GET THE APP

Ordered Chaos: Harnessing Developmental Pathways in Tumor-Induced
Journal of Clinical and Cellular Immunology

Journal of Clinical and Cellular Immunology
Open Access

ISSN: 2155-9899

+44 1223 790975

Review Article - (2014) Volume 5, Issue 5

Ordered Chaos: Harnessing Developmental Pathways in Tumor-Induced Lymphangiogenesis

Francois M1*, Shayan R2 and Karnezis T3,4
1Division of Genomics of Development and Diseases, Institute for Molecular Bioscience, The University of Queensland, St Lucia, QLD 4072, Australia
2O’Brien Institute, Fitzroy, Victoria 3065, Australian Catholic University, Fitzroy, VIC, 3065, Australia
3Tumor Angiogenesis Program, Peter MacCallum Cancer Centre, East Melbourne, Victoria 3002, Australia
4Sir Peter MacCallum Department of Oncology, University of Melbourne, Parkville, VIC 3010, Australia
*Corresponding Author: Dr. Francois M, Division of Genomics of Development and Diseases, Institute for Molecular Bioscience, University of Queensland, 306 Carmody Road, St Lucia 4072, Queensland, Australia, Tel: +61 7 3346 2494 Email:

Abstract

Over the past decade, research investigating the development of the lymphatic vascular system has become a key focus within the fields of developmental biology and cancer biology. Critical discoveries relating to the specification of lymphatic endothelial cells and lymphatic morphogenesis during embryonic development have helped to identify novel molecular drivers of pathological lymphangiogenesis; that is, the sprouting of a new lymphatic vessel from, or the enlargement of, pre-existing lymphatic vessels. These pathways may also constitute potentially useful therapeutic targets, which may be exploited to restrict tumor spread via the lymphatic system. Here, we discuss the current knowledge of both developmental and tumor-induced lymphangiogenesis, and draw parallels between the two processes to describe the molecular pathways that are re-capitulated during the growth of tumor lymphatics, and which promote metastasis.

Keywords: Lymphangiogenesis; Transcription factor; Metastasis; Growth factors; Developmental genetics; Solid cancer

Introduction

The lymphatic vasculature forms a complex network that parallels the blood vascular system to drain fluid from, and thus regulate the homeostasis of, the interstitial tissues. The lymphatic vascular tree is a key feature of vertebrate physiology and plays a central role in lipid and hormone transport as well as immune cell trafficking. For over a century, the cellular origin of lymphatic endothelial cells (LECs) has been debated [1,2]. More recently, molecular advances in developmental genetics and imaging have shed light onto the processes that instruct lymphangiogenesis; both in development and as part of several key human diseases. In adult vertebrates, lymphangiogenesis plays an important role in cancer metastasis, a process that determines patient mortality [3-5]. Discoveries of the molecular pathways that underpin lymphangiogenesis are therefore pivotal to developing novel therapeutic avenues to restrict cancer spread [6]. In this review we highlight parallels between the fundamental molecular processes underlying lymphatic vascular development, and the same processes that become dysregulated during tumor-induced lymphangiogenesis.

Molecular Pathways Driving Embryonic Lymphangiogenesis

In both fish and mammals, the lymphatic vasculature has been shown to arise from venous endothelial cell precursors [7,8]. Major sources of LEC-precursors that contribute to establishing the primitive lymphatic plexus (9.5 day post-coïtum ( dpc)-14.5 dpc in the mouse) have a venous origin and were identified by lineage tracing experiments and advanced imaging methods [9-12]. Vascular beds that establish a pool of LEC-precursors include the cardinal and inter-somitic veins and the superficial venous plexus (Figure 1A).

cellular-immunology-molecular-profile

Figure 1: “Ordered chaos”: Key developmental pathways activated during embryonic development face re-activation under tumor conditions to drive aberrant lymphangiogenesis and metastatic spread. (A) Schematic representation of embryonic lymphangiogenesis and the major genetic pathways that instruct lymphatic endothelial cell fate (SOX18, COUP-TFII, PROX1), maintain venous identity (Notch, RA) and govern morphogenesis of the lymphatic vascular plexus (VEGF-C, VEGF- R3, CCBE1, TGF-B, BMP9 and BMP2) from 9 dpc-14.5 dpc in mice. During development venous endothelial cells provide LEC precursor cells and macrophages control lymphatic vessel calibre. (B) Under tumor-induced conditions, embryonic pathways are re-activated or dysregulated to cause expansion of the pre-existing lymphatic vasculature in and around tumor tissues. Neo-lymphatics form an ‘on- ramp’ for tumor cells and fluid draining to the lymph node basin. Reactivation of the developmental program is influenced by the inflammatory response, components like tumor-infiltrated macrophages produce VEGF-C and change their molecular profile to acquire a LEC-like molecular signature. These factors contribute to produce a pseudo-functional lymphatic vasculature.

In order to shift from a venous endothelial cell identity to a LEC fate, a subset of the venous endothelial cells begins to express a finely-tuned combination of transcription factors and growth factors/receptors that subsequently induce the development and morphogenesis of the lymphatic vascular plexus [13].

Transcriptional control of LEC specification

Gain- and loss-of-function experiments using transgenic mouse models have demonstrated that only a handful of transcription factors have been shown to guide lymphangiogenesis. Following the onset of arterio-venous specification (around 7.75-9 dpc in the mouse), a sub-population of the endothelium that expresses SOX18 (Sry-related HMG box containing-18) and COUP-TFII (chicken ovalbumin upstream promoter or NR2F2) becomes restricted to the dorso-lateral part of the anterior cardinal vein to form pre-lymphatic clusters, which in turn form a reservoir of LEC- precursors (up to 14.5 dpc) [9]. At this time point, the primary lymphatic structure has been established and is composed of lymphatic sacs, also described as pre-thoracic ducts [10] that later form the deeper collecting lymphatic vessels. In contrast, the superficial lymphatic vascular network is still expanding and completes the colonization of organs such as the skin or heart at significantly later embryonic stages (e.g. 17-18 dpc).

A critical transcription factor is SOX18, which -along with the endothelium-specific SOX7 and SOX17 -belongs to the SOX-F group. One function of SOX18 is to directly transactivate a downstream transcription factor, Prox1 [14], which act as the gatekeeper of LEC identity [7,15]. Recent work has shown that gain-of-function of the RAF1/MEK/ERK pathway is able to constitutively modulate SOX18 activity and trigger excessive lymphangiogenesis [16].

Similarly, COUP-TFII, which is essential for maintenance of venous identity, has also been shown to directly induce Prox1 transcription [17,18]. Later during development, COUP-TFII additionally causes direct transactivation of neuropilin-2 (NRP2), a co- receptor of VEGF-R3 that is essential in driving LEC migration within mesenchymal tissues [19].

Once LEC cell fate has been induced by SOX18 and COUP-TFII, the expression of Prox1 is also required for maintenance of LEC identity and for acquiring additional lymphatic-specific markers [20]. Conditional deletion of the Prox1 gene in adult LECs led to a reversal from a LEC identity into a blood endothelial cell phenotype [21]; whereas SOX18 signaling ceases after 14.5 dpc and is not essential to the maintenance of a LEC phenotype.

External cues that influence lymphangiogenesis

Despite the lymphatic potential of the venous endothelium, not all venous cells commit to a LEC fate. Rather, some segments of the vein must remain intact and spared from undergoing lymphangiogenesis in order to preserve the functional integrity of the blood circulatory system. To this end, negative regulation of lymphangiogenesis has been shown to be mediated by at least three independent pathways: Cyp26b1 (a retinoic acid (RA) degradation enzyme), Notch and TGF-β/BMP signaling. Supporting this negative regulatory role, loss of Cyp26b1 function in vivo was shown to lead to an increased local concentration of RA, which in turn stimulated LEC proliferation, aberrant lymphatic vessel formation and enlarged lymph sacs [22,23].

More recently, Notch signaling was also established as an essential factor in confining lymphatic differentiation to the dorso-lateral side of the anterior cardinal veins [24-26]. Targeted gene disruption of Notch1 in LEC precursors gave rise to a localized expansion of lymphatic specific markers in non-lymphangiogenic segments of the veins. Another growth factor, TGF-β was recently reported to perform a dual role during dermal lymphangiogenesis; whilst able to promote lymphatic vessels sprouting and branching complexity, TGF- β also inhibited LEC proliferation [27]. Finally, bone morphogenic proteins BMP2 and BMP9 were also reported as negative modulators of lymphatic vessel growth and differentiation. In vitro, BMP9-induced ALK1 was found to directly inhibit PROX1 activity; while loss-of-function of either Bmp9 or Alk1 lead to dysmorphic lymphatic vessels [28]. Further, BMP2 was reported to inhibit Prox1 via the induction of micro RNAs (miR-31 and miR181a) [29] to negatively regulate LEC identity.

Growth factors and lymphatic vascular remodelling

Embryonic animal models have shown that once the venous endothelium became committed to a LEC fate, the acquisition of lymphatic-specific markers such as podoplanin, LYVE1, VEGF-R3 and NRP2 enabled LECs to become responsive to growth factor stimulation (predominantly VEGF-C), in order to assemble a vascular plexus within the mesenchyme (from 10.5 dpc until 14.5 dpc) [30]. Later, this plexus was remodeled (14.5 dpc until 18.5 dpc) and finally matured to establish a functional lymphatic network. The major signaling axis controlling LEC migration and remodeling was controlled by the VEGF-C/Collagen and Calcium-Binding Epidermal growth factor domains 1 (CCBE1)/VEGF-R3 pathway -a signaling axis that promoted LEC migration, proliferation and survival, and was thus indispensible for both embryonic and adult lymphangiogenesis [31]. Vegf-c loss-of-function experiments revealed that, despite LEC specification still occurring in the absence of this key growth factor, the embryos lacked a complete lymphatic vasculature, due to defective lymphangiogenic sprouting. Further, a positive feedback loop between VEGF-R3 and PROX1 has been proposed to maintain the identity and the number of LEC progenitor in the cardinal vein [32].

Studying the maturation process of VEGF-C protein, Jeltsch et al. and Leguen et al. demonstrated a regulatory pathway of VEGF-C activation that varies depending on the activity of CCBE1 [33,34]. It was shown that CCBE1 could promote cleavage from a minimally-active form of VEGF-C to a mature and active form, via the A disintegrin and metalloprotease with thrombospondin motifs-3 (ADAMTS-3) protease [33]. This discovery not only suggested that CCBE1/ADAMTS-3-inhibition may represent an exciting novel therapeutic avenue to limit tumor lymphangiogenesis; but also further re-enforced the concept that tumor location (e.g. next to a source of CCBE1) may be a key determinant of the ability of normal lymphatics to sprout -even in the presence of established pro-lymphangiogenic cues -during both embryonic and pathological lymphangiogenesis [35].

Bridging the Gap between Embryonic and Tumor-Induced Lymphangiogenesis

Re-activation of embryonic pathways during tumor-induced lymphangiogenesis

As tumor lymphatics mimic normal initial lymphatic capillaries in their role as the entry point for fluid absorption and cellular escape into the lymphatic system, it is reasonable to draw parallels between major embryonic pathways that instruct lymphatic formation and morphogenesis and the pathways that become dysregulated and/or reactivated under tumor conditions (Figure 1B).

A key mechanism found to contribute to tumor lymphangiogenesis was outgrowth from pre-existing initial lymphatic vessels in and around the primary tumor mass in response to multiple stimuli, including VEGF-C, VEGF-D, PDGF-β, FGF2 and angiopoietin [35]. Similar to their functions during embryonic development, these growth factors were seen to trigger peri-tumoral lymphangiogenesis and increase metastasis in mice [36]. Each of these signaling molecules was reported to play key functions during embryonic angiogenesis and lymphangiogenesis, in both mouse and fish model systems [31].

Within the tumor microenvironment, multiple cellular sources of VEGF-C have been identified. Moussai et al. and Schoppmann et al. showed that tumor-infiltrating macrophages secrete this growth factor [37,38], while in cancer cells such as MCF7 human breast cancer cell line, SIX1 was also shown to promote VEGF-C gene expression, which stimulated cancer metastasis [39]. Further, LECs were demonstrated to act in an autocrine fashion in response to VEGF-C signaling, by up- regulating chemokines such as CCL21, which enhanced tumor chemo-invasion [40].

Other developmental lymphangiogenic pathways ‘re-awakened’ in cancer-related lymphangiogenesis include NRP2, a receptor for class III semaphorins and VEGF-C co-receptor [19] that can stimulate both VEGFR-2 and VEGFR-3 signaling [41]. Yuan et al. demonstrated that genetic disruption of Nrp2 results in reduced LEC proliferation and tissue lymphatic density during development [41]. In contrast to the situation found in developing lymphatics, Nrp2 is not expressed in quiescent lymphatics; however, Caunt et al. demonstrated not only that Nrp2 became expressed during active lymphangiogenesis in and around a primary tumor -but that blockade of NRP2 signaling restricted tumoral lymphangiogenesis [42].

LEC: Lymphatic Endothelial Cells; VEGF: Vascular Endothelial Growth Factor; TGF: Transforming Growth Factor; BMP: Bone Morphogenetic Protein; COUP-TFII: Chicken-Ovalbumin Upstream Transcription Factor II; PROX1: Prospero-Related Homeobox-1; SOX18: SRY-Related HMG Box containing Domain 18; CCBE1: Collagen and Calcium Binding EGF Domain: RA: Retinoic Acid; CCL21: Chemokine (C-C motif) Ligand 21; VEGF-R3: VEGF- Receptor3. ISV: Inter-Somitic Vessels; PGD: Prostaglandin-D

Fewer functional tumor lymphatics were generated in this tumor model, thereby contributing to reduced metastasis to lymph nodes and distant organs [42]. Similarly critical for lymphatic development, the ligand for endothelial Tie2 receptor tyrosine kinase angiopoietin 2 (Ang2; also known as Angpt2 or Agpt2) was shown to be specifically required for developmental lymphatic patterning [43]. It was subsequently demonstrated that angiopoietin expression induced in animal models additionally promoted peri-tumoral lymphangiogenesis, and that an ANG2-inhibitory antibody reduced tumor lymphangiogenesis and metastasis to both regional lymph nodes and the lungs [44,45]. Cao et al. used a mouse corneal model to investigate a developmental role for FGF2 in lymphangiogenesis [46]. They found that FGF2 activity was mediated by FGFR-1 expressed on LECS; however, that this lymphangiogenic pathway required synergistic VEGFR-3 activation by VEGF-C - particularly for tip cell initiation of lymphatic sprouting [46]. In the analagous process of lymphangiogenesis induced in a tumor model, the authors found that VEGFR-3-mediated lymphangiogenesis incorporated synergistic VEGF-C/FGF2-driven tumoral lymphatic formation, which was associated with tumor metastasis. Neutralising antibody against the common VEGFR-3 pathway was able to restrict lymphatic ingrowth and tumor spread [46].

Finally, numerous other genes have been linked with abnormal developmental remodeling or maturation, and altered resulting phenotypeslymphatics, without yet being implicated in tumor lymphangiogenesis. Examples include genes such as angiopoietin-like 4 (Angptl4)] [47] apoptosis stimulating protein of p53 (Aspp1; Ppp1r13b) [48] and T-synthase (C1galt1) [49]. Other abnormalities such as dysfunctional pericyte recruitment, impaired valve formation [50], atypical patterning or hypoplastic/hyperplasia have been linked to chylothorax and lymphoedematous phenotypes, some of which have been linked to analogous human syndromes (such as lymphoedema distichiasis [51-53]. Collectively, these pathways and are now considered potential targets for molecular therapeutics to restrict tumor-induced lymphangiogenesis [36]; while restricting the negative regulators of these pathways may facilitate therapeutic lymphangiogenesis in patients suffering from secondary lymphedema following lymph node surgery [54].

The cellular origin of LECs in tumor lymphatics

The utilisation of fate mapping experiments during embryonic development has established, at least in the early steps of lymphangiogenesis that LEC precursors arise from multiple venous vascular beds [55]. Wilting et al. showed in an a vein limb bud grafting experiment that homotopically grafted distal wing buds of chick into quail embryos formed lymphatics composed of both chick and quail endothelial cells; suggesting that the lymphatics of the wing bud do not exclusively develop from sprouts from nearby lymph sacs, but also involve recruitment of local so-called ‘lymphangioblasts’ [55].

In the tumor micro-environment several additional factors such as inflammation may alter the manifestation of these developmental processes. Tumor-induced neo- lymphangiogenesis is thought to originate predominantly from outgrowths from pre-existing lymphatic vessels. However, it still remains to be established whether cell- autonomous contributions to neo-lymphatic formation in the tumor setting are complemented by trans-differentiation from other cell types [56].

A potential non-endothelial LEC-progenitor proposed in the literature are macrophages or circulating endothelial progenitor [57,58]. The role of tumor- associated macrophages (TAMs) in promoting angiogenesis and acquiring an angiogenic phenotype has been well established [59]; but it was not until recently that TAM were suggested as a potential source of LECs. This hypothesis derived from observations that TAMs -which expressed integrin family member cluster of differentiation molecule 11b (Cd11b) -acquire lymphatic specific markers including Prox1, Lyve1, Podoplanin and Vegf-r3; meanwhile undergoing an observable down-regulation of myeloid markers within the same cells [60]. These TAMs were able to integrate into growing lymphatic vessels in an experimental mouse tumor model [60]. This partial, transient reprogramming of the myeloid identity into a LEC profile suggested that TAMs may contribute to the lymphatic vasculature [61]. Conversely, studies in developing mouse embryos suggested that macrophages solely contributed to modulation of lymphatic vessel caliber by regulating LEC proliferation, but that they did not trans-differentiate to integrate into forming vessels or act as a cellular reservoir for LEC-precursors [62]. More detailed fate mapping experiments in the tumor setting and advanced in vivo live imaging will enable researchers to answer definitely whether or not TAM trans-differentiate into LECs, or if they merely act as intermediate cell clusters to bridge newly-formed intra- and peri-tumoral lymphatics.

A key feature of lymphangiogenesis during embryonic development is the trans- differentiation of venous endothelial cells into LECs [14]. Despite advances our in understanding of the genetic pathways that govern LEC differentiation, studies have yet to explore the hypothesis that a subset of neo-lymphatics could also arise from pre-existing veins or other blood vascular structures in the adult. Several studies support this concept: COUP-TFII and SOX18 were each shown to be individually required for tumor-induced lymphangiogenesis [63,64]. Further, VEGF-R3 expression has been shown to become reactivated in a subset of the blood vessels [65-67]. Future studies based on fate mapping experiments are required to identify the cellular origin of LEC progenitors in a tumor setting.

Remodeling of Pre-Existing Adult Lymphatic Vessels in Solid Tumors

Peripheral and central tumoral lymphatic sprouting/remodelling

Whilst originally considered a passive conduit, the lymphatic system has been more recently acknowledged as an active, dynamic participant in cancer metastasis [68]. Further, the ways in which the individual vessel subtypes within the hierarchical lymphatic network respond to, and interact with, external cues have also become recognized as critical to cancer spread [35]. Neo-lymphatics formed within the primary tumor have been the focus of much animal and human research, and have been shown to be associated with enhanced rates of metastasis [69]. The location within the tumor (whether central or peripheral) in which the lymphatics form, has also aroused interest; both in terms of what role vessels in each location might play in metastasis and what the different lymphangiogenic mechanisms favoring neo- lymphatics in each location might be [35].

In normal physiology, interstitial fluid bathing the extravascular (interstitial) tissues cycles via the lymphatics back into the blood vascular circulation. Absorbed by thin- walled ‘initial’ or capillary lymphatics, lymph is transported via progressively enlarging vessels that also adopt a more developed mural structure, consisting of muscular and adventitial layers surrounding LEC-lined lumens. They exhibit both valves and a contraction system that aid lymph movement against gravity [69-71]. For the most part, tumor-induced lymphatics have been shown to resemble a disorganized version of initial lymphatics [72] both mimicking the absorptive function as an entry point to the lymphatic network and sharing structural similarities with their parent vessel of origin [35]. Tumor lymphatics derived from nearby normal initial lymphatics sprout then undergo directional in-growth toward tumor-derived lymphangiogenic cues reminiscent of those seen during development. Further, normal lymphatics surrounding the tumor undergo remodeling in the form of dilatation and increased caliber in response to VEGF-D [72,73] and VEGF-C [74], potentially augmenting the flow into the collecting lymphatic vessels and onward to the draining lymph nodes (Figure 2); a feature associated with enhanced metastasis [68,75].

cellular-immunology-smooth-muscle

Figure 2: The remodeling of lymphatic vessels influences tumor metastasis. The proximity of the primary tumor to small, plastic and hyperdense initial or capillary vessels is a key parameter that modulates metastatic events. In small vessels the lymphatic endothelium is only anchored to the matrix by filaments and zipper-like cell-cell junctions that favor the entry of tumor cells. Conversely larger lymphatic vessels surrounded by a layer of smooth-muscle cells and adventitia are less prone to sprouting, however they remodel via alternative prostaglandin mediated mechanisms. The combination of various components from the tumor micro-environment, such as fluid pressure, chemokines and growth factor signaling stimulate neo-lymphangiogenesis and remotely prepare a lymphovascular niche in the tumor draining lymph nodes.

Remodeling of collecting lymphatics

The small caliber initial lymphatics and larger collecting vessels that make up the lymphatic vasculature differ in their morphology, structure and anatomical locations; and perform distinct specialized functions that contribute toward their respective normal physiological roles [76,77]. Most research investigating mechanisms behind lymphogenous metastasis of cancer cells has focused on elucidating the influences of lymphangiogenic growth factors on the initial lymphatic vessel subtypes within or around a primary tumor, whilst the collecting lymphatics that drain tumor tissues towards regional lymph nodes, have remained largely ignored [78]. Whereas ambient smaller lymphatics largely respond to lymphangiogenic factors by proliferation and/or sprouting to generate tumor-associated lymphangiogenesis (Figure 2), the larger collecting lymphatics respond to the same lymphangiogenic stimuli in a unique and quite distinct manner [35,75,79,80]. Far from passive conduits of metastatic cells, more recent findings suggest that collecting vessels undergo a significant remodeling of their own, which contributes critically to the process of tumor spread [68,75] (Figure 2). Observations of murine models of VEGF-C-over-expressing metastatic tumors demonstrated that drainage from the primary tumor via the collecting lymphatics performs an active role in enhancing tumor dissemination, through increasing fluid flow by dilating-an increase in collecting lymphatic diameter attributed to LEC proliferation [75,79,80]. Similarly, it was shown that VEGF-D secreted by a flank xenograft tumor model also induced dilatation of the collecting lymphatics draining from the primary tumor to the axilla that was critical for cancer spread to the sentinel lymph node [68]. In contrast to the VEGF-C model and to the effect that VEGF-D had on initial lymphatics [72]; however, the mechanism by which circumferential dilation occurs is not by endothelial proliferation but through specific prostaglandin-mediated responses to VEGF-D [68]. Importantly, treatment with non- steroidal anti-inflammatory drugs not only reversed the VEGF-D-driven morphological remodeling of collection lymphatics, but also reduced the rate of tumor metastasis to draining lymph nodes and distant organs [68].

The influence of tissue forces on lymphatic remodelling

In addition to growth factors, pro-lymphangiogenic cues may also include mechano-biological stimuli such as hydrostatic pressure and tissue flow, which then become translated into lymphangiogenic molecular pathways [71,81]. Whilst the importance of these forces in generating tumor lymphatics is less well understood, emerging work indicates that mechano-induction may also be an important additional determinant of the density and nature of lymphatics generated within different regions of the tumor [71,81]. Both mechano-induction and lymphangiogenic signaling pathways are process ‘borrowed’ from embryogenesis. During development, mechanosensory complexes formed by Integrin ß1 and VEGFR-3 can translate increased interstitial/ECM stiffness into lymphangiogenic signals [70,71]. Increased interstitial fluid pressure leads to activation of Integrin ß1, which in turn induces VEGFR-3 tyrosine phosphorylation [82]. This VEGFR-3 activation is thought to be mediated by Src family kinases (SFKs) in a VEGF-C-independent pathway [83], thus resulting in VEGF-C-independent LEC proliferation [82]. Further, fibrillin anchoring-filaments were shown to be capable of binding transmembrane integrin glycoproteins to activate intracellular signaling pathways [77,84]. These signals translate ECM stiffening into cytoskeletal alterations that increase cell permeability in order to facilitate improved fluid uptake and flow [77,84]. In a tumor, higher fluid pressure within the centre or increased lymph flow toward the periphery may also determine the nature of lymphatics generated in different regions. High intra-tumoral pressure due to leaky nascent lymphatics and blood vessels may both induce intra-tumoral vessels to grow and lead to their collapse (due to poorly developed supporting structures) and dysfunction, contributing to further fluid accumulation [85]. This accumulation of non-absorbed fluid in turn produces a pressure gradient favoring flow towards the lower pressure tumor periphery, where the vessels remain functional due to lower ambient pressure and VEGF-C/D-mediated dilatation [73] (Figure 2). Whilst a specific role for hydrostatic pressure in driving tumor lymphangiogenesis remains to be defined in cancer, fluid flow gradients have been shown to drive tumor lymphangiogenesis [86].

Regardless of the underlying stimulus, however, tumors that recreate developmental lymphangiogenesis can enhance metastasis. The respective contributions of newly- formed vessels and the pre-formed mature lymphatics to the metastatic process has been debated [69]. The role/functionality of the nascent lymphatics within the substance of the tumor, and the degree of vessel dilation/collapse compared with those induced in peripheral areas of the tumor (or immediately adjacent tissues), remains controversial [87,88]. Intratumoral lymphatics were shown to be present in a murine model, yet predominanly collapsed -in contrast to the apparently dilated peri-tumoral lymphatics [87]. Analysis of human tumors showed that peri-tumoral lymphatics were the most important for metastasis; while intra-tumoral lymphatics exhibit proliferative markers [88], these vessels were collapsed and unable to transport tumor cells, despite a higher vessels density [42,75,89]. Whether peri-tumoral lymphatics represented pre-existing vessels compressed by an expanding tumor, or ‘neo-lymphatics’ generated through lymphangiogenesis remains unclear [89,90]. Authors postulate that there is both a greater concentration of stromal and inflammatory cells secreting VEGF-family members, and a greater density of pre-existing vessels to provide a ‘source’ from which tumor neo-lymphatics can sprout, in the periperal tumor microenvironment [91-93]. Studies of human melanomas found that peri-tumoral vessel density and caliber were significantly increased in metastatic lesions and were associated with regional metastasis, poor disease-free and overall patient survival [91,94,95]. A more recent study matching melanoma samples for all other prognostic indicators found that a high ratio of peri-tumoral-to-intra-tumoral lymphatic vessel density was associated with a higher rate of metastasis to the draining lymph node basin [35]. Human breast cancer specimens also exhibited collapsed intra-tumoral vessels (poorly-staining with proliferation markers) and increased densities of peri- tumoral lymphatics, that contained tumor emboli [88,96].

Lymphangiogenesis in the sentinel lymph node

Comparatively little is known about the development of lymph nodes or other lymphoid tissue. During development, these structures are situated along collecting lymphatics within a lymph node ‘anlagen’ or precursor tissue, at the site of future lymph nodes. Lymph nodes originate from connective tissues protruding into primative lymph sacs, and integrate lymphatics and blood vessels with haemopoetic cells and stromal supportive cells [97,98]. This process incorporates mesenchymal cells differentiation into aggregates of CD45+CD4+CD3+ lymphoid ‘tissue inducer cells’ and stromal organiser cells [98]. These distinct cell types interact within the anlagen to stimulate adhesion molecule expression on stromal organiser cells, and release chemokines such as CCL19, CCL21 and CXCL13; signals which, in turn, attract additional lymphoid ‘tissue-inducer cells’ and other haemopoietic cells [98,99]. Importantly, mouse models showed that lymph node development is genetically independent from lymphatic vascular embryogenesis [100].

A lymph node and distant organ metastatic niche

Lymph nodes represent a key ‘staging-post’ in tumor dissemination from the primary cancer toward distant organs. The pro-metastatic effects of tumor-derived lymphangiogenic growth factors are not limited only to the primary tumor microenvironment and may also induce lymphangiogenesis and modulate existing lymphatics and blood vessels within draining lymph nodes ‘downstream’ of the tumor; often prior to the arrival of metastatic cells [101,102]. Harrell et al. observed that lymphangiogenesis and enhanced lymph flow preceded melanoma metastasis, and was associated with increased levels of β-lymphocytes within the nodes [103]. These observations are consistent with the ‘seed’ and ‘soil’ theory of metastasis articulated by Paget in 1889. He suggested that primary cancer cells can spread preferentially to specific distant areas. Thus, increase lymphatic vessel density in a sentinel lymph node prior to any detectable metastasis could promote enhanced tumor transport to the lymph nodes and could serve an accurate predictor of lymphogenous spread [104] (Figure 2). The enhancement of the lymphatic network within these draining nodes is referred as the ‘lymphvascular niche’, which is akin to the ‘vascular niche’ seen in the formation and maintenance of hemopoietic stem cells within lymph nodes and bone marrow [104,105]. The lymphvascular niche may be able to tailor the microenvironment and, therefore, immune responses against cancer cells transiting toward the lymph node. Additionally, this niche may influence the conditions that preempt then support the survival and growth of metastatic deposits [106].

Future Approaches to Manipulating Cancer Lymphatics

Most of the novel anti-lymphangiogenic strategies to restrict tumor metastasis revolve around the VEGF-C/VEGF-R3 signaling axis, with the view to complementing anti-angiogenic therapies. Data from pre-clinical models suggest a clear benefit in targeting tumor-induced lymphangiogenesis to reduce solid tumor metastasis [107]. The next step is now to assess the clinical outcome of targeting the lymphatic vasculature in human cancer. Some anti-angiogenic molecules that target pathways involving the VEGF-family of growth factors also restrict lymphangiogenesis via the inhibition of the VEGFR-3 pathway; however, specific benefit of anti-lymphangiogenic therapies in this setting needs to be assessed more thoroughly, aside from evaluating tumor burden in the draining lymph nodes. In rare lymphatic vascular disorders, it is possible to manipulate LEC proliferation; as demonstrated by using Sildenafil (Viagra) treatment to restrict the growth of pediatric orbital lymphangioma [108]. So far, only two active phase I clinical trials targeting the VEGF-C/VEGF-R3 pathway have been initiated. The first combined Bevacizumab (avastin, anti-VEGF-A) with VGX-100 (anti-VEGF-C blocking antibody, NCT01514123) and another targeted VEGF-R3 (NCT01288989) in advanced solid adult malignancy, however, the outcomes are still pending.

Most of the novel anti-lymphangiogenic strategies to restrict tumor metastasis revolve around the VEGF-C/VEGF-R3 signaling axis, with the view to complementing anti-angiogenic therapies. Data from pre-clinical models suggest a clear benefit in targeting tumor-induced lymphangiogenesis to reduce solid tumor metastasis [107]. The next step is now to assess the clinical outcome of targeting the lymphatic vasculature in human cancer. Some anti-angiogenic molecules that target pathways involving the VEGF-family of growth factors also restrict lymphangiogenesis via the inhibition of the VEGFR-3 pathway; however, specific benefit of anti-lymphangiogenic therapies in this setting needs to be assessed more thoroughly, aside from evaluating tumor burden in the draining lymph nodes. In rare lymphatic vascular disorders, it is possible to manipulate LEC proliferation; as demonstrated by using Sildenafil (Viagra) treatment to restrict the growth of pediatric orbital lymphangioma [108]. So far, only two active phase I clinical trials targeting the VEGF-C/VEGF-R3 pathway have been initiated. The first combined Bevacizumab (avastin, anti-VEGF-A) with VGX-100 (anti-VEGF-C blocking antibody, NCT01514123) and another targeted VEGF-R3 (NCT01288989) in advanced solid adult malignancy, however, the outcomes are still pending. Despite major advances in our understanding of the molecular pathways involved in VEGF biology, the development of anti-VEGF drugs in human cancer has already started to show some limitations. Long-term clinical benefit of VEGF-based therapies has remained marginal due to resistance mechanisms mediated by a combination of both VEGF-axis-dependent alterations and non-VEGF-dependent pathways [109]. In some cases, anti-angio/lymphangiogenic approaches even led to the opposite effect and enhanced tumor metastasis [110]. To complement anti-VEGF-A strategies, other monoclonal antibody inhibitors have been developed against targets such as angiopoeitin signaling antagonists [111,112]. In order to avoid developing resistance phenomena similar to limitations as ‘Avastin-escape’ seen in anti-VEGF-A therapies, it will be essential to further explore the potential of VEGF-C- and VEGF- D-independent strategies. Additionally, immuno-modulatory regulating effects of the VEGF protein growth factor family is an aspect of their biology that remains relatively neglected in designing anti-angiogenic therapeutic strategies; for example VEGF-A was shown to promote priming of natural killer cells in mouse tumor models [113,114], while VEGF-C was reported to enhance immunity in a B16 melanoma model [115]. Further, manipulating the remodeling of the lymphatic network may also affect immune cell trafficking and therefore influence the inflammatory micro- environment of the tumor. These findings have led to the reassessment of the approaches used to target the VEGF family of proteins in the tumor setting.

Lastly, the heterogeneity of the both the blood and lymphatic tumor vasculature may also contribute to limiting the success of the anti-VEGF therapies, as only a subset of the remodeling vasculature is affected by the blockade of VEGF signaling [116]. Further, the broadening of the therapeutic focus from simply targeting newly-sprouting tumor lymphatics and nearby initial lymphatics under the influence of the tumor microenvironment, to also encompass the collecting lymphatics may provide further avenues for therapeutic intervention. A more detailed understanding of the biological characteristics of the different lymphatic subtypes and their role in metastasis is thus, fundamental. An example of detailed mechanistic understanding leading to novel therapeutic approaches is the elucidation of the role of PGE2 in metastasis resulting from VEGF-D-mediated down-regulation of the enzyme prostaglandin dehydrogenase by Karnezis et al.; and the inhibition of metastasis by using non-steroidal anti-inflammatory drugs [68]. Several subsequent clinical studies have corroborated an anti-metastatic benefit to breast cancer patients treated with Aspirin [117,118].

Overall, therefore, it is critical to further explore the developmental pathways that govern lymphangiogenesis in order to facilitate the identification, characterisation and optimisation of molecular targets in tumor lymphangiogenesis, and to provide new directions in treatments to restrict the spread of cancer.

References

  1. Kampmeier OF (1912) The development of the thoracic duct in the pig. Am J Anat 13: 401-475.
  2. Sabin FR (1902) On the origin of the lymphatic system from the veins, and the development of the lymph hearts and thoracic duct in the pig. Am J Anat 1:367-389.
  3. Mandriota SJ,Jussila L, Jeltsch M, Compagni A, Baetens D, et al. (2001) Vascular endothelial growth factor-C-mediated lymphangiogenesis promotes tumour metastasis. EMBO J 20: 672-682.
  4. Skobe M,Hawighorst T, Jackson DG, Prevo R, Janes L, et al. (2001) Induction of tumor lymphangiogenesis by VEGF-C promotes breast cancer metastasis. Nat Med 7: 192-198.
  5. Stacker SA, Caesar C, Baldwin ME, Thornton GE, Williams RA, et al. (2001) VEGF-D promotes the metastatic spread of tumor cells via the lymphatics. Nat Med 7: 186-191.
  6. Christiansen A,Detmar M (2011) Lymphangiogenesis and cancer. Genes Cancer 2: 1146-1158.
  7. Wigle JT, Oliver G (1999) Prox1 function is required for the development of the murine lymphatic system. Cell 98: 769-778.
  8. Yaniv K,Isogai S, Castranova D, Dye L, Hitomi J, et al. (2006) Live imaging of lymphatic development in the zebrafish. Nat Med 12: 711-716.
  9. François M, Short K, Secker GA, Combes A, Schwarz Q, et al. (2012) Segmental territories along the cardinal veins generate lymph sacs via a ballooning mechanism during embryonic lymphangiogenesis in mice. DevBiol 364: 89-98.
  10. Hägerling R,Pollmann C, Andreas M, Schmidt C, Nurmi H, et al. (2013) A novel multistep mechanism for initial lymphangiogenesis in mouse embryos based on ultramicroscopy. EMBO J 32: 629-644.
  11. Srinivasan RS, Dillard ME, Lagutin OV, Lin FJ, Tsai S, et al. (2007) Lineage tracing demonstrates the venous origin of the mammalian lymphatic vasculature. Genes Dev 21: 2422-2432.
  12. Yang Y,García-Verdugo JM, Soriano-Navarro M, Srinivasan RS, Scallan JP, et al. (2012) Lymphatic endothelial progenitors bud from the cardinal vein and intersomitic vessels in mammalian embryos. Blood 120: 2340-2348.
  13. Tammela T,Alitalo K (2010) Lymphangiogenesis: Molecular mechanisms and future promise. Cell 140: 460-476.
  14. François M,Caprini A, Hosking B, Orsenigo F, Wilhelm D, et al. (2008) Sox18 induces development of the lymphatic vasculature in mice. Nature 456: 643-647.
  15. Wigle JT, Harvey N, Detmar M, Lagutina I, Grosveld G, et al. (2002) An essential role for Prox1 in the induction of the lymphatic endothelial cell phenotype. EMBO J 21: 1505-1513.
  16. Deng Y, Simons M (2013) Lymphatic fate determination: playing RAF with ERK. Cell Cycle 12: 1157-1158.
  17. Srinivasan RS,Geng X, Yang Y, Wang Y, Mukatira S, et al. (2010) The nuclear hormone receptor Coup-TFII is required for the initiation and early maintenance of Prox1 expression in lymphatic endothelial cells. Genes Dev 24: 696-707.
  18. Yamazaki T,Yoshimatsu Y, Morishita Y, Miyazono K, Watabe T (2009) COUP-TFII regulates the functions of Prox1 in lymphatic endothelial cells through direct interaction. Genes Cells 14: 425-434.
  19. Xu Y, Yuan L, Mak J, Pardanaud L, Caunt M, et al. (2010) Neuropilin-2 mediates VEGF-C-induced lymphatic sprouting together with VEGFR3. J Cell Biol 188: 115-130.
  20. Srinivasan RS, Oliver G (2011) Prox1 dosage controls the number of lymphatic endothelial cell progenitors and the formation of the lymphovenous valves. Genes Dev 25: 2187-2197.
  21. Johnson NC, Dillard ME, Baluk P, McDonald DM, Harvey NL, et al. (2008) Lymphatic endothelial cell identity is reversible and its maintenance requires Prox1 activity. Genes Dev 22: 3282-3291.
  22. Bowles J, Secker G, Nguyen C,Kazenwadel J, Truong V, et al. (2014) Control of retinoid levels by CYP26B1 is important for lymphatic vascular development in the mouse embryo. DevBiol 386: 25-33.
  23. Marino D,Dabouras V, Brändli AW, Detmar M (2011) A role for all-trans-retinoic acid in the early steps of lymphatic vasculature development. J Vasc Res 48: 236-251.
  24. Fatima A, Culver A, Culver F, Liu T, Dietz WH, et al. (2014) Murine Notch1 is required for lymphatic vascular morphogenesis during development. DevDyn 243: 957-964.
  25. Geudens I,Herpers R, Hermans K, Segura I, Ruiz de Almodovar C, et al. (2010) Role of delta-like-4/Notch in the formation and wiring of the lymphatic network in zebrafish. ArteriosclerThrombVascBiol 30: 1695-1702.
  26. Murtomaki A, Uh MK, Choi YK, Kitajewski C, Borisenko V, et al. (2013) Notch1 functions as a negative regulator of lymphatic endothelial cell differentiation in the venous endothelium. Development 140: 2365-2376.
  27. James JM,Nalbandian A, Mukouyama YS (2013) TGFβ signaling is required for sprouting lymphangiogenesis during lymphatic network development in the skin. Development 140: 3903-3914.
  28. Yoshimatsu Y, Lee YG, Akatsu Y, Taguchi L, Suzuki HI, et al. (2013) Bone morphogenetic protein-9 inhibits lymphatic vessel formation via activin receptor-like kinase 1 during development and cancer progression. ProcNatlAcadSci U S A 110: 18940-18945.
  29. Dunworth WP, Cardona-Costa J, Bozkulak EC, Kim JD, Meadows S, et al. (2014) Bone morphogenetic protein 2 signaling negatively modulates lymphatic development in vertebrate embryos. Circ Res114: 56-66.
  30. Karkkainen MJ,Haiko P, Sainio K, Partanen J, Taipale J, et al. (2004) Vascular endothelial growth factor C is required for sprouting of the first lymphatic vessels from embryonic veins. Nat Immunol 5: 74-80.
  31. Koltowska K,Betterman KL, Harvey NL, Hogan BM (2013) Getting out and about: the emergence and morphogenesis of the vertebrate lymphatic vasculature. Development 140: 1857-1870.
  32. Srinivasan RS, Escobedo N, Yang Y,Interiano A, Dillard ME, et al. (2014) The Prox1-Vegfr3 feedback loop maintains the identity and the number of lymphatic endothelial cell progenitors. Genes Dev 28: 2175-2187.
  33. Jeltsch M,Jha SK, Tvorogov D, Anisimov A, Leppänen VM, et al. (2014) CCBE1 enhances lymphangiogenesis via A disintegrin and metalloprotease with thrombospondin motifs-3-mediated vascular endothelial growth factor-C activation. Circulation 129: 1962-1971.
  34. Le Guen L,Karpanen T, Schulte D, Harris NC, Koltowska K, et al. (2014) Ccbe1 regulates Vegfc-mediated induction of Vegfr3 signaling during embryonic lymphangiogenesis. Development 141: 1239-1249.
  35. Shayan R,Inder R, Karnezis T, Caesar C, Paavonen K, et al. (2013) Tumor location and nature of lymphatic vessels are key determinants of cancer metastasis. ClinExp Metastasis 30: 345-356.
  36. Stacker SA, Williams SP,Karnezis T,Shayan R, Fox SB, et al. (2014) Lymphangiogenesis and lymphatic vessel remodelling in cancer. Nat Rev Cancer 14: 159-172.
  37. Moussai D, Mitsui H, Pettersen JS, Pierson KC, Shah KR, et al. (2011) The human cutaneous squamous cell carcinoma microenvironment is characterized by increased lymphatic density and enhanced expression of macrophage-derived VEGF-C. J Invest Dermatol 131: 229-236.
  38. Schoppmann SF, Birner P, Stöck J, Kalt R, Ullrich R, et al. (2002) Tumor-associated macrophages express lymphatic endothelial growth factors and are related to peritumoral lymphangiogenesis. Am J Pathol 161: 947-956.
  39. Wang CA,Jedlicka P, Patrick AN, Micalizzi DS, Lemmer KC, et al. (2012) SIX1 induces lymphangiogenesis and metastasis via upregulation of VEGF-C in mouse models of breast cancer. J Clin Invest 122: 1895-1906.
  40. Issa A, Le TX, Shoushtari AN, Shields JD, Swartz MA (2009) Vascular endothelial growth factor-C and C-C chemokine receptor 7 in tumor cell-lymphatic cross-talk promote invasive phenotype. Cancer Res 69: 349-357.
  41. Yuan L,Moyon D, Pardanaud L, Bréant C, Karkkainen MJ, et al. (2002) Abnormal lymphatic vessel development in neuropilin 2 mutant mice. Development 129: 4797-4806.
  42. Caunt M,Mak J, Liang WC, Stawicki S, Pan Q, et al. (2008) Blocking neuropilin-2 function inhibits tumor cell metastasis. Cancer Cell 13: 331-342.
  43. Gale NW, Thurston G, Hackett SF, Renard R, Wang Q, et al. (2002) Angiopoietin-2 is required for postnatal angiogenesis and lymphatic patterning, and only the latter role is rescued by Angiopoietin-1. Dev Cell 3: 411-423.
  44. Fagiani E, Lorentz P, Kopfstein L, Christofori G (2011) Angiopoietin-1 and -2 exert antagonistic functions in tumor angiogenesis, yet both induce lymphangiogenesis. Cancer Res 71: 5717-5727.
  45. Holopainen T,Saharinen P, D'Amico G, Lampinen A, Eklund L, et al. (2012) Effects of angiopoietin-2-blocking antibody on endothelial cell-cell junctions and lung metastasis. J Natl Cancer Inst 104: 461-475.
  46. Cao R,Ji H, Feng N, Zhang Y, Yang X, et al. (2012) Collaborative interplay between FGF-2 and VEGF-C promotes lymphangiogenesis and metastasis. ProcNatlAcadSci U S A 109: 15894-15899.
  47. Bäckhed F, Crawford PA, O'Donnell D, Gordon JI (2007) Postnatal lymphatic partitioning from the blood vasculature in the small intestine requires fasting-induced adipose factor. ProcNatlAcadSci U S A 104: 606-611.
  48. Hirashima M, Sano K, Morisada T, Murakami K, Rossant J, et al. (2008) Lymphatic vessel assembly is impaired in Aspp1-deficient mouse embryos. DevBiol 316: 149-159.
  49. Fu J, Gerhardt H, McDaniel JM, Xia B, Liu X, et al. (2008) Endothelial cell O-glycan deficiency causes blood/lymphatic misconnections and consequent fatty liver disease in mice. J Clin Invest 118: 3725-3737.
  50. Pepper MS,Skobe M (2003) Lymphatic endothelium: morphological, molecular and functional properties. J Cell Biol 163: 209-213.
  51. Brooks BP,Dagenais SL, Nelson CC, Glynn MW, Caulder MS, et al. (2003) Mutation of the FOXC2 gene in familial distichiasis. J AAPOS 7: 354-357.
  52. Dagenais SL,Hartsough RL, Erickson RP, Witte MH, Butler MG, et al. (2004) Foxc2 is expressed in developing lymphatic vessels and other tissues associated with lymphedema-distichiasis syndrome. Gene Expr Patterns 4: 611-619.
  53. Kriederman BM,Myloyde TL, Witte MH, Dagenais SL, Witte CL, et al. (2003) FOXC2 haploinsufficient mice are a model for human autosomal dominant lymphedema-distichiasis syndrome. Hum Mol Genet 12: 1179-1185.
  54. DiSipio T, Rye S, Newman B, Hayes S (2013) Incidence of unilateral arm lymphoedema after breast cancer: a systematic review and meta-analysis. Lancet Oncol 14: 500-515.
  55. Wilting J,Neeff H, Christ B (1999) Embryonic lymphangiogenesis. Cell Tissue Res 297: 1-11.
  56. Battula VL, Evans KW, Hollier BG, Shi Y, Marini FC, et al. (2010) Epithelial-mesenchymal transition-derived cells exhibit multilineage differentiation potential similar to mesenchymal stem cells. Stem Cells 28: 1435-1445.
  57. Religa P, Cao R, Bjorndahl M, Zhou Z, Zhu Z, et al. (2005) Presence of bone marrow-derived circulating progenitor endothelial cells in the newly formed lymphatic vessels. Blood 106: 4184-4190.
  58. Salven P,Mustjoki S, Alitalo R, Alitalo K, Rafii S (2003) VEGFR-3 and CD133 identify a population of CD34+ lymphatic/vascular endothelial precursor cells. Blood 101: 168-172.
  59. Lamagna C,Aurrand-Lions M, Imhof BA (2006) Dual role of macrophages in tumor growth and angiogenesis. J LeukocBiol 80: 705-713.
  60. Hall KL, Volk-Draper LD, Flister MJ, Ran S (2012) New model of macrophage acquisition of the lymphatic endothelial phenotype. PLoS One 7: e31794.
  61. Zumsteg A,Baeriswyl V, Imaizumi N, Schwendener R, Rüegg C, et al. (2009) Myeloid cells contribute to tumor lymphangiogenesis. PLoS One 4: e7067.
  62. Gordon EJ,Rao S, Pollard JW, Nutt SL, Lang RA, et al. (2010) Macrophages define dermal lymphatic vessel calibre during development by regulating lymphatic endothelial cell proliferation. Development 137: 3899-3910.
  63. Duong T,Proulx ST, Luciani P, Leroux JC, Detmar M, et al. (2012) Genetic ablation of SOX18 function suppresses tumor lymphangiogenesis and metastasis of melanoma in mice. Cancer Res 72: 3105-3114.
  64. Lin J, Lalani AS, Harding TC, Gonzalez M, Wu WW et al. (2005) Inhibition of lymphogenous metastasis using adeno-associated virus-mediated gene transfer of a soluble VEGFR-3 decoy receptor. Cancer Res 65:6901-6909.
  65. Paavonen K,Puolakkainen P, Jussila L, Jahkola T, Alitalo K (2000) Vascular endothelial growth factor receptor-3 in lymphangiogenesis in wound healing. Am J Pathol 156: 1499-1504.
  66. Partanen TA,Alitalo K, Miettinen M (1999) Lack of lymphatic vascular specificity of vascular endothelial growth factor receptor 3 in 185 vascular tumors. Cancer 86: 2406-2412.
  67. Valtola R,Salven P, Heikkilä P, Taipale J, Joensuu H, et al. (1999) VEGFR-3 and its ligand VEGF-C are associated with angiogenesis in breast cancer. Am J Pathol 154: 1381-1390.
  68. Karnezis T,Shayan R, Caesar C, Roufail S, Harris NC, et al. (2012) VEGF-D promotes tumor metastasis by regulating prostaglandins produced by the collecting lymphatic endothelium. Cancer Cell 21: 181-195.
  69. Shayan R,Achen MG, Stacker SA (2006) Lymphatic vessels in cancer metastasis: bridging the gaps. Carcinogenesis 27: 1729-1738.
  70. Leak LV, Burke JF (1968), Electron microscopic study of lymphatic capillaries in the removal of connective tissue fluids and particulate substances. Lymphology 1: 39-52.
  71. Sabine A,Agalarov Y, Maby-El Hajjami H, Jaquet M, Hägerling R, et al. (2012) Mechanotransduction, PROX, and FOXC2 cooperate to control connexin37 and calcineurin during lymphatic-valve formation. Dev Cell 22: 430-445.
  72. Shayan R,Karnezis T, Tsantikos E, Williams SP, Runting AS, et al. (2007) A system for quantifying the patterning of the lymphatic vasculature. Growth Factors 25: 417-425.
  73. Paquet-Fifield S, Levy SM, Sato T, Shayan R, Karnezis T, et al. (2013)vascular endothelial growth factor-d modulates caliber and function of initial lymphatics in the dermis. J Invest Dermatol 133: 2074-2084.
  74. Baluk P,Tammela T, Ator E, Lyubynska N, Achen MG, et al. (2005) Pathogenesis of persistent lymphatic vessel hyperplasia in chronic airway inflammation. J Clin Invest 115: 247-257.
  75. He Y,Rajantie I, Pajusola K, Jeltsch M, Holopainen T, et al. (2005) Vascular endothelial cell growth factor receptor 3-mediated activation of lymphatic endothelium is crucial for tumor cell entry and spread via lymphatic vessels. Cancer Res 65: 4739-4746.
  76. Mäkinen T, Adams RH, Bailey J, Lu Q, Ziemiecki A, et al. (2005) PDZ interaction site in ephrinB2 is required for the remodeling of lymphatic vasculature. Genes Dev 19: 397-410.
  77. Scavelli C, Weber E, Aglianò M, Cirulli T, Nico B, et al. (2004) Lymphatics at the crossroads of angiogenesis and lymphangiogenesis. J Anat 204: 433-449.
  78. Wick N, Haluza D, Gurnhofer E, Raab I, Kasimir MT, et al. (2008) Lymphatic precollectors contain a novel, specialized subpopulation of podoplanin low, CCL27-expressing lymphatic endothelial cells. Am J Pathol173: 1202-1209.
  79. Hoshida T, Isaka N, Hagendoorn J, di Tomaso E, Chen YL, et al. (2006) Imaging steps of lymphatic metastasis reveals that vascular endothelial growth factor-C increases metastasis by increasing delivery of cancer cells to lymph nodes: therapeutic implications. Cancer Res 66: 8065-8075.
  80. Wirzenius M,Tammela T, Uutela M, He Y, Odorisio T, et al. (2007) Distinct vascular endothelial growth factor signals for lymphatic vessel enlargement and sprouting. J Exp Med 204: 1431-1440.
  81. Planas-Paz L,Lammert E (2013) Mechanical forces in lymphatic vascular development and disease. Cell Mol Life Sci 70: 4341-4354.
  82. Planas-Paz L,Lammert E (2014) Mechanosensing in developing lymphatic vessels. AdvAnatEmbryol Cell Biol 214: 23-40.
  83. Galvagni F, Pennacchini S, Salameh A, Rocchigiani M, Neri F, et al. (2010) Endothelial cell adhesion to the extracellular matrix induces c-Src-dependent VEGFR-3 phosphorylation without the activation of the receptor intrinsic kinase activity. Circ Res106: 1839-1848.
  84. Solito R, Alessandrini C, Fruschelli M, Pucci AM, Gerli R (1997) An immunological correlation between the anchoring filaments of initial lymph vessels and the neighboring elastic fibers: a unified morphofunctional concept. Lymphology30: 194-202.
  85. Wu M,Frieboes HB, McDougall SR, Chaplain MA, Cristini V, et al. (2013) The effect of interstitial pressure on tumor growth: coupling with the blood and lymphatic vascular systems. J TheorBiol 320: 131-151.
  86. Swartz MA, Lund AW (2012) Lymphatic and interstitial flow in the tumour microenvironment: linking mechanobiology with immunity. Nat Rev Cancer 12: 210-219.
  87. Leu AJ, Berk DA, Lymboussaki A, Alitalo K, Jain RK (2000) Absence of functional lymphatics within a murine sarcoma: a molecular and functional evaluation. Cancer Res 60: 4324-4327.
  88. Padera TP,Kadambi A, di Tomaso E, Carreira CM, Brown EB, et al. (2002) Lymphatic metastasis in the absence of functional intratumor lymphatics. Science 296: 1883-1886.
  89. Alitalo K,Tammela T, Petrova TV (2005) Lymphangiogenesis in development and human disease. Nature 438: 946-953.
  90. Yan G, Zhou XY, Cai SJ, Zhang GH, Peng JJ, et al. (2008) Lymphangiogenic and angiogenicmicrovessel density in human primary sporadic colorectal carcinoma. World J Gastroenterol 14: 101-107.
  91. Dadras SS,Detmar M (2004) Angiogenesis and lymphangiogenesis of skin cancers. HematolOncolClin North Am 18: 1059-1070, viii.
  92. Kyzas PA,Stefanou D, Agnantis NJ (2005) COX-2 expression correlates with VEGF-C and lymph node metastases in patients with head and neck squamous cell carcinoma. Mod Pathol 18: 153-160.
  93. Yokoyama Y, Charnock-Jones DS, Licence D, Yanaihara A, Hastings JM,, et al. (2003) Expression of vascular endothelial growth factor (VEGF)-D and its receptor, VEGF receptor , as a prognostic factor in endometrial carcinoma. Clin Cancer Res 9: 1361-1369.
  94. Takeuchi H, Fujimoto A, Tanaka M, Yamano T, Hsueh E, et al. (2004) CCL21 chemokine regulates chemokine receptor CCR7 bearing malignant melanoma cells. Clin Cancer Res 10: 2351-2358.
  95. Valencak J, Heere-Ress E, Kopp T, Schoppmann SF, Kittler H, et al. (2004) Selective immunohistochemical staining shows significant prognostic influence of lymphatic and blood vessels in patients with malignant melanoma. Eur J Cancer40: 358-364.
  96. Vleugel MM,Bos R, van der Groep P, Greijer AE, Shvarts A, et al. (2004) Lack of lymphangiogenesis during breast carcinogenesis. J ClinPathol 57: 746-751.
  97. Cupedo T, Kraal G, Mebius RE (2002) The role of CD45+CD4+CD3- cells in lymphoid organ development. Immunol Rev 189: 41-50.
  98. Cupedo T,Mebius RE (2005) Cellular interactions in lymph node development. J Immunol 174: 21-25.
  99. Cupedo T,Mebius RE (2003) Role of chemokines in the development of secondary and tertiary lymphoid tissues. SeminImmunol 15: 243-248.
  100. Fütterer A, Mink K, Luz A, Kosco-Vilbois MH, Pfeffer K (1998) The lymphotoxin beta receptor controls organogenesis and affinity maturation in peripheral lymphoid tissues. Immunity 9: 59-70.
  101. Farnsworth RH,Karnezis T, Shayan R, Matsumoto M, Nowell CJ, et al. (2011) A role for bone morphogenetic protein-4 in lymph node vascular remodeling and primary tumor growth. Cancer Res 71: 6547-6557.
  102. Hirakawa S, Brown LF, Kodama S, Paavonen K, Alitalo K, et al. (2007) VEGF-C-induced lymphangiogenesis in sentinel lymph nodes promotes tumor metastasis to distant sites. Blood 109: 1010-1017.
  103. Harrell MI,Iritani BM, Ruddell A (2007) Tumor-induced sentinel lymph node lymphangiogenesis and increased lymph flow precede melanoma metastasis. Am J Pathol 170: 774-786.
  104. Hirakawa S,Detmar M, Kerjaschki D, Nagamatsu S, Matsuo K, et al. (2009) Nodal lymphangiogenesis and metastasis: Role of tumor-induced lymphatic vessel activation in extramammary Paget's disease. Am J Pathol 175: 2235-2248.
  105. Arai F,Suda T (2007) Maintenance of quiescent hematopoietic stem cells in the osteoblastic niche. Ann N Y AcadSci 1106: 41-53.
  106. Kaplan RN,Riba RD, Zacharoulis S, Bramley AH, Vincent L, et al. (2005) VEGFR1-positive haematopoietic bone marrow progenitors initiate the pre-metastatic niche. Nature 438: 820-827.
  107. Duong T,Koopman P, Francois M (2012) Tumor lymphangiogenesis as a potential therapeutic target. J Oncol 2012: 204946.
  108. Gandhi NG, Lin LK, O'Hara M (2013) Sildenafil for pediatric orbital lymphangioma. JAMA Ophthalmol 131: 1228-1230.
  109. Clarke JM, Hurwitz HI (2013) Understanding and targeting resistance to anti-angiogenic therapies. J GastrointestOncol 4: 253-263.
  110. Ebos JM, Lee CR, Cruz-Munoz W, Bjarnason GA, Christensen JG, et al. (2009) Accelerated metastasis after short-term treatment with a potent inhibitor of tumor angiogenesis. Cancer Cell 15: 232-239.
  111. Cascone T,Heymach JV (2012) Targeting the angiopoietin/Tie2 pathway: cutting tumor vessels with a double-edged sword? J ClinOncol 30: 441-444.
  112. D'Amico G, Korhonen EA, Anisimov A, Zarkada G, Holopainen T, et al. (2014) Tie1 deletion inhibits tumor growth and improves angiopoietin antagonist therapy. J Clin Invest 124: 824-834.
  113. Doloff JC, Waxman DJ (2012) VEGF receptor inhibitors block the ability of metronomically dosed cyclophosphamide to activate innate immunity- induced tumor regression. Cancer Res72: 1103-1115.
  114. Vecchiarelli-Federico LM,Cervi D, Haeri M, Li Y, Nagy A, et al. (2010) Vascular endothelial growth factor--a positive and negative regulator of tumor growth. Cancer Res 70: 863-867.
  115. Lund AW,Duraes FV, Hirosue S, Raghavan VR, Nembrini C, et al. (2012) VEGF-C promotes immune tolerance in B16 melanomas and cross-presentation of tumor antigen by lymph node lymphatics. Cell Rep 1: 191-199.
  116. Sitohy B, Nagy JA, Dvorak HF (2012) Anti-VEGF/VEGFR therapy for cancer: reassessing the target. Cancer Res 72: 1909-1914.
  117. Fraser DM, Sullivan FM, Thompson AM,McCowan C4 (2014) Aspirin use and survival after the diagnosis of breast cancer: a population-based cohort study. Br J Cancer 111: 623-627.
  118. Holmes MD, Chen WY, Li L, Hertzmark E, Spiegelman D, et al. (2010) Aspirin intake and survival after breast cancer. J ClinOncol 28: 1467-1472.
Citation: Francois M, Shayan R, Karnezis T (2014) Ordered Chaos: Harnessing Developmental Pathways in Tumor-Induced Lymphangiogenesis. J Clin Cell Immunol 5:270.

Copyright: © 2014 Francois M, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top