  56. Bates, R.C. Colorectal cancer progression: Integrin αvβ6 and the epithelial-mesenchymal transition (EMT). Cell Cycle 2005, 4, 1350–1352. [Google Scholar] [CrossRef] [PubMed]
  57. Bates, R.C.; Mercurio, A.M. The epithelial-mesenchymal transition (EMT) and colorectal cancer progression. Cancer Biol. Ther. 2005, 4, 365–370. [Google Scholar] [CrossRef] [PubMed]
  58. Ramos, D.M.; Dang, D.; Sadler, S. The role of the integrin αvβ6 in regulating the epithelial to mesenchymal transition in oral cancer. Anticancer Res. 2009, 29, 125–130. [Google Scholar] [PubMed]
  59. Munshi, H.G.; Stack, M.S. Reciprocal interactions between adhesion receptor signaling and MMP regulation. Cancer Metastasis Rev. 2006, 25, 45–56. [Google Scholar] [CrossRef] [PubMed]
  60. Böger, C.; Warneke, V.S.; Behrens, H.M.; Kalthoff, H.; Goodman, S.L.; Becker, T.; Röcken, C. Integrins αvβ3 and αvβ5 as prognostic, diagnostic, and therapeutic targets in gastric cancer. Gastric Cancer 2015, 18, 784–795. [Google Scholar] [CrossRef] [PubMed]
  61. Schnell, O.; Krebs, B.; Wagner, E.; Romagna, A.; Beer, A.J.; Grau, S.J.; Thon, N.; Goetz, C.; Kretzschmar, H.A.; Tonn, J.C.; et al. Expression of integrin αvβ3 in gliomas correlates with tumor grade and is not restricted to tumor vasculature. Brain Pathol. 2008, 18, 378–386. [Google Scholar] [CrossRef] [PubMed]
  62. Berghoff, A.S.; Kovanda, A.K.; Melchardt, T.; Bartsch, R.; Hainfellner, J.A.; Sipos, B.; Schittenhelm, J.; Zielinski, C.C.; Widhalm, G.; Dieckmann, K.; et al. αvβ3, αvβ5 and αvβ6 integrins in brain metastases of lung cancer. Clin. Exp. Metastasis 2014, 31, 841–851. [Google Scholar] [CrossRef] [PubMed]
  63. Böger, C.; Kalthoff, H.; Goodman, S.L.; Behrens, H.M.; Röcken, C. Integrins and their ligands are expressed in non-small cell lung cancer but not correlated with parameters of disease progression. Virchows Arch. 2014, 464, 69–78. [Google Scholar] [CrossRef] [PubMed]
  64. Fabricius, E.M.; Wildner, G.P.; Kruse-Boitschenko, U.; Hoffmeister, B.; Goodman, S.L.; Raguse, J.D. Immunohistochemical analysis of integrins αvβ3, αvβ5 and α5β1, and their ligands, fibrinogen, fibronectin, osteopontin and vitronectin, in frozen sections of human oral head and neck squamous cell carcinomas. Exp. Ther. Med. 2011, 2, 9–19. [Google Scholar] [CrossRef] [PubMed]
  65. Hosotani, R.; Kawaguchi, M.; Masui, T.; Koshiba, T.; Ida, J.; Fujimoto, K.; Wada, M.; Doi, R.; Imamura, M. Expression of integrin αvβ3 in pancreatic carcinoma: Relation to MMP-2 activation and lymph node metastasis. Pancreas 2002, 25, e30–e35. [Google Scholar] [CrossRef] [PubMed]
  66. Heß, K.; Böger, C.; Behrens, H.M.; Röcken, C. Correlation between the expression of integrins in prostate cancer and clinical outcome in 1284 patients. Ann. Diagn. Pathol. 2014, 18, 343–350. [Google Scholar] [CrossRef] [PubMed]
  67. Marsh, D.; Dickinson, S.; Neill, G.W.; Marshall, J.F.; Hart, I.R.; Thomas, G.J. αvβ6 integrin promotes the invasion of morphoeic basal cell carcinoma through stromal modulation. Cancer Res. 2008, 68, 3295–3303. [Google Scholar] [CrossRef] [PubMed]
  68. Moore, K.M.; Thomas, G.J.; Duffy, S.W.; Warwick, J.; Gabe, R.; Chou, P.; Ellis, I.O.; Green, A.R.; Haider, S.; Brouilette, K.; et al. Therapeutic targeting of integrin αvβ6 in breast cancer. J. Natl. Cancer Inst. 2014, 106. [Google Scholar] [CrossRef] [PubMed]
  69. Yang, G.Y.; Guo, S.; Dong, C.Y.; Wang, X.Q.; Hu, B.Y.; Liu, Y.F.; Chen, Y.W.; Niu, J.; Dong, J.H. Integrin αvβ6 sustains and promotes tumor invasive growth in colon cancer progression. World J. Gastroenterol. 2015, 21, 7457–7467. [Google Scholar] [CrossRef] [PubMed]
  70. Hecht, J.L.; Dolinski, B.M.; Gardner, H.A.; Violette, S.M.; Weinreb, P.H. Overexpression of the αvβ6 integrin in endometrial cancer. Appl. Immunohistochem. Mol. Morphol. 2008, 16, 543–547. [Google Scholar] [CrossRef] [PubMed]
  71. Elayadi, A.N.; Samli, K.N.; Prudkin, L.; Liu, Y.H.; Bian, A.; Xie, X.J.; Wistuba, I.I.; Roth, J.A.; McGuire, M.J.; Brown, K.C. A peptide selected by biopanning identifies the integrin αvβ6 as a prognostic biomarker for nonsmall cell lung cancer. Cancer Res. 2007, 67, 5889–5895. [Google Scholar] [CrossRef] [PubMed]
  72. Impola, U.; Uitto, V.J.; Hietanen, J.; Hakkinen, L.; Zhang, L.; Larjava, H.; Isaka, K.; Saarialho-Kere, U. Differential expression of matrilysin-1 (MMP-7), 92 kd gelatinase (MMP-9), and metalloelastase (MMP-12) in oral verrucous and squamous cell cancer. J. Pathol. 2004, 202, 14–22. [Google Scholar] [CrossRef] [PubMed]
  73. Steiger, K.; Schlitter, A.M.; Weichert, W.; Esposito, I.; Wester, H.J.; Notni, J. Perspective of αvβ6-integrin imaging for clinical management of pancreatic carcinoma and its precursor lesions. Mol. Imaging 2017, in press. [Google Scholar] [CrossRef] [PubMed]
  74. Sawada, K.; Mitra, A.K.; Radjabi, A.R.; Bhaskar, V.; Kistner, E.O.; Tretiakova, M.; Jagadeeswaran, S.; Montag, A.; Becker, A.; Kenny, H.A.; et al. Loss of E-cadherin promotes ovarian cancer metastasis via α5-integrin, which is a therapeutic target. Cancer Res. 2008, 68, 2329–2339. [Google Scholar] [CrossRef] [PubMed]
  75. Daley, W.P.; Peters, S.B.; Larsen, M. Extracellular matrix dynamics in development and regenerative medicine. J. Cell Sci. 2008, 121, 255–264. [Google Scholar] [CrossRef] [PubMed]
  76. Jordan, C.T.; Guzman, M.L.; Noble, M. Cancer stem cells. N. Engl. J. Med. 2006, 355, 1253–1261. [Google Scholar] [CrossRef] [PubMed]
  77. Taipale, J.; Beachy, P.A. The Hedgehog and Wnt signalling pathways in cancer. Nature 2001, 411, 349–354. [Google Scholar] [CrossRef] [PubMed]
  78. Hurt, E.M.; Chan, K.; Serrat, M.A.; Thomas, S.B.; Veenstra, T.D.; Farrar, W.L. Identification of vitronectin as an extrinsic inducer of cancer stem cell differentiation and tumor formation. Stem Cells 2010, 28, 390–398. [Google Scholar] [CrossRef] [PubMed]
  79. Seguin, L.; Kato, S.; Franovic, A.; Camargo, M.F.; Lesperance, J.; Elliott, K.C.; Yebra, M.; Mielgo, A.; Lowy, A.M.; Husain, H.; et al. An integrin β3-KRAS-RalB complex drives tumour stemness and resistance to EGFR inhibition. Nat. Cell Biol. 2014, 16, 457–468. [Google Scholar] [CrossRef] [PubMed]
  80. Vellon, L.; Menendez, J.A.; Liu, H.; Lupu, R. Upregulation of αvβ3 integrin expression is a novel molecular response to chemotherapy-induced cell damage in a heregulin-dependent manner. Differentiation 2007, 75, 819–830. [Google Scholar] [CrossRef] [PubMed]
  81. Sheldrake, H.M.; Patterson, L.H. Function and antagonism of β3 integrins in the development of cancer therapy. Curr. Cancer Drug Targets 2009, 9, 519–540. [Google Scholar] [CrossRef] [PubMed]
  82. Gianni, T.; Massaro, R.; Campadelli-Fiume, G. Dissociation of HSV gL from gH by αvβ6- or αvβ8-integrin promotes gH activation and virus entry. Proc. Natl. Acad. Sci. USA 2015, 112, E3901–E3910. [Google Scholar] [CrossRef] [PubMed]
  83. Chesnokova, L.S.; Nishimura, S.L.; Hutt-Fletcher, L.M. Fusion of epithelial cells by Epstein-Barr Virus proteins is triggered by binding of viral glycoproteins gHgL to integrins αvβ6 or αvβ8. Proc. Natl. Acad. Sci. USA 2009, 106, 20464–20469. [Google Scholar] [CrossRef] [PubMed]
  84. Hutt-Fletcher, L.M.; Chesnokova, L.S. Integrins as triggers of Epstein-barr virus fusion and epithelial cell infection. Virulence 2010, 1, 395–398. [Google Scholar] [CrossRef] [PubMed]
  85. Bear, J.E.; Haugh, J.M. Directed migration of mesenchymal cells: Where signaling and the cytoskeleton meet. Curr. Opin. Cell Biol. 2014, 30, 74–82. [Google Scholar] [CrossRef] [PubMed]
  86. Parsons, J.T.; Horwitz, A.R.; Schwartz, M.A. Cell adhesion: Integrating cytoskeletal dynamics and cellular tension. Nat. Rev. Mol. Cell Biol. 2010, 11, 633–643. [Google Scholar] [CrossRef] [PubMed]
  87. Missirlis, D.; Haraszti, T.; Scheele, C.; Wiegand, T.; Diaz, C.; Neubauer, S.; Rechenmacher, F.; Kessler, H.; Spatz, J.P. Substrate engagement of integrins α5β1 and αvβ3 is necessary, but not sufficient, for high directional persistence in migration on fibronectin. Sci. Rep. 2016, 6, 23258. [Google Scholar] [CrossRef] [PubMed]
  88. Ganguly, K.K.; Pal, S.; Moulik, S.; Chatterjee, A. Integrins and metastasis. Cell Adhes. Migr. 2013, 7, 251–261. [Google Scholar] [CrossRef] [PubMed]
  89. Lawson, C.D.; Burridge, K. The on-off relationship of Rho and Rac during integrin-mediated adhesion and cell migration. Small GTPases 2014, 5, e27958. [Google Scholar] [CrossRef] [PubMed]
  90. Cao, Z.; Livas, T.; Kyprianou, N. Anoikis and EMT: Lethal “liaisons” during cancer progression. Crit. Rev. Oncol. 2016, 21, 155–168. [Google Scholar] [CrossRef] [PubMed]
  91. Kalluri, R.; Weinberg, R.A. The basics of epithelial-mesenchymal transition. J. Clin. Investig. 2009, 119, 1420–1428. [Google Scholar] [CrossRef] [PubMed]
  92. Bartsch, J.E.; Staren, E.D.; Appert, H.E. Adhesion and migration of extracellular matrix-stimulated breast cancer. J. Surg. Res. 2003, 110, 287–294. [Google Scholar] [CrossRef]
  93. Pytela, R.; Pierschbacher, M.D.; Ruoslahti, E. Identification and isolation of a 140 kd cell surface glycoprotein with properties expected of a fibronectin receptor. Cell 1985, 40, 191–198. [Google Scholar] [CrossRef]
  94. Stefansson, S.; Lawrence, D.A. The serpin PAI-1 inhibits cell migration by blocking integrin αvβ3 binding to vitronectin. Nature 1996, 383, 441–443. [Google Scholar] [CrossRef] [PubMed]
  95. Lian, J.; Guoping, C.; Shapiro, S.S.; Tran, L.P.; Beacham, D.A. Glycoprotein Ibα can mediate endothelial cell migration on von Willebrand factor-containing substrata. Exp. Cell Res. 1999, 252, 114–122. [Google Scholar] [CrossRef] [PubMed]
  96. Hamidi, H.; Pietilä, M.; Ivaska, J. The complexity of integrins in cancer and new scopes for therapeutic targeting. Br. J. Cancer 2016, 115, 1017–1023. [Google Scholar] [CrossRef] [PubMed]
  97. Gehler, S.; Ponik, S.M.; Riching, K.M.; Keely, P.J. Bi-directional signaling: Extracellular matrix and integrin regulation of breast tumor progression. Crit. Rev. Eukaryot. Gene Expr. 2013, 23, 139–157. [Google Scholar] [CrossRef] [PubMed]
  98. Upheber, S.; Karle, A.; Miller, J.; Schlaugk, S.; Gross, E.; Reuning, U. Alternative splicing of KAI1 abrogates its tumor-suppressive effects on integrin αvβ3-mediated ovarian cancer biology. Cell Signal. 2015, 27, 652–662. [Google Scholar] [CrossRef] [PubMed]
  99. Hapke, S.; Kessler, H.; Luber, B.; Benge, A.; Hutzler, P.; Höfler, H.; Schmitt, M.; Reuning, U. Ovarian cancer cell proliferation and motility is induced by engagement of integrin αvβ3/vitronectin interaction. Biol. Chem. 2003, 384, 1073–1083. [Google Scholar] [CrossRef] [PubMed]
  100. Löffek, S.; Franzke, C.W.; Helfrich, I. Tension in Cancer. Int. J. Mol. Sci. 2016, 17, 1910. [Google Scholar] [CrossRef] [PubMed]
  101. Morgan, M.R.; Byron, A.; Humphries, M.J.; Bass, M.D. Giving off mixed signals-distinct functions of α5β1 and αvβ3 integrins in regulating cell behaviour. IUBMB Life 2009, 61, 731–738. [Google Scholar] [CrossRef] [PubMed]
  102. Roca-Cusachs, P.; Iskratsch, T.; Sheetz, M.P. Finding the weakest link: Exploring integrin-mediated mechanical molecular pathways. J. Cell Sci. 2012, 125, 3025–3038. [Google Scholar] [CrossRef] [PubMed]
  103. McCabe, N.P.; De, S.; Vasanji, A.; Brainard, J.; Byzova, T.V. Prostate cancer specific integrin αvβ3 modulates bone metastatic growth and tissue remodeling. Oncogene 2007, 26, 6238–6243. [Google Scholar] [CrossRef] [PubMed]
  104. Mierke, C.T.; Frey, B.; Fellner, M.; Herrmann, M.; Fabry, B. Integrin α5β1 facilitates cancer cell invasion through enhanced contractile forces. J. Cell Sci. 2011, 124, 369–383. [Google Scholar] [CrossRef] [PubMed]
  105. Paul, N.R.; Allen, J.L.; Chapman, A.; Morlan-Mairal, M.; Zindy, E.; Jacquemet, G.; Fernandez del Ama, L.; Ferizovic, N.; Green, D.M.; Howe, J.D.; et al. α5β1 integrin recycling promotes Arp2/3-independent cancer cell invasion via the formin FHOD3. J. Cell Biol. 2015, 210, 1013–1031. [Google Scholar] [CrossRef] [PubMed]
  106. Breuss, J.M.; Gillett, N.; Lu, L.; Sheppard, D.; Pytela, R. Restricted distribution of integrin beta 6 mRNA in primate epithelial tissues. J. Histochem. Cytochem. 1993, 41, 1521–1527. [Google Scholar] [CrossRef] [PubMed]
  107. Niu, G.; Chen, X. Why integrin as a primary target for imaging and therapy. Theranostics 2011, 1, 30–47. [Google Scholar] [CrossRef] [PubMed]
  108. Breuss, J.M.; Gallo, J.; DeLisser, H.M.; Klimanskaya, I.V.; Folkesson, H.G.; Pittet, J.F.; Nishimura, S.L.; Aldape, K.; Landers, D.V.; Carpenter, W. Expression of the beta 6 integrin subunit in development, neoplasia and tissue repair suggests a role in epithelial remodeling. J. Cell Sci. 1995, 108, 2241–2251. [Google Scholar] [PubMed]
  109. Huang, R.; Zong, X. Aberrant cancer metabolism in epithelial-mesenchymal transition and cancer metastasis: Mechanisms in cancer progression. Crit. Rev. Oncol. Hematol. 2017, 115, 13–22. [Google Scholar] [CrossRef] [PubMed]
  110. Jiang, X.; Teng, M.; Guo, X.; Zhang, D.; Zhang, Q.; Zhang, J.; Huang, Y. Switch from αvβ5 to αvβ6 integrin is required for CD9-regulated keratinocyte migration and MMP-9 activation. FEBS Lett. 2014, 588, 4044–4052. [Google Scholar] [CrossRef] [PubMed]
  111. Guo, W.; Giancotti, F.G. Integrin signalling during tumour progression. Nat. Rev. Mol. Cell Biol. 2004, 5, 816–826. [Google Scholar] [CrossRef] [PubMed]
  112. Balzac, F.; Avolio, M.; Degani, S.; Kaverina, I.; Torti, M.; Silengo, L.; Small, J.V.; Retta, S.F. E-cadherin endocytosis regulates the activity of Rap1: A traffic light GTPase at the crossroads between cadherin and integrin function. J. Cell Sci. 2005, 118, 4765–4783. [Google Scholar] [CrossRef] [PubMed]
  113. Oloumi, A.; McPhee, T.; Dedhar, S. Regulation of E-cadherin expression and beta-catenin/Tcf transcriptional activity by the integrin-linked kinase. Biochim. Biophys. Acta 2004, 1691, 1–15. [Google Scholar] [CrossRef] [PubMed]
  114. Yang, G.Y.; Xu, K.S.; Pan, Z.Q.; Zhang, Z.Y.; Mi, Y.T.; Wang, J.S.; Chen, R.; Niu, J. Integrin αvβ6 mediates the potential for colon cancer cells to colonize in and metastasize to the liver. Cancer Sci. 2008, 99, 879–887. [Google Scholar] [CrossRef] [PubMed]
  115. Saldanha, R.G.; Molloy, M.P.; Bdeir, K.; Cines, D.B.; Song, X.; Uitto, P.M.; Weinreb, P.H.; Violette, S.M.; Baker, M.S. Proteomic identification of lynchpin urokinase plasminogen activator receptor protein interactions associated with epithelial cancer malignancy. J. Proteome Res. 2007, 6, 1016–1028. [Google Scholar] [CrossRef] [PubMed]
  116. Li, X.; Yang, Y.; Hu, Y.; Dang, D.; Regezi, J.; Schmidt, B.L.; Atakilit, A.; Chen, B.; Ellis, D.; Ramos, D.M. Alphavbeta6-fyn signaling promotes oral cancer progression. J. Biol. Chem. 2003, 278, 41646–41653. [Google Scholar] [CrossRef] [PubMed]
  117. Al-Hazmi, N.; Thomas, G.J.; Speight, P.M.; Whawell, S.A. The 120 kDa cell-binding fragment of fibronectin up-regulates migration of αvβ6-expressing cells by increasing matrix metalloproteinase-2 and -9 secretion. Eur. J. Oral Sci. 2007, 115, 454–458. [Google Scholar] [CrossRef] [PubMed]
  118. Morgan, M.R.; Thomas, G.J.; Russell, A.; Hart, I.R.; Marshall, J.F. The integrin cytoplasmic-tail motif ekqkvdlstdc is sufficient to promote tumor cell invasion mediated by matrix metalloproteinase (MMP)-2 or MMP-9. J. Biol. Chem. 2004, 279, 26533–26539. [Google Scholar] [CrossRef] [PubMed]
  119. Thomas, G.J.; Hart, I.R.; Speight, P.M.; Marshall, J.F. Binding of TGF-β1 latency-associated peptide (LAP) to αvβ6 integrin modulates behaviour of squamous carcinoma cells. Br. J. Cancer 2002, 87, 859–867. [Google Scholar] [CrossRef] [PubMed]
  120. Thomas, G.J.; Lewis, M.P.; Hart, I.R.; Marshall, J.F.; Speight, P.M. αvβ6 integrin promotes invasion of squamous carcinoma cells through up-regulation of matrix metalloproteinase-9. Int. J. Cancer 2001, 92, 641–650. [Google Scholar] [CrossRef]
  121. Niu, J.; Dorahy, D.J.; Gu, X.; Scott, R.J.; Draganic, B.; Ahmed, N.; Agrez, M.V. Integrin expression in colon cancer cells is regulated by the cytoplasmic domain of the β6 integrin subunit. Int. J. Cancer 2002, 99, 529–537. [Google Scholar] [CrossRef] [PubMed]
  122. Ramsay, A.G.; Keppler, M.D.; Jazayeri, M.; Thomas, G.J.; Parsons, M.; Violette, S.; Weinreb, P.; Hart, I.R.; Marshall, J.F. HS1-associated protein X-1 regulates carcinoma cell migration and invasion via clathrin-mediated endocytosis of integrin αvβ6. Cancer Res. 2007, 67, 5275–5284. [Google Scholar] [CrossRef] [PubMed]
  123. Lanzetti, L.; Di Fiore, P.P. Endocytosis and Cancer: An ‘Insider’ Network with Dangerous Liaisons. Traffic 2008, 9, 2011–2021. [Google Scholar] [CrossRef] [PubMed]
  124. Nishimura, S.L.; Sheppard, D.; Pytela, R. Integrin αvβ8. Interaction with vitronectin and functional divergence of the β8 cytoplasmic domain. J. Biol. Chem. 1994, 269, 28708–28715. [Google Scholar] [PubMed]
  125. Worthington, J.J.; Klementowicz, J.E.; Travis, M.A. TGF-β: A sleeping giant awoken by integrins. Trends Biochem. Sci. 2011, 36, 47–54. [Google Scholar] [CrossRef] [PubMed]
  126. Aluwihare, P.; Mu, Z.; Zhao, Z.; Yu, D.; Weinreb, P.H.; Horan, G.S.; Violette, S.M.; Munger, J.S. Mice that lack activity of αvβ6- and αvβ8-integrins reproduce the abnormalities of TGF-β1- and TGF-β3-null mice. J. Cell Sci. 2009, 122, 227–232. [Google Scholar] [CrossRef] [PubMed]
  127. Ozawa, A.; Sato, Y.; Imabayashi, T.; Uemura, T.; Takagi, J.; Sekiguchi, K. Molecular basis of the ligand binding specificity of αvβ8 integrin. J. Biol. Chem. 2016, 291, 11551–11565. [Google Scholar] [CrossRef] [PubMed]
  128. Mu, D.; Cambier, S.; Fjellbirkeland, L.; Baron, J.L.; Munger, J.S.; Kawakatsu, H.; Sheppard, D.; Broaddus, V.C.; Nishimura, S.L. The integrin αvβ8 mediates epithelial homeostasis through MT1-MMP-dependent activation of TGF-β1. J. Cell Biol. 2002, 157, 493–507. [Google Scholar] [CrossRef] [PubMed]
  129. McCarty, J.H.; Cook, A.A.; Hynes, R.O. An interaction between αvβ8 integrin and band 4.1b via a highly conserved region of the band 4.1 c-terminal domain. Proc. Natl. Acad. Sci. USA 2005, 102, 13479–13483. [Google Scholar] [CrossRef] [PubMed]
  130. Cheerathodi, M.; Avci, N.G.; Guerrero, P.A.; Tang, L.K.; Popp, J.; Morales, J.E.; Chen, Z.; Carnero, A.; Lang, F.F.; Ballif, B.A.; et al. The cytoskeletal adapter protein spinophilin regulates invadopodia dynamics and tumor cell invasion in glioblastoma. Mol. Cancer Res. 2016, 14, 1277–1287. [Google Scholar] [CrossRef] [PubMed]
  131. Neurohr, C.; Nishimura, S.L.; Sheppard, D. Activation of transforming growth factor-beta by the integrin αvβ8 delays epithelial wound closure. Am. J. Respir. Cell Mol. Biol. 2006, 35, 252–259. [Google Scholar] [CrossRef] [PubMed]
  132. Milner, R.; Huang, X.; Wu, J.; Nishimura, S.; Pytela, R.; Sheppard, D.; Ffrench-Constant, C. Distinct roles for astrocyte a αvβ5 and αvβ8 integrins in adhesion and migration. J. Cell Sci. 1999, 112, 4271–4279. [Google Scholar] [PubMed]
  133. Cambier, S.; Gline, S.; Mu, D.; Collins, R.; Araya, J.; Dolganov, G.; Einheber, S.; Boudreau, N.; Nishimura, S.L. Integrin αvβ8-mediated activation of transforming growth factor-β by perivascular astrocytes: An angiogenic control switch. Am. J. Pathol. 2005, 166, 1883–1894. [Google Scholar] [CrossRef]
  134. Hanahan, D.; Weinberg, R.A. The hallmarks of cancer. Cell 2000, 100, 57–70. [Google Scholar] [CrossRef]
  135. Cruet-Hennequart, S.; Maubant, S.; Luis, J.; Gauduchon, P.; Staedel, C.; Dedhar, S. αv integrins regulate cell proliferation through integrin-linked kinase (ILK) in ovarian cancer cells. Oncogene 2003, 22, 1688–1702. [Google Scholar] [CrossRef] [PubMed]
  136. Wei, Q.; Pohl, T.L.; Seckinger, A.; Spatz, J.P.; Cavalcanti-Adam, E.A. Regulation of integrin and growth factor signaling in biomaterials for osteodifferentiation. Beilstein J. Org. Chem. 2015, 11, 773–783. [Google Scholar] [CrossRef] [PubMed]
  137. Streuli, C.H.; Akhtar, N. Signal co-operation between integrins and other receptor systems. Biochem. J. 2009, 418, 491–506. [Google Scholar] [CrossRef] [PubMed]
  138. Hannigan, G.; Troussard, A.A.; Dedhar, S. Integrin-linked kinase. A cancer therapeutic target unique among its ILK. Nat. Rev. Cancer 2005, 5, 51–63. [Google Scholar] [CrossRef] [PubMed]
  139. Stupack, D.G.; Cheresh, D.A. Get a ligand, get a life: Integrins, signaling and cell survival. J. Cell Sci. 2002, 115, 3729–3738. [Google Scholar] [CrossRef] [PubMed]
  140. Lössner, D.; Abou-Ajram, C.; Benge, A.; Reuning, U. Integrin αvβ3 mediates upregulation of epidermal growth-factor receptor expression and activity in human ovarian cancer cells. Int. J. Biochem. Cell. Biol. 2008, 40, 2746–2761. [Google Scholar] [CrossRef] [PubMed]
  141. Schwartz, M.A.; Assoian, R.K. Integrins and cell proliferation: Regulation of cyclin-dependent kinases via cytoplasmic signaling pathways. J. Cell Sci. 2001, 114, 2553–2560. [Google Scholar] [PubMed]
  142. Munger, J.S.; Huang, X.; Kawakatsu, H.; Griffiths, M.J.; Dalton, S.L.; Wu, J.; Pittet, J.F.; Kaminski, N.; Garat, C.; Matthay, M.A.; et al. The integrin αvβ6 binds and activates latent TGF-β1: A mechanism for regulating pulmonary inflammation and fibrosis. Cell 1999, 96, 319–328. [Google Scholar] [CrossRef]
  143. Liang, D.; Xu, W.; Zhang, Q.; Tao, B.B. Study on the effect of Integrin αvβ6 on proliferation and apoptosis of cervical cancer cells. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 2811–2815. [Google Scholar] [PubMed]
  144. Fjellbirkeland, L.; Cambier, S.; Broaddus, V.C.; Hill, A.; Brunetta, P.; Dolganov, G.; Jablons, D.; Nishimura, S.L. Integrin αvβ8-mediated activation of transforming growth factor-beta inhibits human airway epithelial proliferation in intact bronchial tissue. Am. J. Pathol. 2003, 163, 533–542. [Google Scholar] [CrossRef]
  145. Cambier, S.; Mu, D.Z.; O’Connell, D.; Boylen, K.; Travis, W.; Liu, W.H.; Broaddus, V.C.; Nishimura, S.L. A role for the integrin αvβ8 in the negative regulation of epithelial cell growth. Cancer Res. 2000, 60, 7084–7093. [Google Scholar] [PubMed]
  146. Giacotti, F.G.; Ruoslahti, E. Transduction-integrin signaling. Science 1999, 285, 1028–1032. [Google Scholar] [CrossRef]
  147. Hayashido, Y.; Kitano, H.; Sakaue, T.; Fujii, T.; Suematsu, M.; Sakurai, S.; Okamoto, T. Overexpression of integrin αv facilitates proliferation and invasion of oral squamous cell carcinoma cells via MEK/ERK signaling pathway that is activated by interaction of integrin αvβ8 with type collagen. Int. J. Oncol. 2014, 45, 1875–1882. [Google Scholar] [CrossRef] [PubMed]
  148. Hashimoto, M.; Yanagisawa, H.; Minagawa, S.; Sen, D.; Ma, R.; Murray, L.A.; Tsui, P.; Lou, J.; Marks, J.D.; Baron, J.L.; et al. TGF-β-dependent dendritic cell chemokinesis in murine models of airway disease. J. Immunol. 2015, 195, 1182–1190. [Google Scholar] [CrossRef] [PubMed]
  149. Brand, O.J.; Somanath, S.; Moermans, C.; Yanagisawa, H.; Hashimoto, M.; Cambier, S.; Markovics, J.; Bondesson, A.J.; Hill, A.; Jablons, D.; et al. Transforming growth factor-β and interleukin-1β signaling pathways converge on the chemokine CCL20 promoter. J. Biol. Chem. 2015, 290, 14717–14728. [Google Scholar] [CrossRef] [PubMed]
  150. Kitamura, H.; Cambier, S.; Somanath, S.; Barker, T.; Minagawa, S.; Markovics, J.; Goodsell, A.; Publicover, J.; Reichardt, L.; Jablons, D.; et al. Mouse and human lung fibroblasts regulate dendritic cell trafficking, airway inflammation, and fibrosis through integrin αvβ8-mediated activation of TGF-β. J. Clin. Investig. 2011, 121, 2863–2875. [Google Scholar] [CrossRef] [PubMed]
  151. Worthington, J.J.; Kelly, A.; Smedley, C.; Bauche, D.; Campbell, S.; Marie, J.C.; Travis, M.A. Integrin αvβ8-mediated TGF-β activation by effector regulatory T cells is essential for suppression of T-cell-mediated inflammation. Immunity 2015, 42, 903–915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Edwards, J.P.; Thornton, A.M.; Shevach, E.M. Release of active TGF-β1 from the latent TGF-β1/GARP complex on T regulatory cells is mediated by integrin β8. J. Immunol. 2014, 193, 2843–2849. [Google Scholar] [CrossRef] [PubMed]
  153. Cheresh, D.A.; Stupack, D.G. Integrin-mediated death: An explanation of the integrin-knockout phenotype? Nat. Med. 2002, 8, 193–194. [Google Scholar] [CrossRef] [PubMed]
  154. Frisch, S.M.; Ruoslahti, E. Integrins and anoikis. Curr. Opin. Cell Biol. 1997, 9, 701–770. [Google Scholar] [CrossRef]
  155. Buchheit, C.L.; Weigel, K.J.; Schafer, Z.T. Cancer cell survival during detachment from the ECM: Multiple barriers to tumour progression. Nat. Rev. Cancer 2014, 14, 632–641. [Google Scholar] [CrossRef] [PubMed]
  156. Teitz, T.; Wei, T.; Valentine, M.B.; Vanin, E.F.; Grenet, J.; Valentine, V.A.; Behm, F.G.; Look, A.T.; Lahti, J.M.; Kidd, V.J. Caspase 8 is deleted or silenced preferentially in childhood neuroblastomas with amplification of MYCN. Nat. Med. 2000, 6, 529–535. [Google Scholar] [CrossRef] [PubMed]
  157. Alanko, J.; Mai, A.; Jacquemet, G.; Schauer, K.; Kaukonen, R.; Saari, M.; Goud, B.; Ivaska, J. Integrin endosomal signalling suppresses anoikis. Nat. Cell Biol. 2015, 17, 1412–1421. [Google Scholar] [CrossRef] [PubMed]
  158. Plantefaber, L.C.; Hynes, R.O. Changes in integrin receptors on oncogenically transformed cells. Cell 1989, 56, 281–290. [Google Scholar] [CrossRef]
  159. Salmeron-Sanchez, M.; Dalby, M.J. Synergistic growth factor microenvironments. Chem. Commun. Camb. 2016, 52, 13327–13336. [Google Scholar] [CrossRef] [PubMed]
  160. Reginato, M.J.; Mills, K.R.; Paulus, J.K.; Lynch, D.K.; Sgroi, D.C.; Debnath, J.; Muthuswamy, S.K.; Brugge, J.S. Integrins and EGFR coordinately regulate the pro-apoptotic protein BIM to prevent anoikis. Nat. Cell Biol. 2003, 5, 733–740. [Google Scholar] [CrossRef] [PubMed]
  161. Ray, A.-M.; Schaffner, F.; Janouskova, H.; Noulet, F.; Rognan, D.; Lelong-Rebel, I.; Choulier, L.; Blandin, A.-F.; Lehmann, M.; Martin, S.; et al. Single cell tracking assay reveals an opposite effect of selective small non-peptidic α5β1 or αvβ3/β5 integrins antagonists in U87MG glioma cells. Biochem. Biophys. Acta 2014, 1840, 2978–2987. [Google Scholar] [CrossRef] [PubMed]
  162. Renner, G.; Janouskova, H.; Noulet, F.; Guerin, E.; Bär, S.; Nuesch, J.; Rechenmacher, F.; Neubauer, S.; Kessler, H.; Choulier, L.; et al. Integrin α5β1 and p53 convergent pathways in the control of anti-apoptotic proteins PEA-15 and survivin in high grade glioma. Cell Death Differ. 2016, 23, 640–653. [Google Scholar] [CrossRef] [PubMed]
  163. Paoli, P.; Giannoni, E.; Chiarugi, P. Anoikis molecular pathways and its role in cancer progression. Biochim. Biophys. Acta 2013, 1833, 3481–3498. [Google Scholar] [CrossRef] [PubMed]
  164. Demircioglu, F.; Hodivala-Dilke, K. αvβ3 Integrin and tumour blood vessels-learning from the past to shape the future. Curr. Opin. Cell Biol. 2016, 42, 121–127. [Google Scholar] [CrossRef] [PubMed]
  165. Avraamides, C.J.; Garmy-Susini, B.; Varner, J.A. Integrins in angiogenesis and lymphangiogenesis. Nat. Rev. Cancer 2008, 8, 604–617. [Google Scholar] [CrossRef] [PubMed]
  166. Bergers, G.; Benjamin, L.E. Tumorigenesis and the angiogenic switch. Nat. Rev. Cancer 2003, 3, 401–410. [Google Scholar] [CrossRef] [PubMed]
  167. Kim, S.; Bell, K.; Mousa, S.A.; Varner, J.A. Regulation of angiogenesis in vivo by ligation of integrin α5β1 with the central cell-binding domain of fibronectin. Am. J. Pathol. 2000, 156, 1345–1362. [Google Scholar] [CrossRef]
  168. Stefanidakis, M.; Koivunen, E. Cell-surface association between matrix metalloproteinases and integrins: Role of the complexes in leukocyte migration and cancer progression. Blood 2006, 108, 1441–1450. [Google Scholar] [CrossRef] [PubMed]
  169. Mengele, K.; Napieralski, R.; Magdolen, V.; Reuning, U.; Gkazepis, A.; Sweep, F.; Brunner, N.; Foekens, J.; Harbeck, N.; Schmitt, M. Characteristics of the level-of-evidence-1 disease forecast cancer biomarkers uPA and its inhibitor PAI-1. Expert Rev. Mol. Diagn. 2010, 10, 947–962. [Google Scholar] [CrossRef] [PubMed]
  170. Wang, G.L.; Jiang, B.H.; Rue, E.A.; Semenza, G.L. Hypoxia-inducible factor 1 is a basic-helix-loop-helix-pas heterodimer regulated by cellular O2 tension. Proc. Natl. Acad. Sci. USA 1995, 92, 5510–5514. [Google Scholar] [CrossRef] [PubMed]
  171. Krock, B.L.; Skuli, N.; Simon, M.C. Hypoxia-induced angiogenesis: Good and evil. Genes Cancer 2011, 2, 1117–1133. [Google Scholar] [CrossRef] [PubMed]
  172. Cowden Dahl, K.D.; Robertson, S.E.; Weaver, V.M.; Simon, M.C. Hypoxia-inducible factor regulates αvβ3 integrin cell surface expression. Mol. Biol. Cell 2005, 16, 1901–1912. [Google Scholar] [CrossRef] [PubMed]
  173. Keely, S.; Glover, L.E.; MacManus, C.F.; Campbell, E.L.; Scully, M.M.; Furuta, G.T.; Colgan, S.P. Selective induction of integrin β1 by hypoxia-inducible factor: Implications for wound healing. FASEB J. 2009, 23, 1338–1346. [Google Scholar] [CrossRef] [PubMed]
  174. Brooks, D.L.; Schwab, L.P.; Krutilina, R.; Parke, D.N.; Sethuraman, A.; Hoogewijs, D.; Schorg, A.; Gotwald, L.; Fan, M.; Wenger, R.H.; et al. Itga6 is directly regulated by hypoxia-inducible factors and enriches for cancer stem cell activity and invasion in metastatic breast cancer models. Mol. Cancer 2016, 15, 26. [Google Scholar] [CrossRef] [PubMed]
  175. Chou, C.C.; Chuang, H.C.; Salunke, S.B.; Kulp, S.K.; Chen, C.S. A novel HIF-1 alpha-integrin-linked kinase regulatory loop that facilitates hypoxia-induced HIF-1alpha expression and epithelial-mesenchymal transition in cancer cells. Oncotarget 2015, 6, 8271–8285. [Google Scholar] [CrossRef] [PubMed]
  176. Liu, W.; Shen, S.M.; Zhao, X.Y.; Chen, G.Q. Targeted genes and interacting proteins of hypoxia inducible factor-1. Int. J. Biochem. Mol. Biol. 2012, 3, 165–178. [Google Scholar] [PubMed]
  177. Han, Z.B.; Ren, H.; Zhao, H.; Chi, Y.; Chen, K.; Zhou, B.; Liu, Y.J.; Zhang, L.; Xu, B.; Liu, B.; et al. Hypoxia-inducible factor (HIF)-1 alpha directly enhances the transcriptional activity of stem cell factor (scf) in response to hypoxia and epidermal growth factor (EGF). Carcinogenesis 2008, 29, 1853–1861. [Google Scholar] [CrossRef] [PubMed]
  178. Kelly, B.D.; Hackett, S.F.; Hirota, K.; Oshima, Y.; Cai, Z.; Berg-Dixon, S.; Rowan, A.; Yan, Z.; Campochiaro, P.A.; Semenza, G.L. Cell type-specific regulation of angiogenic growth factor gene expression and induction of angiogenesis in nonischemic tissue by a constitutively active form of hypoxia-inducible factor 1. Circ. Res. 2003, 93, 1074–1081. [Google Scholar] [CrossRef] [PubMed]
  179. Nikitenko, L.L.; Smith, D.M.; Bicknell, R.; Rees, M.C. Transcriptional regulation of the crlr gene in human microvascular endothelial cells by hypoxia. FASEB J. 2003, 17, 1499–1501. [Google Scholar] [CrossRef] [PubMed]
  180. Weis, S.M.; Cheresh, D.A. αV integrins in angiogenesis and cancer. Cold Spring Harb. Perspect. Med. 2011, 1, a006478. [Google Scholar] [CrossRef] [PubMed]
  181. Raguse, J.D.; Gath, H.J.; Bier, J.; Riess, H.; Oettle, H. Cilengitide (emd 121974) arrests the growth of a heavily pretreated highly vascularised head and neck tumour. Oral Oncol. 2004, 40, 228–230. [Google Scholar] [CrossRef] [PubMed]
  182. Reynolds, A.R.; Hart, I.R.; Watson, A.R.; Welti, J.C.; Silva, R.G.; Robinson, S.D.; Da Violante, G.; Gourlaourou, V.; Salih, M.; Jones, M.C.; et al. Stimulation of tumor growth and angiogenesis by low concentrations of RGD-mimetic integrin inhibitors. Nat. Med. 2009, 15, 392–400. [Google Scholar] [CrossRef] [PubMed]
  183. Bader, B.L.; Rayburn, H.; Crowley, D.; Hynes, R.O. Extensive vasculogenesis, angiogenesis, and organogenesis precede lethality in mice lacking all αv integrins. Cell 1998, 95, 507–519. [Google Scholar] [CrossRef]
  184. Reynolds, L.E.; Wyder, L.; Lively, J.C.; Taverna, D.; Robinson, S.D.; Huang, X.; Sheppard, D.; Hynes, R.O.; Hodivala-Dilke, K.M. Enhanced pathological angiogenesis in mice lacking β3 integrin or β3 and β5 integrins. Nat. Med. 2002, 8, 27–34. [Google Scholar] [CrossRef] [PubMed]
  185. Van der Flier, A.; Badu-Nkansah, K.; Whittaker, C.A.; Crowley, D.; Bronson, R.T.; Lacy-Hulbert, A.; Hynes, R.O. Endothelial α5 and αv integrins cooperate in remodeling of the vasculature during development. Development 2010, 137, 2439–2449. [Google Scholar] [CrossRef] [PubMed]
  186. Wong, P.P.; Demircioglu, F.; Ghazaly, E.; Alrawashdeh, W.; Stratford, M.R.L.; Scudamore, C.L.; Cereser, B.; Crnogorac-Jurcevic, T.; McDonald, S.; Elia, G.; et al. Dual-action combination therapy enhances angiogenesis while reducing tumor growth and spread. Cancer Cell 2015, 27, 123–137. [Google Scholar] [CrossRef] [PubMed]
  187. Kostourou, V.; Lechertier, T.; Reynolds, L.E.; Lees, D.M.; Baker, M.; Jones, D.T.; Tavora, B.; Ramjaun, A.R.; Birdsey, G.M.; Robinson, S.D.; et al. FAK-heterozygous mice display enhanced tumour angiogenesis. Nat. Commun. 2013, 4, 2020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Wong, P.P.; Bodrug, N.; Hodivala-Dilke, K.M. Exploring Novel Methods for Modulating Tumor Blood Vessels in Cancer Treatment. Curr. Biol. 2016, 26, R1161–R1166. [Google Scholar] [CrossRef] [PubMed]
  189. Arnold, T.D.; Niaudet, C.; Pang, M.F.; Siegenthaler, J.; Gaengel, K.; Jung, B.; Ferrero, G.M.; Mukouyama, Y.S.; Fuxe, J.; Akhurst, R.; et al. Excessive vascular sprouting underlies cerebral hemorrhage in mice lacking αvβ8-TGFβ signaling in the brain. Development 2014, 141, 4489–4499. [Google Scholar] [CrossRef] [PubMed]
  190. Liu, J.; Zeng, L.; Kennedy, R.M.; Gruenig, N.M.; Childs, S.J. βPix plays a dual role in cerebral vascular stability and angiogenesis, and interacts with integrin αvβ8. Dev. Biol. 2012, 363, 95–105. [Google Scholar] [CrossRef] [PubMed]
  191. Hirota, S.; Liu, Q.; Lee, H.S.; Hossain, M.G.; Lacy-Hulbert, A.; McCarty, J.H. The astrocyte-expressed integrin αvβ8 governs blood vessel sprouting in the developing retina. Development 2011, 138, 5157–5166. [Google Scholar] [CrossRef] [PubMed]
  192. Tchaicha, J.H.; Mobley, A.K.; Hossain, M.G.; Aldape, K.D.; McCarty, J.H. A mosaic mouse model of astrocytoma identifies αvβ8 integrin as a negative regulator of tumor angiogenesis. Oncogene 2010, 29, 4460–4472. [Google Scholar] [CrossRef] [PubMed]
  193. Su, H.; Kim, H.; Pawlikowska, L.; Kitamura, H.; Shen, F.; Cambier, S.; Markovics, J.; Lawton, M.T.; Sidney, S.; Bollen, A.W.; et al. Reduced expression of integrin αvβ8 is associated with brain arteriovenous malformation pathogenesis. Am. J. Pathol. 2010, 176, 1018–1027. [Google Scholar] [CrossRef] [PubMed]
  194. Smith, B.N.; Bhowmick, N.A. Role of EMT in metastasis and therapy resistance. J. Clin. Med. 2016, 5, 17. [Google Scholar] [CrossRef] [PubMed]
  195. Khan, Z.; Marshall, J.F. The role of integrins in TGF-β-activation in the tumour stroma. Cell Tissue Res. 2016, 365, 657–673. [Google Scholar] [CrossRef] [PubMed]
  196. Araki, K.; Shimura, T.; Suzuki, H.; Tsutsumi, S.; Wada, W.; Yajima, T.; Kobayahi, T.; Kubo, N.; Kuwano, H. E/N-cadherin switch mediates cancer progression via TGF-β-induced epithelial-to-mesenchymal transition in extrahepatic cholangiocarcinoma. Br. J. Cancer 2011, 105, 1885–1893. [Google Scholar] [CrossRef] [PubMed]
  197. Dubois, C.M.; Laprise, M.H.; Blanchette, F.; Gentry, L.E.; Leduc, R. Processing of transforming growth factor β1 precursor by human furin convertase. J. Biol. Chem. 1995, 270, 10618–10624. [Google Scholar] [CrossRef] [PubMed]
  198. Lawrence, D.A.; Pircher, R.; Kryceve-Martinerie, C.; Jullien, P. Normal embryo fibroblasts release transforming growth factors in a latent form. J. Cell Physiol. 1984, 121, 184–188. [Google Scholar] [CrossRef] [PubMed]
  199. Wipff, P.J.; Hinz, B. Integrins and the activation of latent transforming growth factor beta1—An intimate relationship. Eur. J. Cell Biol. 2008, 87, 601–615. [Google Scholar] [CrossRef] [PubMed]
  200. Shi, M.; Zhu, J.; Wang, R.; Chen, X.; Mi, L.; Walz, T.; Springer, T.A. Latent TGF-β structure and activation. Nature 2011, 474, 343–349. [Google Scholar] [CrossRef] [PubMed]
  201. Ramirez, F.; Pereira, L. The fibrillins. Int. J. Biochem. Cell Biol. 1999, 31, 255–259. [Google Scholar] [CrossRef]
  202. Ahamed, J.; Burg, N.; Yoshinaga, K.; Janczak, C.A.; Rifkin, D.B.; Coller, B.S. In vitro and in vivo evidence for shear-induced activation of latent transforming growth factor-β1. Blood 2008, 112, 3650–3660. [Google Scholar] [CrossRef] [PubMed]
  203. Annes, J.P.; Chen, Y.; Munger, J.S.; Rifkin, D.B. Integrin αvβ6-mediated activation of latent TGF-β requires the latent TGF-β binding protein-1. J. Cell Biol. 2004, 165, 723–734. [Google Scholar] [CrossRef] [PubMed]
  204. Munger, J.S.; Harpel, J.G.; Giancotti, F.G.; Rifkin, D.B. Interactions between growth factors and integrins: Latent forms of transforming growth factor-β are ligands for the integrin αvβ1. Mol. Biol. Cell 1998, 9, 2627–2638. [Google Scholar] [CrossRef] [PubMed]
  205. Robertson, I.B.; Rifkin, D.B. Regulation of the bioavailability of TGF-β and TGF-β-related proteins. Cold Spring Harb. Perspect. Biol. 2016, 8. [Google Scholar] [CrossRef] [PubMed]
  206. Eke, I.; Zscheppang, K.; Dickreuter, E.; Hickmann, L.; Mazzeo, E.; Unger, K.; Krause, M.; Cordes, N. Simultaneous β1 integrin-EGFR targeting and radiosensitization of human head and neck cancer. J. Natl. Cancer Inst. 2015, 107, dju419. [Google Scholar] [CrossRef] [PubMed]
  207. Tatler, A.L.; Goodwin, A.T.; Gbolahan, O.; Saini, G.; Porte, J.; John, A.E.; Clifford, R.L.; Violette, S.M.; Weinreb, P.H.; Parfrey, H.; et al. Amplification of TGF-β induced ITGB6 gene transcription may promote pulmonary fibrosis. PLoS ONE 2016, 11, e0158047. [Google Scholar] [CrossRef] [PubMed]
  208. Yeh, Y.Y.; Chiao, C.C.; Kuo, W.Y.; Hsiao, Y.C.; Chen, Y.J.; Wei, Y.Y.; Lai, T.H.; Fong, Y.C.; Tang, C.H. TGF-β1 increases motility and αvβ3 integrin up-regulation via PI3K, Akt and NF-κB-dependent pathway in human chondrosarcoma cells. Biochem. Pharmacol. 2008, 75, 1292–1301. [Google Scholar] [CrossRef] [PubMed]
  209. Wendt, M.K.; Smith, J.A.; Schiemann, W.P. Transforming growth factor-β-induced epithelial-mesenchymal transition facilitates epidermal growth factor-dependent breast cancer progression. Oncogene 2010, 29, 6485–6498. [Google Scholar] [CrossRef] [PubMed]
  210. Pechkovsky, D.V.; Scaffidi, A.K.; Hackett, T.L.; Ballard, J.; Shaheen, F.; Thompson, P.J.; Thannickal, V.J.; Knight, D.A. Transforming growth factor β1 induces αvβ3 integrin expression in human lung fibroblasts via a β3 integrin-, c-SRC-, and p38 MAPK-dependent pathway. J. Biol. Chem. 2008, 283, 12898–12908. [Google Scholar] [CrossRef] [PubMed]
  211. Kracklauer, M.P.; Schmidt, C.; Sclabas, G.M. TGF-β1 signaling via αvβ6 integrin. Mol. Cancer 2003, 2, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Mamuya, F.A.; Duncan, M.K. αv integrins and TGF-β-induced EMT: A circle of regulation. J. Cell. Mol. Med. 2012, 16, 445–455. [Google Scholar] [CrossRef] [PubMed]
  213. Heckmann, D.; Kessler, H. Design and chemical synthesis of integrin ligands. Methods Enzymol. 2007, 426, 463–503. [Google Scholar] [CrossRef] [PubMed]
  214. Kapp, T.G.; Rechenmacher, F.; Neubauer, S.; Maltsev, O.V.; Cavalcanti-Adam, A.E.; Zarka, R.; Reuning, U.; Notni, J.; Wester, H.-J.; Mas-Moruno, C.; et al. A comprehensive evaluation of the activity and selectivity profile of ligands for RGD-binding integrins. Sci. Rep. 2017, 7, 39805. [Google Scholar] [CrossRef] [PubMed]
  215. Dechantsreiter, M.A.; Planker, E.; Mathä, B.; Lohof, E.; Hölzemann, G.; Jonczyk, A.; Goodman, S.L.; Kessler, H. N-Methylated cyclic RGD Peptides as highly active and selective αvβ3 integrin antagonists. J. Med. Chem. 1999, 42, 3033–3040. [Google Scholar] [CrossRef] [PubMed]
  216. Mas-Moruno, C.; Rechenmacher, F.; Kessler, H. Cilengitide: The first anti-angiogenic small molecule drug candidate. Design, synthesis and clinical evaluation. Anticancer Agents Med. Chem. 2010, 10, 753–768. [Google Scholar] [CrossRef] [PubMed]
  217. Reardon, D.A.; Neyns, B.; Weller, M.; Tonn, J.C.; Nabors, L.B.; Stupp, R. Cilengitide: An RGD pentapeptide αvβ3 and αvβ5 integrin inhibitor in development for glioblastoma and other malignancies. Future Oncol. 2011, 7, 339–354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Nabors, L.B.; Fink, K.L.; Mikkelsen, T.; Grujicic, D.; Tarnawski, D.H.N.; Mazurkiewicz, M.; Salacz, M.; Ashby, L.; Zagonel, V.; Depenni, R.; et al. Two Cilengitide regimens in combination with standard treatment for patients with newly diagnosed glioblastoma and unmethylated MGMT gene promoter: Results of the open-label, controlled, randomized phase II CORE study. Neuro-Oncology 2015, 17, 708–717. [Google Scholar] [CrossRef] [PubMed]
  219. Mason, W.P. End of the road: Confounding results of the CORE trial terminate the arduous journey of Cilengitide for glioblastoma. Neuro-Oncology 2015, 17, 634–635. [Google Scholar] [CrossRef] [PubMed]
  220. Heckmann, D.; Meyer, A.; Laufer, B.; Zahn, G.; Stragies, R.; Kessler, H. Rational design of highly active and selective ligands for the α5β1 integrin receptor. ChemBioChem 2008, 9, 1397–1407. [Google Scholar] [CrossRef] [PubMed]
  221. Neubauer, S.; Rechenmacher, F.; Brimioulle, R.; Di Leva, F.S.; Bochen, A.; Sobahi, T.R.; Schottelius, M.; Novellino, E.; Mas-Moruno, C.; Marinelli, L.; et al. Pharmacophoric modifications lead to superpotent αvβ3 integrin ligands with suppressed α5β1 activity. J. Med. Chem. 2014, 57, 3410–3417. [Google Scholar] [CrossRef] [PubMed]
  222. Stragies, R.; Osterkamp, F.; Zischinsky, G.; Vossmeyer, D.; Kalkhof, H.; Reimer, U.; Zahn, G. Design and Synthesis of a new class of selective integrin α5β1 antagonists. J. Med. Chem. 2007, 50, 3786–3794. [Google Scholar] [CrossRef] [PubMed]
  223. Maltsev, O.V.; Marelli, U.K.; Kapp, T.G.; Di Leva, F.S.; Di Maro, S.; Nieberler, M.; Reuning, U.; Schwaiger, M.; Novellino, E.; Marinelli, L.; et al. Stable peptides instead of stapled peptides: Highly potent αvβ6-selective integrin ligands. Angew. Chem. Int. Ed. 2016, 55, 1535–1539. [Google Scholar] [CrossRef] [PubMed]
  224. Indrevoll, B.; Kindberg, G.M.; Solbakken, M.; Bjurgert, E.; Johansen, J.H.; Karlsen, H.; Mendizabal, M.; Cuthbertson, A. NC-100717: A versatile RGD peptide scaffold for angiogenesis imaging. Bioorg. Med. Chem. Lett. 2006, 16, 6190–6193. [Google Scholar] [CrossRef] [PubMed]
  225. Hynes, R.O. Integrins: Versatility, modulation, and signaling in cell adhesion. Cell 1992, 69, 11–25. [Google Scholar] [CrossRef]
  226. Aumailley, M.; Gurrath, M.; Müller, G.; Calvete, J.; Timpl, R.; Kessler, H. Arg-Gly-Asp constrained within cyclic pentapeptides. Strong and selective inhibitors of cell adhesion to vitronectin and laminin fragment P1. FEBS Lett. 1991, 291, 50–54. [Google Scholar] [CrossRef]
  227. Haubner, R.; Finsinger, D.; Kessler, H. Stereoisomeric Peptide Libraries and Peptidomimetics for designing selective inhibitors of the αvβ3 integrin for a new cancer therapy. Angew. Chem. Int. Ed. 1997, 36, 1374–1389. [Google Scholar] [CrossRef]
  228. Gurrath, M.; Müller, G.; Kessler, H.; Aumailley, M.; Timpl, R. Conformation/activity studies of rationally designed potent anti-adhesive RGD peptides. Eur. J. Biochem. 1992, 210, 911–921. [Google Scholar] [CrossRef] [PubMed]
  229. Chatterjee, J.; Rechenmacher, F.; Kessler, H. N-Methylation of peptides and proteins: An important element for modulating biological functions. Angew. Chem. Int. Ed. 2013, 52, 254–269. [Google Scholar] [CrossRef] [PubMed]
  230. Coller, B.S.; Shattil, S.J. The GPIIb/IIIa (integrin αIIbβ3) odyssey: A technology-driven saga of a receptor with twists, turns, and even a bend. Blood 2008, 112. [Google Scholar] [CrossRef] [PubMed]
  231. Zhu, J.; Luo, B.-H.; Xiao, T.; Zhang, C.; Nishida, N.; Springer, T.A. Structure of a complete integrin ectodomain in a physiologic resting state and activation and deactivation by applied forces. Mol. Cell 2008, 32, 849–861. [Google Scholar] [CrossRef] [PubMed]
  232. Nagae, M.; Re, S.; Mihara, E.; Nogi, T.; Sugita, Y.; Tagaki, J. Crystal structure of α5β1 integrin ectodomain: Atomic details of the fibronectin receptor. J. Cell Biol. 2012, 187, 131–140. [Google Scholar] [CrossRef] [PubMed]
  233. Pallarola, D.; Platzman, I.; Bochen, A.; Cavalcanti-Adam, E.A.; Axmann, M.; Kessler, H.; Geiger, B.; Spatz, J.P. Focal adhesion stabilization by enhanced integrin-cRGD binding affinity. BioNanoMaterials 2017, 18, 1–2. [Google Scholar] [CrossRef]
  234. Ley, K.; Rivera-Nieves, J.; Sandborn, W.J.; Shattil, S. Integrin-based therapeutics: Biological basis, clinical use and new drugs. Nat. Rev. Drug Discov. 2016, 15, 173–183. [Google Scholar] [CrossRef] [PubMed]
  235. Hovlid, M.L.; Steinmetz, N.F.; Laufer, B.; Lau, J.L.; Kuzelka, J.; Wang QHyypiä, T.; Nemerow, G.R.; Kessler, H.; Manchester, M.; Finn, M.G. Guiding plant virus particles to integrin-displaying cells. Nanoscale 2012, 4, 3698–3705. [Google Scholar] [CrossRef] [PubMed]
  236. Shroff, K.; Kokkoli, E. PEGylated liposomal doxorubicin targeted to α5β1-expressing MDA-MB-231 breast cancer cells. Langmuir 2012, 28, 4729–4736. [Google Scholar] [CrossRef] [PubMed]
  237. Beer, A.J.; Haubner, R.; Goebel, M.; Luderschmidt, S.; Spilker, M.E.; Wester, H.J.; Weber, W.A.; Schwaiger, M. Biodistribution and pharmacokinetics of the αvβ3-selective tracer 18F-galacto-RGD in cancer patients. J. Nucl. Med. 2005, 46, 133–141. [Google Scholar]
  238. Beer, A.J.; Haubner, R.; Sarbia, M.; Goebel, M.; Luderschmidt, S.; Grosu, A.L.; Schnell, O.; Niemeyer, M.; Kessler, H.; Wester, H.J.; et al. Positron emission tomography using 18F-Galacto-RGD identifies the level of integrin αvβ3 expression in man. Clin. Cancer Res. 2006, 12, 3942–3949. [Google Scholar] [CrossRef] [PubMed]
  239. Haubner, R.; Weber, W.A.; Beer, A.J.; Vabuliene, E.; Reim, D.; Sarbia, M.; Becker, K.F.; Goebel, M.; Hein, R.; Wester, H.J.; et al. Noninvasive visualization of the activated αvβ3 integrin in cancer patients by positron emission tomography and 18F-Galacto-RGD. PLoS Med. 2005, 2, e70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Chen, H.; Niu, G.; Wu, H.; Chen, X. Clinical application of radiolabeled RGD peptides for PET imaging of integrin αvβ3. Theranostics 2016, 6, 78–92. [Google Scholar] [CrossRef] [PubMed]
  241. Beer, A.J.; Haubner, R.; Wolf, I.; Goebel, M.; Luderschmidt, S.; Niemeyer, M.; Grosu, A.-L.; Martinez, M.-J.; Wester, H.J.; Weber, W.A.; et al. PET-based human dosimetry of 18F-galacto-RGD, a new radiotracer for imaging αvβ3 expression. J. Nucl. Med. 2006, 47, 763–769. [Google Scholar] [PubMed]
  242. McParland, B.J.; Miller, M.P.; Spinks, T.J.; Kenny, L.M.; Osman, S.; Khela, M.K.; Aboagye, E.; Coombes, R.C.; Hui, A.M.; Cohen, P.S. The biodistribution and radiation dosimetry of the Arg-Gly-Asp peptide 18F-AH111585 in healthy volunteers. J. Nucl. Med. 2008, 49, 1664–1667. [Google Scholar] [CrossRef] [PubMed]
  243. Doss, M.; Kolb, H.C.; Zhang, J.J.; Belanger, M.J.; Stubbs, J.B.; Stabin, M.G.; Hostetler, E.D.; Alpaugh, R.K.; von Mehren, M.; Walsh, J.C.; et al. Biodistribution and radiation dosimetry of the integrin marker 18F-RGD-K5 determined from whole-body PET/CT in monkeys and humans. J. Nucl. Med. 2012, 53, 787–795. [Google Scholar] [CrossRef] [PubMed]
  244. Kim, J.H.; Lee, J.S.; Kang, K.W.; Lee, H.Y.; Han, S.W.; Kim, T.Y.; Lee, Y.S.; Jeong, J.M.; Lee, D.S. Whole-body distribution and radiation dosimetry of 68Ga-NOTA-RGD, a positron emission tomography agent for angiogenesis imaging. Cancer Biother. Radiopharm. 2012, 27, 65–71. [Google Scholar] [CrossRef] [PubMed]
  245. Wan, W.; Guo, N.; Pan, D.; Yu, C.; Weng, Y.; Luo, S.; Ding, H.; Xu, Y.; Wang, L.; Lang, L.; et al. First experience of 18F-Alfatide in lung cancer patients using a new lyophilized kit for rapid radiofluorination. J. Nucl. Med. 2013, 54, 691–698. [Google Scholar] [CrossRef] [PubMed]
  246. Mittra, E.S.; Goris, M.L.; Iagaru, A.H.; Kardan, A.; Burton, L.; Berganos, R.; Chang, E.; Lui, S.; Shen, B.; Chin, F.T.; et al. Pilot pharmacokinetic and dosimetric studies of 18F-FPPRGD2: A PET radiopharmaceutical agent for imaging αvβ3 integrin levels. Radiology 2011, 260, 182–191. [Google Scholar] [CrossRef] [PubMed]
  247. Yu, C.; Pan, D.; Mi, B.; Xu, Y.; Lang, L.; Niu, G.; Yang, M.; Wan, W.; Chen, X. 18F-Alfatide II PET/CT in healthy human volunteers and patients with brain metastases. Eur. J. Nucl. Med. Mol. Imaging 2015, 42, 2021–2028. [Google Scholar] [CrossRef] [PubMed]
  248. Zheng, K.; Liang, N.; Zhang, J.; Lang, L.; Zhang, W.; Li, S.; Zhao, J.; Niu, G.; Li, F.; Zhu, Z.; et al. 68Ga-NOTA-PRGD2 PET/CT for integrin imaging in patients with lung cancer. J. Nucl. Med. 2015, 56, 1823–1827. [Google Scholar] [CrossRef] [PubMed]
  249. Atkinson, S.J.; Ellison, T.S.; Steri, V.; Gould, E.; Robinson, S.D. Redefining the role(s) of endothelial αvβ3-integrin in angiogenesis. Biochem. Soc. Trans. 2014, 42, 1590–1595. [Google Scholar] [CrossRef] [PubMed]
  250. Fässler, R.; Meyer, M. Consequences of lack of β1 integrin gene expression in mice. Genes Dev. 1995, 9, 1896–1908. [Google Scholar] [CrossRef] [PubMed]
  251. Tanjore, H.; Zeisberg, E.M.; Gerami-Naini, B.; Kalluri, R. β1 integrin expression on endothelial cells is required for angiogenesis but not for vasculogenesis. Dev. Dyn. 2007, 237, 75–82. [Google Scholar] [CrossRef] [PubMed]
  252. Neubauer, S.; Rechenmacher, F.; Beer, A.J.; Curnis., F.; Pohle, K.; D’Alessandria, C.; Wester, H.J.; Reuning, U.; Corti, A.; Schwaiger, M.; et al. Selective imaging of the angiogenic relevant integrins α5β1 and αvβ3. Angew. Chem. Int. Ed. 2013, 52, 11656–11659. [Google Scholar] [CrossRef] [PubMed]
  253. D’Alessandria, C.; Pohle., K.; Rechenmacher, S.; Neubauer, S.; Notni, J.; Wester, H.J.; Schwaiger, M.; Kessler, H.; Beer, A.J. In vivo biokinetic and metabolic characterization of the 68Ga-labeled α5β1-selective peptidomimetic FR366. Eur. J. Nucl. Med. Mol. Imaging 2016, 43, 953–963. [Google Scholar] [CrossRef]
  254. Haubner, R.; Maschauer, S.; Einsiedel, J.; Eder, I.E.; Rangger, C.; Gmeiner, P.; Virgolini, I.; Prante, O. H-CRRETAWAC-OH, a lead structure for the development of radiotracer targeting integrin α5β1? Biomed. Res. Int. 2014, 2014, 243185. [Google Scholar] [CrossRef] [PubMed]
  255. Zhao, H.T.; Gao, H.N.; Zhai, L.P.; Liu, X.J.; Jia, B.; Shi, J.Y.; Wang, F. Tc-99m-HisoDGR as a Potential SPECT Probe for orthotopic glioma detection via targeting of integrin α5β1. Bioconjug. Chem. 2016, 27, 1259–1266. [Google Scholar] [CrossRef] [PubMed]
  256. Notni, J.; Steiger, K.; Hoffmann, F.; Reich, D.; Kapp, F.G.; Rechenmacher, T.; Neubauer, S.; Kessler, H.; Wester, H.J. Complementary, selective PET-Imaging of integrin subtypes α5β1 and αvβ3 using Ga-68-aquibeprin and Ga-68-avebetrin. J. Nucl. Med. 2016, 57, 460–466. [Google Scholar] [CrossRef] [PubMed]
  257. Notni, J.; Steiger, K.; Hoffmann, F.; Reich, D.; Schwaiger, M.; Kessler, H.; Wester, H.J. Variation of specific activities of Ga-68-aquibeprin and Ga-68-avebetrin enables selective PET-imaging of different expression levels of integrins α5β1 and αvβ3. J. Nucl. Med. 2016, 57, 1618–1624. [Google Scholar] [CrossRef] [PubMed]
  258. Liu, H.; Wu, Y.; Wang, F.; Liu, Z. Molecular imaging of integrin αvβ6 expression in living subjects. Am. J. Nucl. Med. Mol. Imaging 2014, 4, 333–345. [Google Scholar] [PubMed]
  259. Zhang, Y.; Sun, Y.; Yang, F.; Guo, J.; He, J.; Wu, Q.; Cao, W.; Lv, L.; Zheng, H.; Zhang, Z. Induction of partial protection against foot and mouth disease virus in guinea pigs by neutralization with the integrin β6-1 subunit. Viruses 2013, 5, 1114–1130. [Google Scholar] [CrossRef] [PubMed]
  260. Goodman, S.L.; Hölzemann, G.; Sulyok, G.A.; Kessler, H. Nanomolar small molecule inhibitors for αvβ6, αvβ5, and αvβ3 integrins. J. Med. Chem. 2002, 45, 1045–1051. [Google Scholar] [CrossRef] [PubMed]
  261. Kraft, S.; Diefenbach, B.; Mehta, R.; Jonczyk, A.; Luckenbach, G.A.; Goodman, S.L. Definition of an unexpected ligand recognition motif for αvβ6 integrin. J. Biol. Chem. 1999, 274, 1979–1985. [Google Scholar] [CrossRef] [PubMed]
  262. Hausner, S.H.; DiCara, D.; Marik, J.; Marshall, J.F.; Sutcliffe, J.L. Use of a peptide derived from foot-and-mouth disease virus for the noninvasive imaging of human cancer: Generation and evaluation of 4-[18f]fluorobenzoyl a20fmdv2 for in vivo imaging of integrin αvβ6 expression with positron emission tomography. Cancer Res. 2007, 67, 7833–7840. [Google Scholar] [CrossRef] [PubMed]
  263. Li, S.; McGuire, M.J.; Lin, M.; Liu, Y.H.; Oyama, T.; Sun, X.; Brown, K.C. Synthesis and characterization of a high-affinity αvβ6-specific ligand for in vitro and in vivo applications. Mol. Cancer Ther. 2009, 8, 1239–1249. [Google Scholar] [CrossRef] [PubMed]
  264. Kimura, R.H.; Teed, R.; Hackel, B.J.; Pysz, M.A.; Chuang, C.Z.; Sathirachinda, A.; Willmann, J.K.; Gambhir, S.S. Pharmacokinetically stabilized cystine knot peptides that bind αvβ6 integrin with single-digit nanomolar affinities for detection of pancreatic cancer. Clin. Cancer Res. 2012, 18, 839–849. [Google Scholar] [CrossRef] [PubMed]
  265. Ahmedah, H.T.; Patterson, L.H.; Shnyder, S.D.; Sheldrake, H.M. RGD-binding integrins in head and neck cancers. Cancers 2017, 9, 56. [Google Scholar] [CrossRef] [PubMed]
  266. John, A.E.; Luckett, J.C.; Tatler, A.L.; Awais, R.O.; Desai, A.; Habgood, A.; Ludbrook, S.; Blanchard, A.D.; Perkins, A.C.; Jenkins, R.G.; et al. Preclinical SPECT/CT imaging of αvβ6 integrins for molecular stratification of idiopathic pulmonary fibrosis. J. Nucl. Med. 2013, 54, 2146–2152. [Google Scholar] [CrossRef] [PubMed]
  267. Liu, Z.; Liu, H.; Ma, T.; Sun, X.; Shi, J.; Jia, B.; Sun, Y.; Zhan, J.; Zhang, H.; Zhu, Z.; et al. Integrin αvβ6-targeted SPECT imaging for pancreatic cancer detection. J. Nucl. Med. 2014, 55, 989–994. [Google Scholar] [CrossRef] [PubMed]
  268. Zhu, X.; Li, J.; Hong, Y.; Kimura, R.H.; Ma, X.; Liu, H.; Qin, C.; Hu, X.; Hayes, T.R.; Benny, P.; et al. 99mTC-labeled cystine knot peptide targeting integrin αvβ6 for tumor SPECT imaging. Mol. Pharm. 2014, 11, 1208–1217. [Google Scholar] [CrossRef] [PubMed]
  269. Hausner, S.H.; Abbey, C.K.; Bold, R.J.; Gagnon, M.K.; Marik, J.; Marshall, J.F.; Stanecki, C.E.; Sutcliffe, J.L. Targeted in vivo imaging of integrin αvβ6 with an improved radiotracer and its relevance in a pancreatic tumor model. Cancer Res. 2009, 69, 5843–5850. [Google Scholar] [CrossRef] [PubMed]
  270. Singh, A.N.; McGuire, M.J.; Li, S.; Hao, G.; Kumar, A.; Sun, X.; Brown, K.C. Dimerization of a phage-display selected peptide for imaging of αvβ6-integrin: Two approaches to the multivalent effect. Theranostics 2014, 4, 745–760. [Google Scholar] [CrossRef] [PubMed]
  271. Hausner, S.H.; Bauer, N.; Sutcliffe, J.L. In vitro and in vivo evaluation of the effects of aluminum [18F]fluoride radiolabeling on an integrin αvβ6-specific peptide. Nucl. Med. Biol. 2014, 41, 43–50. [Google Scholar] [CrossRef] [PubMed]
  272. Hausner, S.H.; Bauer, N.; Hu, L.Y.; Knight, L.M.; Sutcliffe, J.L. The effect of bi-terminal PEGylation of an integrin αvβ6-targeted 18F peptide on pharmacokinetics and tumor uptake. J. Nucl. Med. 2015, 56, 784–790. [Google Scholar] [CrossRef] [PubMed]
  273. Hausner, S.H.; Carpenter, R.D.; Bauer, N.; Sutcliffe, J.L. Evaluation of an integrin αvβ6-specific peptide labeled with [18F]fluorine by copper-free, strain-promoted click chemistry. Nucl. Med. Biol. 2013, 40, 233–239. [Google Scholar] [CrossRef] [PubMed]
  274. Altmann, A.; Sauter, M.; Roesch, S.; Mier, W.; Warta, R.; Debus, J.; Dyckhoff, G.; Herold-Mende, C.; Haberkorn, U. Identification of a novel ITGαvβ6-binding peptide using protein separation and phage display. Clin. Cancer Res. 2017. [Google Scholar] [CrossRef] [PubMed]
  275. Notni, J.; Reich, D.; Maltsev, O.V.; Kapp, T.G.; Steiger, K.; Hoffmann, F.; Esposito, I.; Weichert, W.; Kessler, H.; Wester, H.J. In Vivo pet imaging of the “cancer integrin” alphavbeta6 using gallium-68 labelled cyclic RGD nonapeptides. J. Nucl. Med. 2016, 58, 671–677. [Google Scholar] [CrossRef] [PubMed]
  276. Zhang, Z.-Y.; Xu, K.; He, Q.-S.; Niu, W.-B.; Wang, J.-Y.; Mi, Y.-T.; Wang, J.-S.; Wang, G.-Q.; Yang, G.-Y.; Niu, J. Signaling and regulatory mechanisms of integrin αvβ6 on the apoptosis of colon cancer cells. Cancer Lett. 2008, 266, 209–215. [Google Scholar] [CrossRef] [PubMed]
  277. Kogelberg, H.; Tolner, B.; Thomas, G.J.; Di Cara, D.; Minogue, S.; Ramesh, B.; Sodha, S.; Marsh, D.; Lowdell, M.W.; Meyer, T.; et al. Engineering a single-chain fv antibody to αvβ6 integrin using the specificity-determining loop of a foot-and-mouth disease virus. J. Mol. Biol. 2008, 382, 385–401. [Google Scholar] [CrossRef] [PubMed]
  278. Integrin Alpha-v-Beta and [18F]-R01-MG-F2 PET/CT in Measuring Response in Patients with Pancreatic Cancer and Healthy Volunteers. Available online: https://clinicaltrials.gov/ct2/show/NCT02683824 (accessed on 7 August 2017).
  279. First-in-Human Positron Emission Tomography Study Using the 18F-αvβ6-Binding Peptide. Available online: https://clinicaltrials.gov/ct2/show/NCT03164486 (accessed on 7 August 2017).
Figure 1. Schematic illustration of integrin activation and the “inside-out” and “outside-in” signaling mechanism. Integrins in the bent resting conformation reveal low affinity binding to their ECM ligands. The inside-out signaling involves disruption of the intracellular salt bridge, which is established between the cytoplasmic subunits. This induces dissociation of the transmembrane helices, followed by the reorganization and generation of a high affinity binding integrin, plus multimerization in focal adhesions. Conformational changes of the resting integrin state are induced by the integrin binding of ECM ligands causing stronger binding at the focal adhesions. The outside-in signaling requires integrin oligomerization [35].
Figure 1. Schematic illustration of integrin activation and the “inside-out” and “outside-in” signaling mechanism. Integrins in the bent resting conformation reveal low affinity binding to their ECM ligands. The inside-out signaling involves disruption of the intracellular salt bridge, which is established between the cytoplasmic subunits. This induces dissociation of the transmembrane helices, followed by the reorganization and generation of a high affinity binding integrin, plus multimerization in focal adhesions. Conformational changes of the resting integrin state are induced by the integrin binding of ECM ligands causing stronger binding at the focal adhesions. The outside-in signaling requires integrin oligomerization [35].
Cancers 09 00116 g001
Figure 2. Integrin functions are instrumental in tumor biological contexts. Integrin functions are involved in various tumor biological processes, including cell adhesion, proliferation, inhibition of apoptosis/anoikis, induction of angiogenesis, cell invasion and migration. Adhesion: Cell adhesion is mediated by integrin binding to the respective recognition motif of RGD-ligands within the ECM, e.g., fibronectin, osteopontin and vitronectin. Binding of RGD ligands to integrins enables communication between the ECM and intracellular components, such as the cytoskeleton. Proliferation: Cell proliferation, mediated by integrin subtypes, such as αvβ3, αvβ6, and αvβ8 may also be induced upon the binding of the RGD-containing latency associated peptide (LAP) of the inactive TGF-β to integrins. Force between the TGF-β binding proteins (LTBP) and the TGF-β molecule results in the activation of the latent TGF-β. Subsequent binding of TGF-β to the TGF-β receptor (TGF-β-R) induces epithelial-mesenchymal transition (EMT) and cell proliferation. Apoptosis/anoikis: Further, integrin expression allows cells to bind to ECM molecules within a mesenchymal tissue context, thereby inhibiting apoptosis/anoikis of invasive carcinoma cells. Angiogenesis: Upon tumor growth, the expression of hypoxia induced factors and vascular endothelial growth factor (VEGF) results in the induction of integrin-associated vessel sprouting and angiogenesis. Invasion/migration: The integrin switch from αvβ1 and/or αvβ5 to αvβ6 allows cells to migrate and cross the basement membrane (BM) to invade the surrounding ECM as invasive cancer cells. This process is associated with EMT-like alterations of the cell phenotype. The EMT process involves the downregulation of E-cadherin and cell adhesion molecules (CAMs), promoting cell mobility, cancer progression and metastasis.
Figure 2. Integrin functions are instrumental in tumor biological contexts. Integrin functions are involved in various tumor biological processes, including cell adhesion, proliferation, inhibition of apoptosis/anoikis, induction of angiogenesis, cell invasion and migration. Adhesion: Cell adhesion is mediated by integrin binding to the respective recognition motif of RGD-ligands within the ECM, e.g., fibronectin, osteopontin and vitronectin. Binding of RGD ligands to integrins enables communication between the ECM and intracellular components, such as the cytoskeleton. Proliferation: Cell proliferation, mediated by integrin subtypes, such as αvβ3, αvβ6, and αvβ8 may also be induced upon the binding of the RGD-containing latency associated peptide (LAP) of the inactive TGF-β to integrins. Force between the TGF-β binding proteins (LTBP) and the TGF-β molecule results in the activation of the latent TGF-β. Subsequent binding of TGF-β to the TGF-β receptor (TGF-β-R) induces epithelial-mesenchymal transition (EMT) and cell proliferation. Apoptosis/anoikis: Further, integrin expression allows cells to bind to ECM molecules within a mesenchymal tissue context, thereby inhibiting apoptosis/anoikis of invasive carcinoma cells. Angiogenesis: Upon tumor growth, the expression of hypoxia induced factors and vascular endothelial growth factor (VEGF) results in the induction of integrin-associated vessel sprouting and angiogenesis. Invasion/migration: The integrin switch from αvβ1 and/or αvβ5 to αvβ6 allows cells to migrate and cross the basement membrane (BM) to invade the surrounding ECM as invasive cancer cells. This process is associated with EMT-like alterations of the cell phenotype. The EMT process involves the downregulation of E-cadherin and cell adhesion molecules (CAMs), promoting cell mobility, cancer progression and metastasis.
Cancers 09 00116 g002
Figure 3. Development and testing of selective integrin ligands. Depicted are some integrin-specific ligands according to their selectivity profile (IC50-values) for the subtypes αvβ3, αvβ5, αvβ6, αvβ8, α5β1, and αIIbβ3 [214]. The binding affinities of the ligands have been determined in a cell-free, enzyme-linked immunosorbent assay (ELISA)-like assay for comparability. None of the selected compounds has significant affinity for the platelet integrin αIIbβ3. Specificity or subtype with the lowest IC50-values are highlighted. The preferential site for modification with diagnostic or therapeutic agents is indicated by an ellipsoid tag.
Figure 3. Development and testing of selective integrin ligands. Depicted are some integrin-specific ligands according to their selectivity profile (IC50-values) for the subtypes αvβ3, αvβ5, αvβ6, αvβ8, α5β1, and αIIbβ3 [214]. The binding affinities of the ligands have been determined in a cell-free, enzyme-linked immunosorbent assay (ELISA)-like assay for comparability. None of the selected compounds has significant affinity for the platelet integrin αIIbβ3. Specificity or subtype with the lowest IC50-values are highlighted. The preferential site for modification with diagnostic or therapeutic agents is indicated by an ellipsoid tag.
Cancers 09 00116 g003

αvβ3
[nM]
αvβ5
[nM]
αvβ6
[nM]
αvβ8
[nM]
α5β1
[nM]
αIIbβ3
[nM]
Ref.
1a>10,000>10,000433 ± 10137 ± 32.3 ± 0.02>10,000[220]
2a0.65 ± 0.05199 ± 21>10,000>10,000108 ± 27.5>10,000[221]
3>10,000>10,00023 ± 3.48.2 ± 0.522.5 ± 0.4>10,000[222]
4a0.61 ± 0.068.4 ± 2.12050 ± 6402350 ± 43814.9 ± 3.15400 ± 814[216]
5a1200 ± 240>10,0000.28 ± 0.01924 ± 3.173 ± 6>10,000[223]
6a1.1 ± 0.116.7 ± 2.1>10,000>10,000820 ± 156>10,000[224]
Figure 4. Imaging of various integrin subtypes.
Figure 4. Imaging of various integrin subtypes.
Cancers 09 00116 g004
Table 1. Overview table of RGD-binding integrins expressed in various human cancers at the tissue level with respect to the expressing cell type and, if available, the prognostic impact.
Table 1. Overview table of RGD-binding integrins expressed in various human cancers at the tissue level with respect to the expressing cell type and, if available, the prognostic impact.
Integrin Cancer TypeCell TypeMain Expression FeatureReference
αvβ3gastric cancertumor, endothelial and stromal cellslow to moderate expression frequency in tumor cells, high frequency in stroma and endothelia, correlates with phenotype, endothelial expression correlates with survival[60]
gliomaendothelial and tumor cellsexpression correlates with grade[61]
lung cancer brain metastasesendothelial and tumor cellshigh expression frequency in endothelial, low frequency in tumor cells[62]
non small cell lung cancerendothelial and tumor cellshigh expression frequency in endothelia, low frequency in tumor cells, no correlation with survival[63]
oral squamous cell carcinomaendothelial cellshigher expression in intratumoral endothelia compared with control tissue[64]
pancreatic cancertumor cellsmoderate expression frequency, involved in lymph node metastasis[65]
prostate cancerendothelial cellshigh expression frequency peritumoral[66]
αvβ5gastric cancertumor, endothelial and stromal cellsmoderate (to high) frequency in tumor cells, high frequency in stroma and endothelial cells, independent prognostic factor in intestinal-type[60]
lung cancer (with brain metastases)vessel endothelia and tumor cellshigh expression frequency in endothelia, low frequency in tumor cells[62]
non small cell lung cancertumor cells and stromahigh frequency in tumor and stroma cells, no correlation with survival[63]
oral squamous cell carcinomatumor cells and stroma [64]
prostate cancertumor cellsexpression influenced by differentiation[66]
αvβ6basal cell carcinomatumor cellshigher expression frequency in infiltrative subtype[67]
breast cancerexpression correlates with prognosis[68]
colon cancerupregulated at invasive front and in budding tumor cells[69]
endometrial canceroften overexpressed without correlation with occurrence of lymph node metastasis[70]
gastric cancerpotential prognostic marker in early stage carcinoma[53]
liverdifferentiates cholangiocarcinoma from hepatocellular carcinoma [52]
non small cell lung cancerhigh expression frequency with intratumoral heterogeneity, no correlation with survival[71]
lung cancer brain metastaseshigh expression frequency[62]
oral squamous cell carcinomaexpression at invasive front[72]
ovarian cancerexpression correlates with grade[55]
pancreatic cancerhigh expression frequency [51]
[73]
prostate cancernot/weakly expressed[66]
αvβ8non small cell lung cancertumor cellslow to moderate expression frequency, no correlation with survival[63]
prostate cancer not expressed[66]
α5β1oral squamous cell carcinomatumor, endothelial cells, stromastrong expression in stroma, expressed also in tumor and endothelial cells[64]
ovarian cancertumor cellsmoderate expression frequency, correlates with survival[74]

Share and Cite

MDPI and ACS Style

Nieberler, M.; Reuning, U.; Reichart, F.; Notni, J.; Wester, H.-J.; Schwaiger, M.; Weinmüller, M.; Räder, A.; Steiger, K.; Kessler, H. Exploring the Role of RGD-Recognizing Integrins in Cancer. Cancers 2017, 9, 116. https://doi.org/10.3390/cancers9090116

AMA Style

Nieberler M, Reuning U, Reichart F, Notni J, Wester H-J, Schwaiger M, Weinmüller M, Räder A, Steiger K, Kessler H. Exploring the Role of RGD-Recognizing Integrins in Cancer. Cancers. 2017; 9(9):116. https://doi.org/10.3390/cancers9090116

Chicago/Turabian Style

Nieberler, Markus, Ute Reuning, Florian Reichart, Johannes Notni, Hans-Jürgen Wester, Markus Schwaiger, Michael Weinmüller, Andreas Räder, Katja Steiger, and Horst Kessler. 2017. "Exploring the Role of RGD-Recognizing Integrins in Cancer" Cancers 9, no. 9: 116. https://doi.org/10.3390/cancers9090116

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop