Next Article in Journal
Correction: Alessia Stornetta et al. Alcohol-Derived Acetaldehyde Exposure in the Oral Cavity. Cancers 2018, 10, 20
Next Article in Special Issue
The Complex Interplay between Chronic Inflammation, the Microbiome, and Cancer: Understanding Disease Progression and What We Can Do to Prevent It
Previous Article in Journal
Update on Immunohistochemistry for the Diagnosis of Lung Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Coming of Age for Autotaxin and Lysophosphatidate Signaling: Clinical Applications for Preventing, Detecting and Targeting Tumor-Promoting Inflammation

by
Matthew G.K. Benesch
1,2,
Iain T.K. MacIntyre
1,
Todd P.W. McMullen
3 and
David N. Brindley
2,*
1
Discipline of Surgery, Faculty of Medicine, Memorial University of Newfoundland, St. John’s, NL AlB 3V6, Canada
2
Signal Transduction Research Group, Cancer Research Institute of Northern Alberta, Department of Biochemistry, Faculty of Medicine and Dentistry, University of Alberta, Edmonton, AB T6G 2S2, Canada
3
Department of Surgery, Faculty of Medicine and Dentistry, University of Alberta, Edmonton, AB T6G 2G7, Canada
*
Author to whom correspondence should be addressed.
Cancers 2018, 10(3), 73; https://doi.org/10.3390/cancers10030073
Submission received: 23 February 2018 / Revised: 10 March 2018 / Accepted: 12 March 2018 / Published: 15 March 2018
(This article belongs to the Special Issue Inflammation and Cancer)

Abstract

:
A quarter-century after the discovery of autotaxin in cell culture, the autotaxin-lysophosphatidate (LPA)-lipid phosphate phosphatase axis is now a promising clinical target for treating chronic inflammatory conditions, mitigating fibrosis progression, and improving the efficacy of existing cancer chemotherapies and radiotherapy. Nearly half of the literature on this axis has been published during the last five years. In cancer biology, LPA signaling is increasingly being recognized as a central mediator of the progression of chronic inflammation in the establishment of a tumor microenvironment which promotes cancer growth, immune evasion, metastasis, and treatment resistance. In this review, we will summarize recent advances made in understanding LPA signaling with respect to chronic inflammation and cancer. We will also provide perspectives on the applications of inhibitors of LPA signaling in preventing cancer initiation, as adjuncts extending the efficacy of current cancer treatments by blocking inflammation caused by either the cancer or the cancer therapy itself, and by disruption of the tumor microenvironment. Overall, LPA, a simple molecule that mediates a plethora of biological effects, can be targeted at its levels of production by autotaxin, LPA receptors or through LPA degradation by lipid phosphate phosphatases. Drugs for these applications will soon be entering clinical practice.

1. Introduction—Cancer as the Ultimate Disease of Chronic Inflammation

Rudolf Virchow, known as the father of modern pathology, observed under a microscope a high concentration of leukocytes in neoplastic tissues. He proposed in 1863 that the “lymphoreticular infiltrate” reflected the origins of cancer at sites of chronic inflammation [1]. His discovery however remained largely ignored for over a century. In that time, there was an ebb and flow of numerous ideas behind the origins of cancer, including viruses, environmental exposures, and genetic factors. These various ideas have at times pitted scientists against each other in debate as to the origins of cancer. However, given the vast heterogeneity of this disease, it is now recognized that they all are indeed correct for particular cancers. During the 1990s, after over a century of intense investigation, our knowledge about what cancer is as a disease began to coalesce, and it was codified and revised by Hanahan and Weinberg as the hallmarks of cancer [2,3]. Cancer is a disease of sustained proliferative signaling, evasion of growth suppressors, replicative immortality, angiogenesis, resistance to cell death, deregulation of cellular genetics, avoidance of the immune system, and invasion and metastasis [2,3]. These traits are enabled through two main characteristics common to virtually all cancers: genomic instability and mutation, and tumor-promoting inflammation [3].
As Virchow noted, virtually all neoplastic lesions contain immune cells, and it has been long recognized by pathologists that these tumor-associated cells come from both the innate and adaptive arms of the immune system [4]. This immune cell composition is also present in inflamed non-neoplastic tumors, suggestive that the immune system actively attempts to destroy pre-malignant and early cancer cells, at least initially [4]. However, if there is no resolution of the injury, this milieu of perpetually smoldering inflammatory signaling (wounds that do not heal) is exploited by these neoplastic cells to complete the transformation into an established cancer [4,5]. This connection between chronic inflammation in many diseases such as hepatitis and inflammatory bowel diseases and their eventual transformation into cancers, has been coined the extrinsic pathway [6,7] (Figure 1). In this chronic inflammatory state, reactive oxygen species and other mutagenic molecules initiate genetic transformations leading to oncogene activation and loss of tumor suppressor functionality, termed the intrinsic pathway [6,7] (Figure 1). Together, these two pathways serve to upregulate a host of pro-inflammatory transcription factors, namely nuclear factor-κB (NF-κB), signal transducer and activator of transcription 3 (STAT3) and hypoxia-inducible factor 1α (HIF1α) in cancer cells [6]. These same signaling pathways also serve to promote cell survival and evasion from the immune system [6,8] (Figure 1). The overall net result is an increase in the production of cytokines and chemokines by neoplastic cells that spills over into adjacent stromal cells, resulting in more inflammatory mediators being produced, with further leukocyte activation and recruitment [6,9]. Hence, tumor-promoting inflammation both enables and is a product of genetic instability, and accentuates the other hallmarks of cancer [9]. Consequently, there is at least an 80% correlation between tumor leukocyte density and poor patient prognosis [10], with at least 20% of all cancer deaths being linked to underlying inflammatory processes [6].
Over the last decade, it has become well acknowledged that cancer can no longer be considered a disease of just cancer cells, but instead is a disease of unchecked cell growth and spread within a permissive host environment [11]. Now, in the age of targeted therapies, modern cancer therapy regimens need to disrupt the interactions between cancer cells and supportive stromal elements to mitigate both pro-inflammatory and pro-survival signaling and development of therapy resistance [12]. Appreciating and targeting chronic tumor-promoting inflammation through adjunct therapy may ultimately be the biggest breakthrough in cancer treatment in the next decade [13], and may be the most effective modality for cancer prevention, improving cancer therapy efficacy and maintaining remission.
In this review, we propose that targeting chronic inflammation through inhibition of the autotaxin-lysophosphatidate-lipid phosphate phosphatase (ATX-LPA-LPP) axis provides one viable strategy for achieving this breakthrough. As potent anti-inflammatory agents, such inhibitors have the potential to be powerful adjuvant agents for improving the efficacy of cancer treatment and preventing therapy resistance irrespective of the genomic instability within cancer cells [14,15,16]. In the last five years, the literature on LPA signaling axis has nearly doubled. With the advent of multimodality clinical trials for targeting the ATX-LPA-LPP axis in chronic inflammatory diseases, we will summarize new findings to complement our previous reviews on this subject [15,16,17].

2. Overview of LPA Signaling and Its Importance in Reproduction and Embryonic Development

As the simplest bioactive glycerolphospholipid, LPA contains a glycerol backbone with a phosphate head group, and an acyl chain (usually unsaturated) at the sn-1 (or sn-2) position (Figure 2). Extracellular LPA is generated primarily from lysophosphatidylcholine (LPC) in plasma by the lysophospholipase D activity of ATX, as demonstrated by ATX heterozygote mice having half-normal LPA levels [18,19]. ATX inhibition provides >95% reduction of LPA levels in plasma [20]. Conditional gene knockout of ATX in adipose tissues in mice results in up to a 38% decrease in circulating LPA levels, thus demonstrating the importance of adipose tissue in ATX secretion [21]. Some LPA, primarily saturated species, is also produced by hydrolysis of a fatty acid chain from the membrane-derived phosphatidate via phospholipase A1 and A2 activity in inflammatory cells, activated platelets and endothelial cells [22]. LPA mediates its plethora of effects by signaling through at least six G-protein coupled LPA receptors (LPA1–6) [17]. There are other orphan G-protein coupled receptors that may facilitate LPA signaling, including GPR35 [23], GPR87 [24,25], P2Y10 [26], and the receptor for advanced glycation end products (RAGE) [27]. LPA is turned over rapidly (t1/2 of ~1 min) in plasma into monoacylglycerol and inorganic phosphate by the ecto-activity of three lipid phosphate phosphatases (LPP1-3) [17,28] (Figure 2).
This ecto-activity helps regulate the total LPA pool, and in particular reduces LPA concentrations in the immediate cellular microenvironment [29]. This may in part explain why cancer cells, which become dependent on sustained LPA signaling, normally have decreased LPP expression (see Section 6). We refer readers to several other recent reviews for a more detailed description of LPA receptor and LPP biology [29,30,31,32,33].
ATX, an ectonucleotide pyrophosphatase/phosphodiesterase (ENPP) enzyme, is encoded by the ENPP2 gene. This is one in a family of seven mammalian enzymes that belong to the superfamily of ecto-nucleotideases that hydrolyze pyrophosphate and phosphodiester bonds of a wide range of nucleotides and their derivatives [34,35]. ATX is unique in having a specific lysophospholipase D activity owing to its additional hydrophobic binding pocket that interacts with the acyl chain of lysophospholipids [36]. As a secreted glycoprotein of about 125 kDa, full-length ATX is synthesized as a pre-proenzyme and is secreted by the classical secretory pathway under the influence of Akt-signaling [37,38]. ATX was first discovered in A2058 melanoma cell culture in 1992 and it was later shown to hydrolyze LPC preferentially [39,40]. Five isoforms of ATX have since been discovered in human beings [41,42]. Despite this, under physiological conditions, all isoforms have largely similar kinetic effects and thus their biological significance is unknown [42]. Nevertheless, all critical residues and structural elements are highly conserved between mice and human beings, underscoring its physiological importance [43].
In relation to reproduction, endometrial stromal cells in culture are highly sensitive to LPA signaling as demonstrated by growth inhibition with either ATX or LPA receptor inhibitors. The source of this LPA is from autocrine-produced ATX [44]. Multiple LPA receptors are likely involved since equal proliferation occurs in endometrial cells with either LPA1 or LPA3 knockouts compared to wild type cells [44]. Strong correlations have also been found in animal models between ovarian follicular growth and LPA signaling [45,46]. ATX is also a tissue-remodeling factor in regressing corpora lutea [47]. Another study showed that placental transcription of ATX increases progressively during normal pregnancy, and a disturbance in placental ATX production is linked to early pre-eclampsia [48]. Serum ATX levels show a significant correlation with maternal blood pressures and uric acid levels, both of which are parameters for the severity of pre-eclampsia [49]. Targeting ATX/LPA signaling could limit pregnancy complications associated with pre-eclampsia, and this represents an emerging field where LPA-modulating therapies could hold significant therapeutic promise.
It has been known for over a decade that complete knockout of ATX in mice is embryonically lethal at day 9.5, secondary to profound vascular and neural crest defects [18,19]. Later work showed that this lethality was due to the absence of ATX catalytic activity, since a single point mutation of the catalytically active amino acid threonine-210 to alanine produces the same lethal phenotype [50]. Ex vivo, the neural defects were partially ameliorated by addition of LPA to culture [51]. In zebrafish, ATX knockout studies have shown that the ATX-LPA-histone deacetylase (HDAC)1/2 axis regulates oligodendrocyte differentiation in the hindbrain [52,53], and that ATX/LPA3 signaling regulates left-right asymmetry [54]. Therefore, ATX-mediated signaling in neurological development is an evolutionally conserved mechanism from at least zebrafish to rodents [53].
More recently, ATX overexpression in mouse embryos was demonstrated to be lethal at around day 9.5–11.5 with growth retardation, open and kinked neural tubes, abnormal allantois, and vascular defects [55]. In this same study, induced overexpression of ATX in the neonatal period caused vascular instability, a delay in retinal vascularization, and a decrease in vessel branching [55]. Interestingly, LPP3 knockout in mice is also embryonically lethal with very similar vascular defects to those embryos overexpressing ATX [56,57]. Like the knockout studies, overexpression of ATX in the zebrafish embryo leads to distortion of axial midline and left-right asymmetry in the embryo through the Rho/ROCK pathway via LPA1–3 signaling. [58]. Overexpression via injection of ATX messenger RNA (mRNA) in zebrafish embryo also induces cardia bifida [59]. Therefore, it appears that LPA concentrations must be tightly regulated to ensure proper development. This is especially apparent from mouse models, as both ATX knockout and overexpression are embryonically lethal [18,19,55]. With respect however to LPA receptor signaling, there is significant redundancy since no individual, double or triple LPA receptor knockout mouse model is lethal [60].

3. LPA Signaling in Wound Healing and Immunity

In the post-natal organism, one of the most important roles of the ATX-LPA-LPP axis is mediating wound healing and tissue remodeling. LPA is a potent activator of platelet aggregation and it promotes the growth and migration of immune cells, fibroblasts, endothelial cells and keratinocytes into sites of tissue damage. The mechanisms for these effects have been reviewed elsewhere [17,61,62,63].
There have been multiple recent advances in LPA biology, most of which relate to immune system regulation and tissue repair. In the innate immune system, LPA promotes lymphocyte extravasation, which maintains immune homeostasis by stimulating the conversion of CD11b+ murine monocytes into F4/80 macrophages. A similar effect also occurs in human beings through a common transcription factor, peroxisome proliferator-activated receptor (PPAR)γ [62]. ATX is induced in immune cells in response to toll-like receptor activation by lipopolysaccharide, via a type I interferon autocrine-paracrine loop involving the JAK-STAT and PI3K-Akt pathways [64]. ATX is highly expressed in lymph node stromal cells promoting interstitial T cell migration in lymph nodes primarily through LPA2 signaling [65,66,67]. LPA promotes and converts primed pluripotent stem cells into naïve stem cells through LPA1-STAT3 signaling pathways, which has implications in regenerative medicine [68].
Our understanding of the role of LPA in neuronal inflammation and regeneration is also evolving and demonstrates new complexities in LPA receptor signaling not previously well appreciated. Serum LPA concentrations and ATX levels in the central nervous system are decreased in patients with multiple sclerosis (MS) compared to healthy controls [69,70]. LPA levels are restored during symptom-free and recovery intervals in murine models of experimental autoimmune encephalomyelitis (EAE), which mimic MS [69]. These influences appear to be mediated through LPA2 since there is more intense disease in LPA2-deficient mice, and recovery in wild-type mice is improved with LPA2 agonist therapy [69]. This recovery appears to be mediated by increasing T-cell homing and promoting remyelination [69]. This is in line with other work demonstrating that reductions in ATX-LPA-HDAC1/2 signaling may contribute to MS pathology and that LPA is important for oligodendrocyte maturation [53,71]. However, significant reduction of total LPA concentrations with the ATX inhibitor Compound-1 resulted in deceased disease severity in a very similar murine model of EAE [72]. These two contrasting studies highlight the emerging complexity of LPA receptor-mediated signaling. While LPA2 signaling may limit disease intensity, LPA signaling through other receptors may have the opposite effect, meaning that disease burden could be dependent on subtle changes in receptor abundance.
In both zebrafish and mice, ATX-LPA1 signaling contributes to proliferation of chondrocytes by promoting S-phase entry and regulating fibronectin assembly, leading to proper cartilage formation [73]. Finally, C18:1-LPA signaling may be a master transcriptional regulator of primary human gingival fibroblasts that mediate gingival repair, but under chronic inflammatory conditions it promotes periodontal disease [74,75,76].
The induction of ATX by acute inflammation in tissue repair has been well appreciated for many years, but the regulation of this process is only now starting to emerge. We showed that LPA normally downregulates ATX production at the mRNA level through LPA1-PI3K signaling [77]. This suppression is overcome via parallel induction of ATX synthesis by signaling through inflammatory cytokines such as TNFα and IL-1β [77]. ATX is eliminated rapidly from the circulation through a scavenger receptor-mediated process in liver sinusoidal endothelial cells [78]. Therefore, once this inciting injury is resolved, basal expression of ATX, and in turn LPA concentrations, are restored [77].

4. Maladaptive LPA Signaling in Chronic Inflammatory Diseases

This regulation of ATX turnover by inflammatory cytokines also explains why elevated ATX and LPA levels persist in chronic inflammation. In addition, it is also well established that LPP1 and LPP3 are downregulated in cancers, which aggravates the increase in LPA levels caused by the elevated ATX [29]. An emerging hallmark of chronic inflammation is sustained LPA production and signaling, leading to propagation of the disease phenotype. These pathologies are now being elucidated in vivo in conditional ATX tissue knockouts, especially in models of arthritis, pulmonary fibrosis and liver diseases [79,80,81]. The chronic inflammatory states in many of these conditions often progress into cancer, as will be discussed in the next section. Below, we summarize major recent advances in LPA signaling in chronic inflammation.

4.1. ATX and LPA in Inflamed Adipose Tissue and the Metabolic Syndrome

There is considerable evidence that LPA contributes to the chronic low-level inflammatory state associated with a myriad of complications tied to obesity, insulin resistance, diabetes, dyslipidemia, atherosclerosis, and hypertension in the Metabolic Syndrome [82]. About 40% of the ATX in mice is produced by adipose tissue, and this increases further with increased dietary fat consumption [83]. This increased ATX secretion is compounded by the expansion of total adipose tissue mass in obesity. Inflammation of adipose tissue also increases ATX expression [77]. Patients with type 2 diabetes have higher ATX levels in serum and visceral fat than non-diabetic patients, and ATX correlates positively with body fat percentage and increased leptin/adiponectin ratios. This correlates negatively with glucose clearance rates [21,84,85]. Leptin and adiponectin have opposing effects on low-grade adipose tissue inflammation and insulin resistance, with leptin upregulating proinflammatory cytokines, and adiponectin having the opposite effect [86]. Thus, the leptin/adiponectin ratio is being increasingly positively correlated to the co-morbidities of the Metabolic Syndrome [87,88]. Similarly, serum ATX is an independent predictor of insulin resistance in older obese adults [89]. In mice lacking ATX expression in white adipose tissue, adiponectin and GLUT-1 levels were increased, leading to increased glucose tolerance [21]. Transgenic mice overexpressing ATX exhibit reduced expression of thermogenic brown adipose tissue in exchange for peripheral white adipose tissue, and have increased weight gain when fed a high fat diet [90].
LPA is able to attenuate insulin signaling in rat hepatocytes via LPA3 signaling through decreased hepatocyte glucose uptake and glycogen synthesis [91]. In the same study, the authors reported LPA 16:0 concentrations to be about 25 percent higher in obese patients with a body mass index (BMI) >30 compared to normal weight individuals (BMI 18.5–25) [91]. ATX expression in adipocytes has recently been shown to be upregulated through the glycoprotein 130 (gp130)-mediated Janus kinase (JAK)-STAT3 pathway via IL-6 family cytokines [92]. Treating high-fat diet-fed obese mice with the oral gp130 inhibitor SC144 suppressed adipose tissue ATX expression, resulting in decreased plasma ATX, LPA and free fatty acid levels, and increased glucose tolerance [92].
While there is strong in vivo and clinical evidence supporting the effect of ATX/LPA in adiposity, there is still some controversy in the strength of the association. This is because a negative relation between BMI and ATX expression levels has been reported in both adipose tissue and serum of non-diabetic human subjects [93]. However, in the same report, serum ATX levels tended to be higher in diabetic subjects [93], and most evidence supports increased ATX/LPA levels in diabetics [84,94]. These discrepancies might in part be rationalized by the product inhibition of LPA on ATX expression under physiological conditions [77]. While increased adipose tissue may lead to increased ATX production, the subsequent increase in LPA concentrations acts as a rheostat to restore balance, leading to eventual decreased ATX levels. However in a more pro-inflammatory condition, such as diabetes, the additional cytokine stimulation of ATX production is enough to overcome this inhibition, a phenomenon best illustrated in cancer models [77,95]. Further delineation of ATX/LPA signaling pathways in obesity is currently a very active area of research.
Several authors have linked LPA signaling to cardiac diseases. MMP-9 is particularly associated with rupture of atherosclerotic plaques in coronary arteries, and LPA upregulates MMP-9 expression through LPA2 signaling and nuclear factor-κB (NF-κB) pathways in macrophages [96]. Fasting plasma ATX concentration is a novel independent predictor of calcific aortic valve stenosis in patients with coronary artery disease [97,98]. Mechanistically, ATX is transported to the aortic value by apolipoprotein (a) and is secreted by valve interstitial cells, leading to inflammation and calcification [99]. In contrast, LPP3 negatively regulates aortic endothelial cell inflammation and could mitigate the development and complications of atherosclerosis [100]. LPP3 silencing in human primary aortic endothelial cells enhances cytokine secretion and leukocyte adhesion whileimpairing angiogenesis, whereas LPP3 overexpression reverses these effects [101,102]. Similar effects occur via pharmacological inhibition of both LPA and sphingosine 1-phosphate signaling, both substances that are degraded by LPP3 [101].

4.2. ATX and Neurological Disorders

LPA also acts synergistically with adipokines (leptin, resistin, TNFα, IL-1β and IL-6) released by white adipose tissue in promoting inflammation and producing reactive oxygen species that disrupt blood brain barrier permeability, leading to hippocampal atrophy and dementia development [103]. Significantly, increased ATX levels are associated with a 3.5- to 5-fold risk of mild cognitive impairment and Alzheimer’s disease [104]. Other conditions well reviewed elsewhere include neuropathic pain secondary to inflammation-mediated demyelination of axons [105,106,107].

4.3. ATX and Arthritis

ATX concentrations in plasma and synovial fluid correlate with the severity of knee osteoarthritis [108]. Mice fed high fat diets exhibit both accelerated surgically-induced and age-related osteoarthritis [109]. In these mice, leptin contributes to metabolic and catabolic changes in articular cartilage, resulting in local increased ATX expression. This in turn promotes cartilage degeneration by increased MMP-13 production, which is the major catabolic enzyme responsible for cleaving type II collagen in cartilage [109]. ATX and MMP-13 are significantly increased in leptin-treated chondrocyte cultures compared to vehicle treated cultures and inhibition of ATX activity decreased leptin-induced MMP-13 expression [109]. Increased ATX levels and LPA signaling are also involved in rheumatoid arthritis, in part by promoting synovial hyperplasia, which lead to progressive destruction of cartilage and bone [17,110].

4.4. ATX and LPA in Pulmonary Fibrosis and Asthma

LPA-mediated lung fibrosis is a very established field of research for which at least two LPA receptor antagonists and an ATX inhibitor have entered clinical trials [16] (Section 7). The foundational animal studies behind these clinical trials demonstrated in a bleomycin-induced mouse model of lung fibrosis that fibroblast migration depends on LPA1 signaling. Mice that lacked LPA1 were more resistant to fibrosis via bleomycin challenge [111,112]. In response to bleomycin, ATX in bronchoalveolar lavage fluid increased 17-fold and this ATX appeared to be of plasma origin that enters the alveolar space via vascular leak [113]. In mouse models of lung allograft fibrosis, mesenchymal cells from fibrotic lung allografts have constitutive β-catenin expression [114]. β-catenin expression depends strongly on autocrine ATX signaling, and both ATX and β-catenin are regulated by the transcription factor nuclear factor of activated T-cells (NFAT1) [114]. Consequently, inhibition of LPA1 or ATX, either genetically or pharmacologically, limits the development of fibrosis through decreased active β-catenin and dephosphorylated NFAT1 expression levels [114]. In another model of hyperoxic lung injury, capillary leakage and respiratory distress were limited by Brp-LPA, a commonly used pan-LPA receptor/ATX inhibitor [115]. Other studies, which were reviewed previously, have associated asthma with upregulation of ATX expression [17,111]. This has now been shown to be mediated primarily through LPA2 signaling [111]. Polyunsaturated LPAs, primarily C22:5-LPA and C22:6-LPA, are being investigated as potential biomarkers for the severity of allergy-mediated airway inflammation and disease in asthma [116].

4.5. ATX and LPA in Autoimmune and Retinal Diseases

Other chronic inflammatory and autoimmune diseases that are now recognized to be mediated by LPA signaling, include Sjogren’s syndrome, a condition causing dry eyes and mouth. In mice, Sjogren’s syndrome is exacerbated by LPA1/3-mediated production of IL-17 in a dose-dependent manner. Saliva volume is restored by treatment with the LPA1/3 antagonist, Ki16425 [117]. In retinal disease, macular edema is medically treated with steroids and anti-vascular endothelial growth factor (VEGF) agents [118]. LPA and ATX concentrations are positively correlated with the proinflammatory mediators, IL-6, IL-8, MCP-1 and VEGF in patients with macular edema secondary to retinal vein occlusion [119], this finding opens the possibility for ATX/LPA targeted therapy.

4.6. ATX and LPA in Hepatic Diseases

Pathogenic LPA signaling is increasingly implicated in a host of hepatic disease conditions. The severity of one of these conditions, biliary atresia, a progressive fibro-inflammatory liver disease, correlates independently with both liver stiffness and serum ATX levels [120]. In a later follow-up report, and the first to study ATX promoter methylation, DNA was extracted from both circulating leukocytes and liver tissues of biliary atresia patients and age-matched controls [121]. Patients with biliary atresia have decreased ATX promoter methylation that was even lower in patients with advanced disease [121]. Consequently, relative ATX mRNA and serum protein levels are both inversely correlated with overall ATX methylation level and positivity associated with jaundice status, hepatic dysfunction and liver stiffness [121]. Mechanistically, fibrosis by LPA signaling is a two-step process whereby ATX expression is largely confined to hepatocytes in response to injury, and LPA1 is highly expressed in the collagen-secreting hepatic stellate cells [122]. Patients with chronic liver disease of various origins (viral, alcoholic, or fatty liver) have both increased ATX and LPA6 expression [81]. This same pattern has been previously correlated to tumorigenicity in hepatocellular carcinoma [123].
Serum ATX may also be a potential biomarker for non-alcoholic fatty liver disease (NAFLD) in obese non-diabetic women. This correlates significantly with other markers of NAFLD including alkaline phosphatase and fasting glucose, insulin and triglyceride levels [124].
Many observational studies and reviews have linked ATX to cholestatic pruritus and liver injury [125,126]. In the first prospective study of its kind, patients with primary biliary cholangitis and primary sclerosing cholangitis were followed for 60 months [127]. Serum ATX was significantly higher in both groups compared to controls and in patients with either cirrhosis or longer disease duration [127]. ATX activity correlated strongly with Model for End-Stage Liver Disease (MELD) scores, Mayo Risk scores, and worse disease-specific health-related quality of life (HRQoL) aspects [127]. High ATX levels are a negative predictor of survival, with such patients at a 2.6- to 4-fold increased risk of death or of requiring liver transplantation [127].
Chronic viral hepatitis infection that has progressed to cirrhosis accounts for 70% of all hepatocellular carcinomas, which represents 90% of all liver cancers [128]. Serum ATX is a superior marker for predicting significant fibrosis in both chronic hepatitis B and C patients. This can be analyzed without consideration of food intake prior to phlebotomy [129]. More recent studies of both hepatitis B and C showed that serum ATX was the most reliable marker for all fibrosis stages compared with other serum markers including hyaluronic acid, type IV collagen 7S, aspartate aminotransferase-to-platelet ratio and FIB-4 index [130,131]. Mechanistically, hepatitis C infection stabilizes hypoxia inducible factors, which in turn increased hepatocellular ATX expression. This cycle is reinforced since LPA signaling through PI3K stabilizes hypoxic inducible factor (HIF)-1α [132]. Elevated serum ATX levels in patients with hepatitis C or co-infection with HIV partially normalize within 6 months of starting interferon-free hepatitis C therapy [133]. Therefore, ATX has significant potential to be both a diagnostic and therapeutic target in mitigating the progression of viral hepatitis into cancer [128,134].

4.7. Importance of the Site of ATX Production in Inflammatory Signaling

One of the outstanding problems in understanding the physiological and pathological roles of ATX secretion is whether signaling by LPA is regulated locally rather than by globally produced ATX. It was established from structural studies of ATX that it binds locally to integrins on cells surfaces [63,135] and that this delivers LPA to signal through nearby receptors [36]. The reports that ATX expression in specific cells, such as adipocytes, fibroblasts, hepatocytes, and lung parenchyma, leads to pathology in different organs [79,80,81,136,137] adds weight to the concept that it is locally produced ATX that has a major signaling role. Locally produced ATX interacting with integrins on nearby cells is likely the primary means by which inflammation in wound healing and tissue repair is localized to the site of injury. However, ATX is just as catalytically active when not bound to integrins [138], which may also explain why plasma LPA levels are increased in numerous disease processes [77,124,133]. Overall, the function however of plasma ATX is still uncertain: it could represent transport among organs and/or the ATX that is en route for hepatic degradation.

5. Recent Developments in LPA Signaling in Cancer Biology

Since ATX was first discovered in melanoma cell culture, ATX and LPA signaling have been associated with cancer initiation and progression, survival against cancer therapy, and metastasis for virtually every major cancer type [17,139]. The common theme is that chronic inflammation upregulates ATX production by either the cancer cells themselves or in adjacent tumor stroma. This is often accompanied with a concomitant overexpression of LPA receptors, particularly LPA1–3, and a downregulation of the ecto-activities of the LPP1/3 [17]. The overall net effect of sustained LPA signaling is to increase angiogenesis and cancer cell growth, migration, and survival.

5.1. Mechanisms of ATX Overexpression in Cancers

For those cancers that secrete ATX, several mechanisms of overexpression are now appreciated. Most genomic alterations in the ATX gene are copy number amplifications [22,35]. In other cases, ATX is over-expressed in response to DNA double-strand breaks secondary to oxidative stresses in cancer cells. This depends on activation of the NLRP3 inflammasome and ataxia telangiectasia mutated (ATM) phosphorylation. The ATM-ATX-dependent loop further propagates inflammation and additional double strand breaks, leading to further ATX production [140]. The first study to look at post-transcriptional regulation of ATX demonstrated that the RNA-binding protein ELAV-like protein 1 or human antigen R (HuR) enhances ATX mRNA stability in melanoma cells and thus increases ATX production [141]. LPA signaling in turn increases HuR protein expression [141]. HuR stabilizes the mRNA for many pro-inflammatory proteins and it is itself a prognostic factor for poor outcome in ovarian and breast cancers [142,143]. We showed previously that ATX is significantly upregulated in thyroid cancer in response to sustained autocrine-induced inflammation, and this can distinguish cancer from benign disease [144]. Since then, HuR has shown to be overexpressed in thyroid cancers but not benign nodules [145]. Taken together, these findings suggest that copy number amplifications and inflammatory-induced mRNA stability resulting in increased ATX translation are likely major mechanisms behind autocrine ATX production in cancers.
A seminal study showed in a mouse mammary tumor virus (MMTV) model that overexpression of ATX or any of LPA1-3 increased mammary tumorigenesis, invasion and metastasis [146]. However, breast cancer cells secrete little ATX compared to the basal rate from breast adipose tissue [137]. We reconciled this discrepancy and uncovered a paracrine-model of ATX production in breast tumors. This involved an immunocompetent orthotopic murine breast cancer model. We demonstrated that secretion of inflammatory mediators from the breast tumor induce ATX expression in adjacent mammary tissue. This subsequently increases LPA within the tumor microenvironment [137,147] and it establishes a vicious loop of inflammatory-driven ATX production since LPA increases the production of more inflammatory cytokines and COX-2. This cycle can be broken with a potent ATX inhibitor (Section 7). Likewise, tumor-induced inflammation in mammary adipose tissue increases macrophage recruitment, which leads to further inflammation and ATX production [137]. These bidirectional interactions between breast cancer cells and ATX production in mammary adipose tissue have since been confirmed [148]. Increased ATX production in adipose tissues could provide a possible link between obesity and its contribution to an estimated 20–40% of breast cancers [149,150].

5.2. LPA Signaling in Cancer Progression

Hepatocellular carcinoma (HCC) is well established to be associated with LPA signaling. HCC has a 5-year survival rate of less than 15% primarily due to late detection and poor screening [122]. HCC is usually a consequence of chronic inflammation from viral hepatitis and patients are at an eight-fold increased risk of death if ATX levels are elevated [151]. Higher LPA2 mRNA levels in HCC specimens are correlated with poorer differentiation, and higher LPA6 mRNA levels are correlated with microvascular invasion. These observations suggest an overall higher malignant potential with increased LPA2/6 expression [152]. This study, like others, correlated ATX with disease progression, but more interestingly showed that the mRNA level of phospholipase A1α, another LPA producing enzyme, has no pathological association [152]. Further, hepatocyte-specific ATX-deficient mice were protected from fibrosis and HCC development compared to controls [81]. Even with the decreasing incidence of Hepatitis B and C infections with more efficacious therapies, NASH will become the primary cause of HCC in the western world by 2030 because of increasing obesity rates [153]. Given the well-established roles of LPA signaling in the pathogenesis of metabolic-related syndromes, ATX/LPA signaling inhibitors should also have therapeutic benefit in HCCs from these origins.
Increased serum ATX activity was most significant for exocrine pancreatic cancer in an evaluation of serum ATX levels in a host of gastrointestinal cancers including cancers of the esophagus, stomach, colon, biliary tract and pancreas [22,154]. Exocrine pancreatic cancer is the fifth major cause of cancer death in the developed world, with a 5-year survival rate of less than 5% [155]. Therefore, serum ATX in combination with other biomarkers could one day offer an early diagnostic tool. However, serum ATX activity is not increased in patients with chronic pancreatitis or pancreatic cysts and neither condition is predictive of pancreatic cancer risk [154]. Lastly, in another study comparing the ascites of gastric cancer patients to those of cirrhosis, higher ATX and LPA levels were found in cirrhosis patients as expected [156]. However, higher lysophosphatidylserine and lysophosphatidylglycerol levels were found in gastric cancer ascites, which may act as substrates for LPA generation [156]. This suggests that glycerol-lysophospholipids other than LPA might be involved in pathogenesis of cancer directly or through being converted into LPA [156].
Advances in the LPA field have also been made in gynecological cancers. Ovarian cancers have a high mortality because they metastasize easily and through the development of resistance to chemotherapy [157]. These characteristics are mediated by cancer stem cells, which in ovarian cancer produce abundant ATX and they express high LPA1 [158]. Knockdown of either ATX or LPA1 leads to loss of stem cell markers and decreased tumorgenicity in xenografts [158]. Thus, the ATX-LPA-LPA1-Akt1 axis maintains cancer stem cell characteristics through an autocrine loop [158]. LPA1 and LPA2 are also involved in LPA-induced proliferation and angiogenesis in endometrial cancer tissue with positive correlations between LPA receptor and ATX protein levels with cancer stage [159].
In connection to other cancers, HT1080 fibroscarcoma cells that have been treated chronically with cisplatin have markedly elevated ATX and LPA2 expression and are much more mobile than controls [160]. Knockdown of LPA2 in these long-term treated cells reduced colony formation and autocrine production of ATX [160]. Multiple myeloma cells stimulate mesenchymal stromal cells to produce ATX [161]. In renal cell carcinomas, LPA activates Afr6-regulated mesenchymal pathways via LPA2 that promote cancer cell plasticity, metastasis and drug resistance [162]. Finally, several new studies have expanded our knowledge of LPA biology in melanoma, the first cancer linked to ATX. The transcription factor NFAT1 has roles in both innate and adaptive immune responses, and increases metastatic potential by directly upregulating IL-8 and MMP-3. This is secondary to autocrine ATX production and LPA signaling [163,164]. In work with the highly metastatic B16F10 murine melanoma, mice lacking either LPA1 or LPA5 had significantly less metastasis, indicating the importance of host LPA signaling for establishing a permissible environment for cancer cell seeding [165,166].

6. The Roles of LPPs in Cancers

Evidence is now mounting from experiments in vivo that increasing the low LPP1/3 expression in cancer cells limits tumor progression. Increasing LPP1 expression in syngeneic and xenograft breast and thyroid cancer models decreased tumor growth and metastasis by up to 80% through both increasing extracellular LPA degradation and decreasing the stimulation of Ca2+-transients by LPA [167]. Similarly, increasing the low LPP3 activity in SKOV3 ovarian cancer cells decreased tumor growth in nude mice [168]. Low expression levels of the LPP1 gene (PLPP1) has been identified as one of twelve markers predictive of poor breast cancer survival [169].
The effects of LPP2 on tumor growth are very different from LPP1 and LPP3. LPP2 expression is increased in breast, lung and ovarian cancers where the expression of LPP1 and LPP3 are decreased [29]. Overexpressing LPP2 in fibroblasts leads to premature S-phase entry, and this effect is not seen with LPP1/3 [170]. Knockdown of LPP2 delays entry into S-phase [170] and impairs anchorage-dependent growth [171]. These observations provide preliminary evidence that increasing the low expressions of LPP1 and LPP3 and decreasing the high expression of LPP2 in cancer cells could provide novel targets for cancer therapy.
Our group discovered that one way of increasing LPP activity is through the use of tetracyclines, which increases the expression of the LPPs on the plasma membrane by enhancing LPP protein stability. This results in increased LPA clearance from circulation [28]. We then built on this finding to show that this doxycycline effect of increasing total ecto-activity of the LPPs in vivo also decreases circulating LPA concentrations, NF-κB translocation to the nucleus and IκB phosphorylation in breast cancer cells. This resulted in delayed breast tumor growth, deceased tumor macrophage infiltration and angiogenesis, and an overall reduction in the inflammatory milieu of the tumor environment [172]. Hence, tetracyclines and other potential activators of LPPs could become useful adjuvant therapies for decreasing signaling by LPA in cancers and other inflammatory diseases.

7. Pharmacological Targeting of LPA-Mediated Inflammation and Cancer Progression

As extracellular targets, ATX and LPA receptor inhibitors are ideal pharmacological targets for a plethora of conditions [173,174,175]. Because of redundancies through LPA receptor signaling, much of the research has focused on the development of ATX inhibitors, thereby impeding LPA signaling through all receptors [17]. Numerous ATX inhibitors that are potent in vivo have been developed over the last five years, and it was important to assess the safety of these inhibitors, given the role of LPA signaling in mediating tissue repair and remodeling. Fortunately, the bulk of ATX in murine models is dispensable in the adult organism [176]. Conditional genetic ablation of ATX in the adult mouse shows no deleterious changes in liver, pancreatic or kidney function, and in blood cell populations [176]. High doses of the competitive ATX inhibitor, PF8380, produced no weight loss or histopathological changes in 13 major organs [176]. In the context of sustained ATX inhibition, the production of small amounts of LPA by other enzymes such as phospholipase A1 is likely to provide sufficient LPA for physiological functions, but not enough to drive pathology [15]. Recent progress made in pharmacological targeting of LPA signaling has now led to the first human clinical trials, where the inhibitors appear to be well tolerated (Section 7.4).

7.1. Limiting Fibrosis through Potent ATX Inhibition

Recent studies in vivo have shown that ATX inhibition by PAT-048 limited bleomycin-induced dermal fibrosis in mice, via inhibition of an ATX-LPA-IL-6 amplification loop. Similar results were obtained in cultures of dermal fibroblasts from patients with scleroderma fibrosis [177]. This same group showed prevention of fibrosis in the same models with LPA1 deletion or antagonism [178]. Consistently, toclizumab, a monoclonal antibody against the IL-6 receptor, showed success against systemic sclerosis in Phase II clinical trials [179]. In other fibrotic models, the progression of NAFLD to non-alcoholic steatohepatitis and liver fibrosis has been blocked in animal models with PAT-505, another novel ATX inhibitor [180]. Similar results have been obtained with the ATX inhibitors AM063 and AM095, which mitigated fibrosis progression and reduced development of HCC by disrupting validated HCC risk gene signatures in cirrhosis-driven HCC rat models [181]. With regard to lung pathologies, while ATX concentrations increased in bronchoalveolar lavage fluid following lipopolysaccharide exposure, neither genetic nor pharmacologic targeting or ATX or LPA receptors reduced the severity of acute lung injuries [182]. However, there is ample evidence including work in clinical trials supporting the efficacy of LPA inhibition in chronic lung conditions. This suggests that pharmacological targeting of LPA signaling is most effective in the context of persistent inflammation [182].
Beyond fibrosis, another competitive ATX inhibitor, Compound 1, exhibits a dose-dependent decrease in joint pain in multiple rodent models [183]. This same group also demonstrated anti-inflammatory and analgesic properties with Compound 1 in models of inflammatory bowel disease and multiple sclerosis [72]. In addition to chemical competitive inhibitors, anti-ATX DNA aptamers have also been designed, and the aptamer RB014 exhibits efficacy in a mouse bleomycin-induced model of pulmonary fibrosis [184]. Bile acids also act as partial non-competitive inhibitors of ATX suggesting that steroid-based drugs may have ATX antagonism properties [185].

7.2. Improving Cancer Chemotherapy Effectiveness by Blocking LPA-Mediated Inflammatory Signaling

Given the considerable overlap between fibrotic and carcinoma pathways [136], LPA signaling inhibitors are expected to serve as useful adjuncts in cancer therapy, largely by mitigating the loss of efficacy for chemotherapy and radiotherapy via blockage of the pro-survival benefits of tumor-promoting inflammation. These include limiting the physical barriers to drug deposition in tumors by desmoplastic reactions, which encase tumors in dense extracellular matrices, and vessel hyperpermeability, which compromises tumor accumulation of chemotherapeutics [11]. We have demonstrated in both thyroid xenograft and orthotopic mouse breast cancer models that the potent ATX inhibitor, ONO-8430506, reduces tumor growth and metastasis by up to 70%. This involves reducing the expression of up to 20 inflammatory cytokines and chemokines that drive tumorigenesis themselves and by further increasing ATX secretion and LPA concentrations [137,144,147]. ATX inhibition in turn is synergistic with doxorubicin treatment, in part by decreasing the LPA-induced expression of the transcription factor nuclear factor (erythroid-derived 2)-like-2 (Nrf-2), which activates the anti-oxidant response element. This promotes the synthesis of antioxidant proteins and multidrug-resistance transporters [186]. LPA thereby protects cancer cells from oxidative damage from most common therapies including tamoxifen, paclitaxel, cisplatin, and possibly ionizing radiation through increasing the effects of transcription factors like Nrf-2 [187,188].

7.3. Sensitizing Tumors to Radiation Therapy by Blockage of LPA Signaling

LPA-mediated protection of cancer cells against radiation-induced cell death is partly mediated through LPA2 by stimulation of pro-survival kinase pathways that involve depleting cells of the proapototic signaling protein, Siva-1 [188]. ATM-mediated NF-κB activation occurs in response to radiation-induced DNA double-strand breaks, and plasma ATX and LPA levels increase. The resulting increase in LPA2 signaling accelerates the resolution of γH2AX histones [188]. Consequently, as predicted, LPA2-knockout mice are deficient in DNA damage repair mechanisms and have significant higher residual γH2AX histones after radiation exposure [188]. LPA also promotes the synthesis of numerous genes that promote DNA repair and this involves an increase in Nrf-2 expression [189,190]. Consequently, the Nrf-2 blockade has been proposed as an additional target for increasing the efficacy of radiation therapy [191]. Part of the increase in Nrf-2 expression could be mediated by the observed radiation-induced increased in the expression of ATX and LPA1 [192]. These observations that LPA protects against radiation-induced damage are compatible with studies showing that inhibition of ATX with BrP-LPA (also a pan-LPA receptor antagonist) or PF-8380 increased the sensitivity of heterotopic glioblastomas to radiation in mice [193,194].
Treatment of breast cancer after the surgical removal of the tumor following lumpectomy commonly involves exposure of the whole breast to about 16-25 fractions of 2 to 2.65 Gy of radiation to eliminate remaining cancer cells [195]. Exposure of rat and human adipose tissue to 0.25 to 5 Gy of γ-radiation increases the production of COX-2 and secretion of ATX and multiple inflammatory mediators [192]. This depends on DNA damage and the consequent activation of NF-κB and COX-2 signaling downstream of the activation of ATM serine/threonine kinase (ATM), ataxia telangiectasia and Rad3-related protein (ATR) and poly [ADP-ribose] polymerase 1 (PARP-1) [192] (Figure 3). Although radiation-induced activation of ATM increased the expressions of LPA1 and LPA2, the increases in ATX, COX-2 and inflammatory cytokines depends on activation of ATR [192]. These results show that these responses to radiation are activated by the accumulation of single-stranded DNA intermediates in double strand break resection during homologous recombination, rather than by double strand breaks themselves. The combined effects of radiation in causing inflammation in breast adipose tissue and thereby increasing the expression of ATX and signaling through LPA1 and LPA2 could increase the survival of residual breast cancer cells and decrease the effectiveness of the radiotherapy [192] (Figure 3).
Furthermore, radiation-induced fibrosis is a significant contributor to long-term sequelae for cancer survivors. Blockage of LPA1/3 with the antagonist VPC12249 ameliorated radiation-induced fibrosis and radiation pneumonitis in murine models [196,197]. Taken together with the evidence that blocking LPA signaling can decrease the progression of pulmonary fibrosis, it is likely that inhibitors of LPA signaling could have a significant potential in decreasing the adverse side effects of radiation therapy as well as improving the killing of residual cancer cells.

7.4. Inhibitors of LPA Signaling Entering into Clinical Trials

The number of inhibitors against LPA signaling has now reached a critical mass, such that the first ATX and LPA receptor inhibitors have entered clinical trials [16] (Table 1). Galapagos NV have tested the first ATX inhibitor, GLPG1690, currently in Phase II trials for idiopathic pulmonary fibrosis (IPF) [198] (Table 1). In a blelomycin-induced pulmonary fibrosis murine model, an analog of GLPG1690 reduced extracellular matrix deposition in the lung by nearly 50% and reduced the concentrations 18:2-LPA in bronchoalveolar lavage fluid by nearly 70% [198]. Similarly, Bristol-Myers Squibb has put BMS-986020, a LPA1 antagonist, into Phase II trials for idiopathic pulmonary fibrosis [16,199]. Sanofi SAR100842, a LPA1/3 antagonist, is in Phase II trials for the treatment of systemic sclerosis and related fibrotic diseases [16,199].
In a non-fibrotic disease model, Lpath, Inc. developed an LPA-directed monoclonal antibody which lowered lesion volume in mouse models of traumatic brain injury by decreasing IL-6 levels and conducted a Phase I trial of the antibody as Lpathomab [16,200] (Table 1).

8. Potential Applications for Inhibitors of LPA Signaling

So far, there is no approved cancer therapy that targets the ATX-LPA-inflammatory axis, but this area opens the prospect of exciting new approaches to cancer treatment. It is unlikely that inhibiting LPA signaling alone will prove effective as a cancer mono-therapy. Also, from a pragmatic and ethical point of view, it would impossible to test a new mono-therapy for blocking LPA signaling where an accepted therapy already exists. We envisage that strategies for blocking LPA signaling will be used to improve the efficacies of existing chemotherapies and radiotherapy regimens. This is based on preclinical evidence that blocking the effects of LPA increases the efficacy of taxanes, cisplatin, tamoxifen and doxorubicin in treating cancers [201]. Equally, LPA protects cells from radiation-induced cell death [192,202]. Therefore, blocking inflammation and the radiation–induced expression of ATX should prevent a loss of efficacy of radiotherapy during subsequent fractions of radiation [192,202]. Eventual clinical trials in the cancer field are likely to study LPA signaling inhibitors as adjuvants to existing standard-of-care treatments, looking at standard parameters that measure response to primary treatment. If these inhibitors as adjuvants show benefit, future studies might look at dose reductions of the primary treatment agents (either chemotherapy or radiotherapy), which would potentially decrease cancer therapy side effects.
One such group of major side effects is fatigue and the development of fibrosis secondary to radiotherapy. Fatigue is associated with the production of inflammatory cytokines during radiotherapy. Blocking LPA signaling should decrease this inflammatory response, which could also contribute to decreased scaring [203]. The positive effects of inhibiting ATX or LPA1 activation in treating pulmonary fibrosis provide encouraging evidence that these treatments could decrease radiation-induced fibrosis [196].
One of the major recent advances in cancer treatment has been the introduction of immunotherapy acting as checkpoint inhibitors. So far, there is no preclinical evidence that blocking LPA signaling will improve the efficacy of immunotherapies or combat its side effects. However, this is an intriguing possibility, since LPA and maladaptive inflammation contribute to immune evasion. The most frequent toxicities of immunotherapies include colitis, dermatitis, hepatitis, and pruritus [204,205], all conditions for which pre-clinical evidence exists that inhibition of LPA signaling might have therapeutic benefits.

9. Conclusions

In summary, we are at an exciting time where several therapeutics are in advanced clinical trials for blocking LPA signaling and inflammation. These agents are generally well tolerated and they could be tested as novel strategies for improving the effectiveness of existing cancer therapies. These approaches should be applicable to a wide variety of cancers since they target the tumor environment, which should be relatively independent of the specific mutation in the cancer cells. Overall, as mitigators of chronic inflammation, inhibitors of LPA signaling could become viable therapeutic modalities for preventing cancer initiation, maintaining the efficacy of chemotherapy and radiotherapy and prolonging remission.

Acknowledgments

The work was supported by the Canadian Cancer Society Research Institute, the Canadian Breast Cancer Foundation, the Women and Children’s Health Institute of the University of Alberta (WCHRI) and the Cancer Research Institute of Northern Alberta (CRINA).

Conflicts of Interest

David N. Brindley has received consultancy payments and/or research funding from Ono Pharmaceuticals Ltd. (Osaka, Japan) and Galapagos NV (Mechelen, Belgium). These groups had no role in the design of the review. The other authors declare no conflicts of interest.

References

  1. Balkwill, F.; Mantovani, A. Inflammation and cancer: Back to virchow? Lancet 2001, 357, 539–545. [Google Scholar] [CrossRef]
  2. Hanahan, D.; Weinberg, R.A. The hallmarks of cancer. Cell 2000, 100, 57–70. [Google Scholar] [CrossRef]
  3. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed]
  4. Dvorak, H.F. Tumors: Wounds that do not heal. Similarities between tumor stroma generation and wound healing. N. Engl. J. Med. 1986, 315, 1650–1659. [Google Scholar] [PubMed]
  5. Mantovani, A.; Sica, A. Macrophages, innate immunity and cancer: Balance, tolerance, and diversity. Curr. Opin. Immunol. 2010, 22, 231–237. [Google Scholar] [CrossRef] [PubMed]
  6. Mantovani, A.; Allavena, P.; Sica, A.; Balkwill, F. Cancer-related inflammation. Nature 2008, 454, 436–444. [Google Scholar] [CrossRef] [PubMed]
  7. Colotta, F.; Allavena, P.; Sica, A.; Garlanda, C.; Mantovani, A. Cancer-related inflammation, the seventh hallmark of cancer: Links to genetic instability. Carcinogenesis 2009, 30, 1073–1081. [Google Scholar] [CrossRef] [PubMed]
  8. Zeligs, K.P.; Neuman, M.K.; Annunziata, C.M. Molecular pathways: The balance between cancer and the immune system challenges the therapeutic specificity of targeting nuclear factor-κb signaling for cancer treatment. Clin. Cancer Res. 2016, 22, 4302–4308. [Google Scholar] [CrossRef] [PubMed]
  9. Hanahan, D.; Coussens, L.M. Accessories to the crime: Functions of cells recruited to the tumor microenvironment. Cancer Cell 2012, 21, 309–322. [Google Scholar] [CrossRef] [PubMed]
  10. Qian, B.Z.; Pollard, J.W. Macrophage diversity enhances tumor progression and metastasis. Cell 2010, 141, 39–51. [Google Scholar] [CrossRef] [PubMed]
  11. Gkretsi, V.; Zacharia, L.C.; Stylianopoulos, T. Targeting inflammation to improve tumor drug delivery. Trends Cancer 2017, 3, 621–630. [Google Scholar] [CrossRef] [PubMed]
  12. Kaiser, J. Profile: Bert vogelstein. Cancer genetics with an edge. Science 2012, 337, 282–284. [Google Scholar] [CrossRef] [PubMed]
  13. Roxburgh, C.S.; McMillan, D.C. Cancer and systemic inflammation: Treat the tumour and treat the host. Br. J. Cancer 2014, 110, 1409–1412. [Google Scholar] [CrossRef] [PubMed]
  14. Brindley, D.N.; Benesch, M.G.K.; Murph, M.M. Autotaxin—An enzymatic augmenter of malignant progression linked to inflammation. In Melanoma—Current Clinical Management and Future Therapeutics; InTech Open: London, UK, 2015; pp. 297–324. [Google Scholar] [CrossRef]
  15. Benesch, M.G.; Tang, X.; Venkatraman, G.; Bekele, R.T.; Brindley, D.N. Recent advances in targeting the autotaxin-lysophosphatidate-lipid phosphate phosphatase axis in vivo. J. Biomed. Res. 2016, 30, 272–284. [Google Scholar] [PubMed]
  16. Benesch, M.G.K.; Yang, Z.; Tang, X.; Meng, G.; Brindley, D.N. Lysophosphatidate signaling: The tumor microenvironment’s new nemesis. Trends Cancer 2017, 3, 748–752. [Google Scholar] [CrossRef] [PubMed]
  17. Benesch, M.G.K.; Ko, Y.M.; McMullen, T.P.W.; Brindley, D.N. Autotaxin in the crosshairs: Taking aim at cancer and other inflammatory conditions. FEBS Lett. 2014, 588, 2712–2727. [Google Scholar] [CrossRef] [PubMed]
  18. Van Meeteren, L.A.; Ruurs, P.; Stortelers, C.; Bouwman, P.; van Rooijen, M.A.; Pradere, J.P.; Pettit, T.R.; Wakelam, M.J.; Saulnier-Blache, J.S.; Mummery, C.L.; et al. Autotaxin, a secreted lysophospholipase D, is essential for blood vessel formation during development. Mol. Cell. Biol. 2006, 26, 5015–5022. [Google Scholar] [CrossRef] [PubMed]
  19. Tanaka, M.; Okudaira, S.; Kishi, Y.; Ohkawa, R.; Iseki, S.; Ota, M.; Noji, S.; Yatomi, Y.; Aoki, J.; Arai, H. Autotaxin stabilizes blood vessels and is required for embryonic vasculature by producing lysophosphatidic acid. J. Biol. Chem. 2006, 281, 25822–25830. [Google Scholar] [CrossRef] [PubMed]
  20. Gierse, J.; Thorarensen, A.; Beltey, K.; Bradshaw-Pierce, E.; Cortes-Burgos, L.; Hall, T.; Johnston, A.; Murphy, M.; Nemirovskiy, O.; Ogawa, S.; et al. A novel autotaxin inhibitor reduces lysophosphatidic acid levels in plasma and the site of inflammation. J. Pharmacol. Exp. Ther. 2010, 334, 310–317. [Google Scholar] [CrossRef] [PubMed]
  21. Dusaulcy, R.; Rancoule, C.; Gres, S.; Wanecq, E.; Colom, A.; Guigne, C.; van Meeteren, L.A.; Moolenaar, W.H.; Valet, P.; Saulnier-Blache, J.S. Adipose-specific disruption of autotaxin enhances nutritional fattening and reduces plasma lysophosphatidic acid. J. Lipid Res. 2011, 52, 1247–1255. [Google Scholar] [CrossRef] [PubMed]
  22. Quan, M.; Cui, J.J.; Feng, X.; Huang, Q. The critical role and potential target of the autotaxin/lysophosphatidate axis in pancreatic cancer. Tumour Biol. 2017, 39. [Google Scholar] [CrossRef] [PubMed]
  23. Oka, S.; Ota, R.; Shima, M.; Yamashita, A.; Sugiura, T. GPR35 is a novel lysophosphatidic acid receptor. Biochem. Biophys. Res. Commun. 2010, 395, 232–237. [Google Scholar] [CrossRef] [PubMed]
  24. Tabata, K.; Baba, K.; Shiraishi, A.; Ito, M.; Fujita, N. The orphan GPCR GPR87 was deorphanized and shown to be a lysophosphatidic acid receptor. Biochem. Biophys. Res. Commun. 2007, 363, 861–866. [Google Scholar] [CrossRef] [PubMed]
  25. Ochiai, S.; Furuta, D.; Sugita, K.; Taniura, H.; Fujita, N. GPR87 mediates lysophosphatidic acid-induced colony dispersal in A431 cells. Eur. J. Pharmacol. 2013, 715, 15–20. [Google Scholar] [CrossRef] [PubMed]
  26. Murakami, M.; Shiraishi, A.; Tabata, K.; Fujita, N. Identification of the orphan GPCR, P2Y(10) receptor as the sphingosine-1-phosphate and lysophosphatidic acid receptor. Biochem. Biophys. Res. Commun. 2008, 371, 707–712. [Google Scholar] [CrossRef] [PubMed]
  27. Rai, V.; Toure, F.; Chitayat, S.; Pei, R.; Song, F.; Li, Q.; Zhang, J.; Rosario, R.; Ramasamy, R.; Chazin, W.J.; et al. Lysophosphatidic acid targets vascular and oncogenic pathways via rage signaling. J. Exp. Med. 2012, 209, 2339–2350. [Google Scholar] [CrossRef] [PubMed]
  28. Tang, X.; Zhao, Y.Y.; Dewald, J.; Curtis, J.M.; Brindley, D.N. Tetracyclines increase lipid phosphate phosphatase expression on plasma membranes and turnover of plasma lysophosphatidate. J. Lipid Res. 2016, 57, 597–606. [Google Scholar] [CrossRef] [PubMed]
  29. Tang, X.; Benesch, M.G.; Brindley, D.N. Lipid phosphate phosphatases and their roles in mammalian physiology and pathology. J. Lipid Res. 2015, 56, 2048–2060. [Google Scholar] [CrossRef] [PubMed]
  30. Aikawa, S.; Hashimoto, T.; Kano, K.; Aoki, J. Lysophosphatidic acid as a lipid mediator with multiple biological actions. J. Biochem. 2015, 157, 81–89. [Google Scholar] [CrossRef] [PubMed]
  31. Yung, Y.C.; Stoddard, N.C.; Chun, J. LPA receptor signaling: Pharmacology, physiology, and pathophysiology. J. Lipid Res. 2014, 55, 1192–1214. [Google Scholar] [CrossRef] [PubMed]
  32. Riaz, A.; Huang, Y.; Johansson, S. G-protein-coupled lysophosphatidic acid receptors and their regulation of AKT signaling. Int. J. Mol. Sci. 2016, 17, 215. [Google Scholar] [CrossRef] [PubMed]
  33. Mueller, P.; Ye, S.; Morris, A.; Smyth, S.S. Lysophospholipid mediators in the vasculature. Exp. Cell Res. 2015, 333, 190–194. [Google Scholar] [CrossRef] [PubMed]
  34. Masse, K.; Bhamra, S.; Allsop, G.; Dale, N.; Jones, E.A. Ectophosphodiesterase/nucleotide phosphohydrolase (Enpp) nucleotidases: Cloning, conservation and developmental restriction. Int. J. Dev. Biol. 2010, 54, 181–193. [Google Scholar] [CrossRef] [PubMed]
  35. Federico, L.; Jeong, K.J.; Vellano, C.P.; Mills, G.B. Autotaxin, a lysophospholipase D with pleomorphic effects in oncogenesis and cancer progression. J. Lipid Res. 2016, 57, 25–35. [Google Scholar] [CrossRef] [PubMed]
  36. Nishimasu, H.; Okudaira, S.; Hama, K.; Mihara, E.; Dohmae, N.; Inoue, A.; Ishitani, R.; Takagi, J.; Aoki, J.; Nureki, O. Crystal structure of autotaxin and insight into GPCR activation by lipid mediators. Nat. Struct. Mol. Biol. 2011, 18, 205–212. [Google Scholar] [CrossRef] [PubMed]
  37. Lyu, L.; Wang, B.; Xiong, C.; Zhang, X.; Zhang, X.; Zhang, J. Selective export of autotaxin from the endoplasmic reticulum. J. Biol. Chem. 2017, 292, 7011–7022. [Google Scholar] [CrossRef] [PubMed]
  38. Jansen, S.; Stefan, C.; Creemers, J.W.; Waelkens, E.; Van Eynde, A.; Stalmans, W.; Bollen, M. Proteolytic maturation and activation of autotaxin (NPP2), a secreted metastasis-enhancing lysophospholipase D. J. Cell Sci. 2005, 118, 3081–3089. [Google Scholar] [CrossRef] [PubMed]
  39. Tokumura, A.; Majima, E.; Kariya, Y.; Tominaga, K.; Kogure, K.; Yasuda, K.; Fukuzawa, K. Identification of human plasma lysophospholipase D, a lysophosphatidic acid-producing enzyme, as autotaxin, a multifunctional phosphodiesterase. J. Biol. Chem. 2002, 277, 39436–39442. [Google Scholar] [CrossRef] [PubMed]
  40. Umezu-Goto, M.; Kishi, Y.; Taira, A.; Hama, K.; Dohmae, N.; Takio, K.; Yamori, T.; Mills, G.B.; Inoue, K.; Aoki, J.; et al. Autotaxin has lysophospholipase D activity leading to tumor cell growth and motility by lysophosphatidic acid production. J. Cell Biol. 2002, 158, 227–233. [Google Scholar] [CrossRef] [PubMed]
  41. Stracke, M.L.; Krutzsch, H.C.; Unsworth, E.J.; Arestad, A.; Cioce, V.; Schiffmann, E.; Liotta, L.A. Identification, purification, and partial sequence analysis of autotaxin, a novel motility-stimulating protein. J. Biol. Chem. 1992, 267, 2524–2529. [Google Scholar] [PubMed]
  42. Perrakis, A.; Moolenaar, W.H. Autotaxin: Structure-function and signaling. J. Lipid Res. 2014, 55, 1010–1018. [Google Scholar] [CrossRef] [PubMed]
  43. Tabchy, A.; Tigyi, G.; Mills, G.B. Location, location, location: A crystal-clear view of autotaxin saturating LPA receptors. Nat. Struct. Mol. Biol. 2011, 18, 117–118. [Google Scholar] [CrossRef] [PubMed]
  44. Aikawa, S.; Kano, K.; Inoue, A.; Aoki, J. Proliferation of mouse endometrial stromal cells in culture is highly sensitive to lysophosphatidic acid signaling. Biochem. Biophys. Res. Commun. 2017, 484, 202–208. [Google Scholar] [CrossRef] [PubMed]
  45. Sinderewicz, E.; Grycmacher, K.; Boruszewska, D.; Kowalczyk-Zieba, I.; Staszkiewicz, J.; Slezak, T.; Woclawek-Potocka, I. Bovine ovarian follicular growth and development correlate with lysophosphatidic acid expression. Theriogenology 2017, 106, 1–14. [Google Scholar] [CrossRef] [PubMed]
  46. Sinderewicz, E.; Grycmacher, K.; Boruszewska, D.; Kowalczyk-Zieba, I.; Staszkiewicz, J.; Slezak, T.; Woclawek-Potocka, I. Expression of factors involved in apoptosis and cell survival is correlated with enzymes synthesizing lysophosphatidic acid and its receptors in granulosa cells originating from different types of bovine ovarian follicles. Reprod. Biol. Endocrinol. 2017, 15, 72. [Google Scholar] [CrossRef] [PubMed]
  47. Masuda, K.; Haruta, S.; Orino, K.; Kawaminami, M.; Kurusu, S. Autotaxin as a novel, tissue-remodeling-related factor in regressing corpora lutea of cycling rats. FEBS J. 2013, 280, 6600–6612. [Google Scholar] [CrossRef] [PubMed]
  48. Ichikawa, M.; Nagamatsu, T.; Schust, D.J.; Kawai-Iwasawa, Y.; Kawana, K.; Yamashita, T.; Osuga, Y.; Aoki, J.; Yatomi, Y.; Fujii, T. Placental autotaxin expression is diminished in women with pre-eclampsia. J. Obstet. Gynaecol. Res. 2015, 41, 1406–1411. [Google Scholar] [CrossRef] [PubMed]
  49. Erenel, H.; Yilmaz, N.; Cift, T.; Bulut, B.; Sozen, I.; Aslan Cetin, B.; Gezer, A.; Ekmekci, H.; Kaya, B.; Tuten, A. Maternal serum autotaxin levels in early- and late-onset preeclampsia. Hypertens. Pregnancy 2017, 36, 310–314. [Google Scholar] [CrossRef] [PubMed]
  50. Ferry, G.; Giganti, A.; Coge, F.; Bertaux, F.; Thiam, K.; Boutin, J.A. Functional invalidation of the autotaxin gene by a single amino acid mutation in mouse is lethal. FEBS Lett. 2007, 581, 3572–3578. [Google Scholar] [CrossRef] [PubMed]
  51. Fotopoulou, S.; Oikonomou, N.; Grigorieva, E.; Nikitopoulou, I.; Paparountas, T.; Thanassopoulou, A.; Zhao, Z.; Xu, Y.; Kontoyiannis, D.L.; Remboutsika, E.; et al. ATX expression and LPA signalling are vital for the development of the nervous system. Dev. Biol. 2010, 339, 451–464. [Google Scholar] [CrossRef] [PubMed]
  52. Yuelling, L.W.; Waggener, C.T.; Afshari, F.S.; Lister, J.A.; Fuss, B. Autotaxin/ENPP2 regulates oligodendrocyte differentiation in vivo in the developing zebrafish hindbrain. Glia 2012, 60, 1605–1618. [Google Scholar] [CrossRef] [PubMed]
  53. Wheeler, N.A.; Lister, J.A.; Fuss, B. The autotaxin-lysophosphatidic acid axis modulates histone acetylation and gene expression during oligodendrocyte differentiation. J. Neurosci. 2015, 35, 11399–11414. [Google Scholar] [CrossRef] [PubMed]
  54. Lai, S.L.; Yao, W.L.; Tsao, K.C.; Houben, A.J.; Albers, H.M.; Ovaa, H.; Moolenaar, W.H.; Lee, S.J. Autotaxin/Lpar3 signaling regulates Kupffer’s vesicle formation and left-right asymmetry in zebrafish. Development 2012, 139, 4439–4448. [Google Scholar] [CrossRef] [PubMed]
  55. Yukiura, H.; Kano, K.; Kise, R.; Inoue, A.; Aoki, J. Autotaxin overexpression causes embryonic lethality and vascular defects. PLoS ONE 2015, 10, e0126734. [Google Scholar] [CrossRef] [PubMed]
  56. Escalante-Alcalde, D.; Hernandez, L.; Le Stunff, H.; Maeda, R.; Lee, H.S.; Cheng, G., Jr.; Sciorra, V.A.; Daar, I.; Spiegel, S.; Morris, A.J.; et al. The lipid phosphatase LPP3 regulates extra-embryonic vasculogenesis and axis patterning. Development 2003, 130, 4623–4637. [Google Scholar] [CrossRef] [PubMed]
  57. Chatterjee, I.; Baruah, J.; Lurie, E.E.; Wary, K.K. Endothelial lipid phosphate phosphatase-3 deficiency that disrupts the endothelial barrier function is a modifier of cardiovascular development. Cardiovasc. Res. 2016, 111, 105–118. [Google Scholar] [CrossRef] [PubMed]
  58. Frisca, F.; Colquhoun, D.; Goldshmit, Y.; Anko, M.L.; Pebay, A.; Kaslin, J. Role of ectonucleotide pyrophosphatase/phosphodiesterase 2 in the midline axis formation of zebrafish. Sci. Rep. 2016, 6, 37678. [Google Scholar] [CrossRef] [PubMed]
  59. Nakanaga, K.; Hama, K.; Kano, K.; Sato, T.; Yukiura, H.; Inoue, A.; Saigusa, D.; Tokuyama, H.; Tomioka, Y.; Nishina, H.; et al. Overexpression of autotaxin, a lysophosphatidic acid-producing enzyme, enhances cardia bifida induced by hypo-sphingosine-1-phosphate signaling in zebrafish embryo. J. Biochem. 2014, 155, 235–241. [Google Scholar] [CrossRef] [PubMed]
  60. Nakanaga, K.; Hama, K.; Aoki, J. Autotaxin—An LPA producing enzyme with diverse functions. J. Biochem. 2010, 148, 13–24. [Google Scholar] [CrossRef] [PubMed]
  61. Knowlden, S.; Georas, S.N. The autotaxin-LPA axis emerges as a novel regulator of lymphocyte homing and inflammation. J. Immunol. 2014, 192, 851–857. [Google Scholar] [CrossRef] [PubMed]
  62. Ray, R.; Rai, V. Lysophosphatidic acid converts monocytes into macrophages in both mice and humans. Blood 2017, 129, 1177–1183. [Google Scholar] [CrossRef] [PubMed]
  63. Fulkerson, Z.; Wu, T.; Sunkara, M.; Kooi, C.V.; Morris, A.J.; Smyth, S.S. Binding of autotaxin to integrins localizes lysophosphatidic acid production to platelets and mammalian cells. J. Biol. Chem. 2011, 286, 34654–34663. [Google Scholar] [CrossRef] [PubMed]
  64. Song, J.; Guan, M.; Zhao, Z.; Zhang, J. Type I interferons function as autocrine and paracrine factors to induce autotaxin in response to TLR activation. PLoS ONE 2015, 10, e0136629. [Google Scholar] [CrossRef] [PubMed]
  65. Katakai, T.; Kondo, N.; Ueda, Y.; Kinashi, T. Autotaxin produced by stromal cells promotes LFA-1-independent and Rho-dependent interstitial T cell motility in the lymph node paracortex. J. Immunol. 2014, 193, 617–626. [Google Scholar] [CrossRef] [PubMed]
  66. Knowlden, S.A.; Capece, T.; Popovic, M.; Chapman, T.J.; Rezaee, F.; Kim, M.; Georas, S.N. Regulation of T cell motility in vitro and in vivo by LPA and LPA2. PLoS ONE 2014, 9, e101655. [Google Scholar] [CrossRef] [PubMed]
  67. Takeda, A.; Kobayashi, D.; Aoi, K.; Sasaki, N.; Sugiura, Y.; Igarashi, H.; Tohya, K.; Inoue, A.; Hata, E.; Akahoshi, N.; et al. Fibroblastic reticular cell-derived lysophosphatidic acid regulates confined intranodal T-cell motility. eLife 2016, 5, e10561. [Google Scholar] [CrossRef] [PubMed]
  68. Kime, C.; Sakaki-Yumoto, M.; Goodrich, L.; Hayashi, Y.; Sami, S.; Derynck, R.; Asahi, M.; Panning, B.; Yamanaka, S.; Tomoda, K. Autotaxin-mediated lipid signaling intersects with LIF and BMP signaling to promote the naive pluripotency transcription factor program. Proc. Natl. Acad. Sci. USA 2016, 113, 12478–12483. [Google Scholar] [CrossRef] [PubMed]
  69. Schmitz, K.; Brunkhorst, R.; de Bruin, N.; Mayer, C.A.; Haussler, A.; Ferreiros, N.; Schiffmann, S.; Parnham, M.J.; Tunaru, S.; Chun, J.; et al. Dysregulation of lysophosphatidic acids in multiple sclerosis and autoimmune encephalomyelitis. Acta Neuropathol. Commun. 2017, 5, 42. [Google Scholar] [CrossRef] [PubMed]
  70. Raddatz, B.B.; Hansmann, F.; Spitzbarth, I.; Kalkuhl, A.; Deschl, U.; Baumgartner, W.; Ulrich, R. Transcriptomic meta-analysis of multiple sclerosis and its experimental models. PLoS ONE 2014, 9, e86643. [Google Scholar] [CrossRef] [PubMed]
  71. Nogaroli, L.; Yuelling, L.M.; Dennis, J.; Gorse, K.; Payne, S.G.; Fuss, B. Lysophosphatidic acid can support the formation of membranous structures and an increase in MBP mRNA levels in differentiating oligodendrocytes. Neurochem. Res. 2009, 34, 182–193. [Google Scholar] [CrossRef] [PubMed]
  72. Thirunavukkarasu, K.; Tan, B.; Swearingen, C.A.; Rocha, G.; Bui, H.H.; McCann, D.J.; Jones, S.B.; Norman, B.H.; Pfeifer, L.A.; Saha, J.K. Pharmacological characterization of a potent inhibitor of autotaxin in animal models of inflammatory bowel disease and multiple sclerosis. J. Pharmacol. Exp. Ther. 2016, 359, 207–214. [Google Scholar] [CrossRef] [PubMed]
  73. Nishioka, T.; Arima, N.; Kano, K.; Hama, K.; Itai, E.; Yukiura, H.; Kise, R.; Inoue, A.; Kim, S.H.; Solnica-Krezel, L.; et al. ATX-LPA1 axis contributes to proliferation of chondrocytes by regulating fibronectin assembly leading to proper cartilage formation. Sci. Rep. 2016, 6, 23433. [Google Scholar] [CrossRef] [PubMed]
  74. George, J.; Headen, K.V.; Ogunleye, A.O.; Perry, G.A.; Wilwerding, T.M.; Parrish, L.C.; McVaney, T.P.; Mattson, J.S.; Cerutis, D.R. Lysophosphatidic acid signals through specific lysophosphatidic acid receptor subtypes to control key regenerative responses of human gingival and periodontal ligament fibroblasts. J. Periodontol. 2009, 80, 1338–1347. [Google Scholar] [CrossRef] [PubMed]
  75. Cerutis, D.R.; Weston, M.D.; Ogunleye, A.O.; McVaney, T.P.; Miyamoto, T. Lysophosphatidic acid (LPA) 18:1 transcriptional regulation of primary human gingival fibroblasts. Genom. Data 2014, 2, 375–377. [Google Scholar] [CrossRef] [PubMed]
  76. Cerutis, D.R.; Weston, M.D.; Alnouti, Y.; Bathena, S.P.; Nunn, M.E.; Ogunleye, A.O.; McVaney, T.P.; Headen, K.V.; Miyamoto, T. A major human oral lysophosphatidic acid species, LPA 18:1, regulates novel genes in human gingival fibroblasts. J. Periodontol. 2015, 86, 713–725. [Google Scholar] [CrossRef] [PubMed]
  77. Benesch, M.G.K.; Zhao, Y.Y.; Curtis, J.M.; McMullen, T.P.; Brindley, D.N. Regulation of autotaxin expression and secretion by lysophosphatidate and sphingosine 1-phosphate. J. Lipid Res. 2015, 56, 1134–1144. [Google Scholar] [CrossRef] [PubMed]
  78. Jansen, S.; Andries, M.; Vekemans, K.; Vanbilloen, H.; Verbruggen, A.; Bollen, M. Rapid clearance of the circulating metastatic factor autotaxin by the scavenger receptors of liver sinusoidal endothelial cells. Cancer Lett. 2009, 284, 216–221. [Google Scholar] [CrossRef] [PubMed]
  79. Nikitopoulou, I.; Oikonomou, N.; Karouzakis, E.; Sevastou, I.; Nikolaidou-Katsaridou, N.; Zhao, Z.; Mersinias, V.; Armaka, M.; Xu, Y.; Masu, M.; et al. Autotaxin expression from synovial fibroblasts is essential for the pathogenesis of modeled arthritis. J. Exp. Med. 2012, 209, 925–933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Oikonomou, N.; Mouratis, M.A.; Tzouvelekis, A.; Kaffe, E.; Valavanis, C.; Vilaras, G.; Karameris, A.; Prestwich, G.D.; Bouros, D.; Aidinis, V. Pulmonary autotaxin expression contributes to the pathogenesis of pulmonary fibrosis. Am. J. Respir. Cell Mol. Biol. 2012, 47, 566–574. [Google Scholar] [CrossRef] [PubMed]
  81. Kaffe, E.; Katsifa, A.; Xylourgidis, N.; Ninou, I.; Zannikou, M.; Harokopos, V.; Foka, P.; Dimitriadis, A.; Evangelou, K.; Moulas, A.N.; et al. Hepatocyte autotaxin expression promotes liver fibrosis and cancer. Hepatology 2017, 65, 1369–1383. [Google Scholar] [CrossRef] [PubMed]
  82. Choi, J.; Cha, Y.J.; Koo, J.S. Adipocyte biology in breast cancer: From silent bystander to active facilitator. Prog. Lipid Res. 2017, 69, 11–20. [Google Scholar] [CrossRef] [PubMed]
  83. Rancoule, C.; Dusaulcy, R.; Treguer, K.; Gres, S.; Guigne, C.; Quilliot, D.; Valet, P.; Saulnier-Blache, J.S. Depot-specific regulation of autotaxin with obesity in human adipose tissue. J. Physiol. Biochem. 2012, 68, 635–644. [Google Scholar] [CrossRef] [PubMed]
  84. Nakamura, R.; Okura, T.; Fujioka, Y.; Sumi, K.; Matsuzawa, K.; Izawa, S.; Ueta, E.; Kato, M.; Taniguchi, S.I.; Yamamoto, K. Serum fatty acid-binding protein 4 (FABP4) concentration is associated with insulin resistance in peripheral tissues, a clinical study. PLoS ONE 2017, 12, e0179737. [Google Scholar] [CrossRef] [PubMed]
  85. D’Souza, K.; Kane, D.A.; Touaibia, M.; Kershaw, E.E.; Pulinilkunnil, T.; Kienesberger, P.C. Autotaxin is regulated by glucose and insulin in adipocytes. Endocrinology 2017, 158, 791–803. [Google Scholar] [CrossRef] [PubMed]
  86. Lopez-Jaramillo, P.; Gomez-Arbelaez, D.; Lopez-Lopez, J.; Lopez-Lopez, C.; Martinez-Ortega, J.; Gomez-Rodriguez, A.; Triana-Cubillos, S. The role of leptin/adiponectin ratio in metabolic syndrome and diabetes. Horm. Mol. Biol. Clin. Investig. 2014, 18, 37–45. [Google Scholar] [CrossRef] [PubMed]
  87. Stern, J.H.; Rutkowski, J.M.; Scherer, P.E. Adiponectin, leptin, and fatty acids in the maintenance of metabolic homeostasis through adipose tissue crosstalk. Cell Metab. 2016, 23, 770–784. [Google Scholar] [CrossRef] [PubMed]
  88. Fruhbeck, G.; Catalan, V.; Rodriguez, A.; Ramirez, B.; Becerril, S.; Salvador, J.; Portincasa, P.; Colina, I.; Gomez-Ambrosi, J. Involvement of the leptin-adiponectin axis in inflammation and oxidative stress in the metabolic syndrome. Sci. Rep. 2017, 7, 6619. [Google Scholar] [CrossRef] [PubMed]
  89. Reeves, V.L.; Trybula, J.S.; Wills, R.C.; Goodpaster, B.H.; Dube, J.J.; Kienesberger, P.C.; Kershaw, E.E. Serum Autotaxin/ENPP2 correlates with insulin resistance in older humans with obesity. Obesity 2015, 23, 2371–2376. [Google Scholar] [CrossRef] [PubMed]
  90. Federico, L.; Ren, H.; Mueller, P.A.; Wu, T.; Liu, S.; Popovic, J.; Blalock, E.M.; Sunkara, M.; Ovaa, H.; Albers, H.M.; et al. Autotaxin and its product lysophosphatidic acid suppress brown adipose differentiation and promote diet-induced obesity in mice. Mol. Endocrinol. 2012, 26, 786–797. [Google Scholar] [CrossRef] [PubMed]
  91. Fayyaz, S.; Japtok, L.; Schumacher, F.; Wigger, D.; Schulz, T.J.; Haubold, K.; Gulbins, E.; Voller, H.; Kleuser, B. Lysophosphatidic acid inhibits insulin signaling in primary rat hepatocytes via the LPA3 receptor subtype and is increased in obesity. Cell. Physiol. Biochem. 2017, 43, 445–456. [Google Scholar] [CrossRef] [PubMed]
  92. Sun, S.; Wang, R.; Song, J.; Guan, M.; Li, N.; Zhang, X.; Zhao, Z.; Zhang, J. Blocking gp130 signaling suppresses autotaxin expression in adipocytes and improves insulin sensitivity in diet-induced obesity. J. Lipid Res. 2017, 58, 2102–2113. [Google Scholar] [CrossRef] [PubMed]
  93. Nishimura, S.; Nagasaki, M.; Okudaira, S.; Aoki, J.; Ohmori, T.; Ohkawa, R.; Nakamura, K.; Igarashi, K.; Yamashita, H.; Eto, K.; et al. ENPP2 contributes to adipose tissue expansion and insulin resistance in diet-induced obesity. Diabetes 2014, 63, 4154–4164. [Google Scholar] [CrossRef] [PubMed]
  94. Shimizu, M.; Furuichi, K.; Toyama, T.; Yamahana, J.; Ohkawa, R.; Igarashi, K.; Aoki, J.; Kaneko, S.; Yatomi, Y.; Wada, T. Serum autotaxin levels are associated with proteinuria and kidney lesions in japanese type 2 diabetic patients with biopsy-proven diabetic nephropathy. Intern. Med. 2016, 55, 215–221. [Google Scholar] [CrossRef] [PubMed]
  95. Valdes-Rives, S.A.; Gonzalez-Arenas, A. Autotaxin-lysophosphatidic acid: From inflammation to cancer development. Mediat. Inflamm. 2017, 2017, 9173090. [Google Scholar] [CrossRef] [PubMed]
  96. Gu, C.; Wang, F.; Zhao, Z.; Wang, H.; Cong, X.; Chen, X. Lysophosphatidic acid is associated with atherosclerotic plaque instability by regulating NF-κb dependent matrix metalloproteinase-9 expression via LPA2 in macrophages. Front. Physiol. 2017, 8, 266. [Google Scholar] [CrossRef] [PubMed]
  97. Nsaibia, M.J.; Mahmut, A.; Boulanger, M.C.; Arsenault, B.J.; Bouchareb, R.; Simard, S.; Witztum, J.L.; Clavel, M.A.; Pibarot, P.; Bosse, Y.; et al. Autotaxin interacts with lipoprotein(a) and oxidized phospholipids in predicting the risk of calcific aortic valve stenosis in patients with coronary artery disease. J. Intern. Med. 2016, 280, 509–517. [Google Scholar] [CrossRef] [PubMed]
  98. Mathieu, P.; Arsenault, B.J.; Boulanger, M.C.; Bosse, Y.; Koschinsky, M.L. Pathobiology of Lp(a) in calcific aortic valve disease. Expert Rev. Cardiovasc. Ther. 2017, 15, 797–807. [Google Scholar] [CrossRef] [PubMed]
  99. Bouchareb, R.; Mahmut, A.; Nsaibia, M.J.; Boulanger, M.C.; Dahou, A.; Lepine, J.L.; Laflamme, M.H.; Hadji, F.; Couture, C.; Trahan, S.; et al. Autotaxin derived from lipoprotein(a) and valve interstitial cells promotes inflammation and mineralization of the aortic valve. Circulation 2015, 132, 677–690. [Google Scholar] [CrossRef] [PubMed]
  100. Smyth, S.S.; Mueller, P.; Yang, F.; Brandon, J.A.; Morris, A.J. Arguing the case for the autotaxin-lysophosphatidic acid-lipid phosphate phosphatase 3-signaling nexus in the development and complications of atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2014, 34, 479–486. [Google Scholar] [CrossRef] [PubMed]
  101. Touat-Hamici, Z.; Weidmann, H.; Blum, Y.; Proust, C.; Durand, H.; Iannacci, F.; Codoni, V.; Gaignard, P.; Therond, P.; Civelek, M.; et al. Role of lipid phosphate phosphatase 3 in human aortic endothelial cell function. Cardiovasc. Res. 2016, 112, 702–713. [Google Scholar] [CrossRef] [PubMed]
  102. Panchatcharam, M.; Salous, A.K.; Brandon, J.; Miriyala, S.; Wheeler, J.; Patil, P.; Sunkara, M.; Morris, A.J.; Escalante-Alcalde, D.; Smyth, S.S. Mice with targeted inactivation of Ppap2b in endothelial and hematopoietic cells display enhanced vascular inflammation and permeability. Arterioscler. Thromb. Vasc. Biol. 2014, 34, 837–845. [Google Scholar] [CrossRef] [PubMed]
  103. Parimisetty, A.; Dorsemans, A.C.; Awada, R.; Ravanan, P.; Diotel, N.; Lefebvre d’Hellencourt, C. Secret talk between adipose tissue and central nervous system via secreted factors-an emerging frontier in the neurodegenerative research. J. Neuroinflamm. 2016, 13, 67. [Google Scholar] [CrossRef] [PubMed]
  104. McLimans, K.E.; Willette, A.A. Autotaxin is related to metabolic dysfunction and predicts alzheimer’s disease outcomes. J. Alzheimers Dis. 2017, 56, 403–413. [Google Scholar] [CrossRef] [PubMed]
  105. Velasco, M.; O’Sullivan, C.; Sheridan, G.K. Lysophosphatidic acid receptors (LPARs): Potential targets for the treatment of neuropathic pain. Neuropharmacology 2017, 113, 608–617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. McDougall, J.J.; Albacete, S.; Schuelert, N.; Mitchell, P.G.; Lin, C.; Oskins, J.L.; Bui, H.H.; Chambers, M.G. Lysophosphatidic acid provides a missing link between osteoarthritis and joint neuropathic pain. Osteoarthr. Cartil. 2017, 25, 926–934. [Google Scholar] [CrossRef] [PubMed]
  107. Herr, D.R.; Ong, J.H.; Ong, W.Y. Potential therapeutic applications for inhibitors of autotaxin, a bioactive lipid-producing lysophospholipase D, in disorders affecting the nervous system. ACS Chem. Neurosci. 2018. [Google Scholar] [CrossRef] [PubMed]
  108. Mabey, T.; Taleongpong, P.; Udomsinprasert, W.; Jirathanathornnukul, N.; Honsawek, S. Plasma and synovial fluid autotaxin correlate with severity in knee osteoarthritis. Clin. Chim. Acta 2015, 444, 72–77. [Google Scholar] [CrossRef] [PubMed]
  109. Datta, P.; Zhang, Y.; Parousis, A.; Sharma, A.; Rossomacha, E.; Endisha, H.; Wu, B.; Kacprzak, I.; Mahomed, N.N.; Gandhi, R.; et al. High-fat diet-induced acceleration of osteoarthritis is associated with a distinct and sustained plasma metabolite signature. Sci. Rep. 2017, 7, 8205. [Google Scholar] [CrossRef] [PubMed]
  110. Orosa, B.; Garcia, S.; Conde, C. The autotaxin-lysophosphatidic acid pathway in pathogenesis of rheumatoid arthritis. Eur. J. Pharmacol. 2015, 765, 228–233. [Google Scholar] [CrossRef] [PubMed]
  111. Chu, X.; Wei, X.; Lu, S.; He, P. Autotaxin-LPA receptor axis in the pathogenesis of lung diseases. Int. J. Clin. Exp. Med. 2015, 8, 17117–17122. [Google Scholar] [PubMed]
  112. Tager, A.M.; LaCamera, P.; Shea, B.S.; Campanella, G.S.; Selman, M.; Zhao, Z.; Polosukhin, V.; Wain, J.; Karimi-Shah, B.A.; Kim, N.D.; et al. The lysophosphatidic acid receptor LPA1 links pulmonary fibrosis to lung injury by mediating fibroblast recruitment and vascular leak. Nat. Med. 2008, 14, 45–54. [Google Scholar] [CrossRef] [PubMed]
  113. Black, K.E.; Berdyshev, E.; Bain, G.; Castelino, F.V.; Shea, B.S.; Probst, C.K.; Fontaine, B.A.; Bronova, I.; Goulet, L.; Lagares, D.; et al. Autotaxin activity increases locally following lung injury, but is not required for pulmonary lysophosphatidic acid production or fibrosis. FASEB J. 2016, 30, 2435–2450. [Google Scholar] [CrossRef] [PubMed]
  114. Cao, P.; Aoki, Y.; Badri, L.; Walker, N.M.; Manning, C.M.; Lagstein, A.; Fearon, E.R.; Lama, V.N. Autocrine lysophosphatidic acid signaling activates β-catenin and promotes lung allograft fibrosis. J. Clin. Investig. 2017, 127, 1517–1530. [Google Scholar] [CrossRef] [PubMed]
  115. Nowak-Machen, M.; Lange, M.; Exley, M.; Wu, S.; Usheva, A.; Robson, S.C. Lysophosphatidic acid generation by pulmonary NKT cell ENPP-2/autotaxin exacerbates hyperoxic lung injury. Purinergic Signal. 2015, 11, 455–461. [Google Scholar] [CrossRef] [PubMed]
  116. Ackerman, S.J.; Park, G.Y.; Christman, J.W.; Nyenhuis, S.; Berdyshev, E.; Natarajan, V. Polyunsaturated lysophosphatidic acid as a potential asthma biomarker. Biomark. Med. 2016, 10, 123–135. [Google Scholar] [CrossRef] [PubMed]
  117. Park, E.; Kim, D.; Lee, S.M.; Jun, H.S. Inhibition of lysophosphatidic acid receptor ameliorates Sjogren’s syndrome in NOD mice. Oncotarget 2017, 8, 27240–27251. [Google Scholar] [PubMed]
  118. Moisseiev, E.; Loewenstein, A. Diabetic macular edema: Emerging strategies and treatment algorithms. Dev. Ophthalmol. 2017, 60, 165–174. [Google Scholar] [PubMed]
  119. Dacheva, I.; Ullmer, C.; Ceglowska, K.; Nogoceke, E.; Hartmann, G.; Muller, S.; Rejdak, R.; Nowomiejska, K.; Reich, M.; Nobl, M.; et al. Lysophosphatidic acids and autotaxin in retinal vein occlusion. Retina 2016, 36, 2311–2318. [Google Scholar] [CrossRef] [PubMed]
  120. Udomsinprasert, W.; Honsawek, S.; Anomasiri, W.; Chongsrisawat, V.; Vejchapipat, P.; Poovorawan, Y. Serum autotaxin levels correlate with hepatic dysfunction and severity in postoperative biliary atresia. Biomarkers 2015, 20, 89–94. [Google Scholar] [CrossRef] [PubMed]
  121. Udomsinprasert, W.; Kitkumthorn, N.; Mutirangura, A.; Chongsrisawat, V.; Poovorawan, Y.; Honsawek, S. Association between promoter hypomethylation and overexpression of autotaxin with outcome parameters in biliary atresia. PLoS ONE 2017, 12, e0169306. [Google Scholar] [CrossRef] [PubMed]
  122. Erstad, D.J.; Tager, A.M.; Hoshida, Y.; Fuchs, B.C. The autotaxin-lysophosphatidic acid pathway emerges as a therapeutic target to prevent liver cancer. Mol. Cell. Oncol. 2017, 4, e1311827. [Google Scholar] [CrossRef] [PubMed]
  123. Mazzocca, A.; Dituri, F.; De Santis, F.; Filannino, A.; Lopane, C.; Betz, R.C.; Li, Y.Y.; Mukaida, N.; Winter, P.; Tortorella, C.; et al. Lysophosphatidic acid receptor LPAR6 supports the tumorigenicity of hepatocellular carcinoma. Cancer Res. 2015, 75, 532–543. [Google Scholar] [CrossRef] [PubMed]
  124. Rachakonda, V.P.; Reeves, V.L.; Aljammal, J.; Wills, R.C.; Trybula, J.S.; DeLany, J.P.; Kienesberger, P.C.; Kershaw, E.E. Serum autotaxin is independently associated with hepatic steatosis in women with severe obesity. Obesity 2015, 23, 965–972. [Google Scholar] [CrossRef] [PubMed]
  125. Sun, Y.; Zhang, W.; Evans, J.F.; Floreani, A.; Zou, Z.; Nishio, Y.; Qi, R.; Leung, P.S.; Bowlus, C.L.; Gershwin, M.E. Autotaxin, pruritus and primary biliary cholangitis (PBC). Autoimmun. Rev. 2016, 15, 795–800. [Google Scholar] [CrossRef] [PubMed]
  126. Ikeda, H.; Yatomi, Y. Autotaxin in liver fibrosis. Clin. Chim. Acta 2012, 413, 1817–1821. [Google Scholar] [CrossRef] [PubMed]
  127. Wunsch, E.; Krawczyk, M.; Milkiewicz, M.; Trottier, J.; Barbier, O.; Neurath, M.F.; Lammert, F.; Kremer, A.E.; Milkiewicz, P. Serum autotaxin is a marker of the severity of liver injury and overall survival in patients with cholestatic liver diseases. Sci. Rep. 2016, 6, 30847. [Google Scholar] [CrossRef] [PubMed]
  128. Ringelhan, M.; McKeating, J.A.; Protzer, U. Viral hepatitis and liver cancer. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2017, 372. [Google Scholar] [CrossRef]
  129. Ikeda, H.; Kobayashi, M.; Kumada, H.; Enooku, K.; Koike, K.; Kurano, M.; Sato, M.; Nojiri, T.; Kobayashi, T.; Ohkawa, R.; et al. Annals express: Performance of autotaxin as a serum marker of liver fibrosis. Ann. Clin. Biochem. 2017. [Google Scholar] [CrossRef]
  130. Joshita, S.; Ichikawa, Y.; Umemura, T.; Usami, Y.; Sugiura, A.; Shibata, S.; Yamazaki, T.; Fujimori, N.; Komatsu, M.; Matsumoto, A.; et al. Serum autotaxin is a useful liver fibrosis marker in patients with chronic hepatitis B virus infection. Hepatol. Res. 2018, 48, 275–285. [Google Scholar] [CrossRef] [PubMed]
  131. Yamazaki, T.; Joshita, S.; Umemura, T.; Usami, Y.; Sugiura, A.; Fujimori, N.; Shibata, S.; Ichikawa, Y.; Komatsu, M.; Matsumoto, A.; et al. Association of serum autotaxin levels with liver fibrosis in patients with chronic hepatitis C. Sci. Rep. 2017, 7, 46705. [Google Scholar] [CrossRef] [PubMed]
  132. Farquhar, M.J.; Humphreys, I.S.; Rudge, S.A.; Wilson, G.K.; Bhattacharya, B.; Ciaccia, M.; Hu, K.; Zhang, Q.; Mailly, L.; Reynolds, G.M.; et al. Autotaxin-lysophosphatidic acid receptor signalling regulates hepatitis C virus replication. J. Hepatol. 2017, 66, 919–929. [Google Scholar] [CrossRef] [PubMed]
  133. Kostadinova, L.; Shive, C.L.; Judge, C.; Zebrowski, E.; Compan, A.; Rife, K.; Hirsch, A.; Falck-Ytter, Y.; Schlatzer, D.M.; Li, X.; et al. During Hepatitis C Virus (HCV) infection and HCV-HIV coinfection, an elevated plasma level of autotaxin is associated with lysophosphatidic acid and markers of immune activation that normalize during interferon-free HCV therapy. J. Infect. Dis. 2016, 214, 1438–1448. [Google Scholar] [CrossRef] [PubMed]
  134. Fujiwara, N.; Friedman, S.L.; Goossens, N.; Hoshida, Y. Risk factors and prevention of hepatocellular carcinoma in the era of precision medicine. J. Hepatol. 2018, 68, 526–549. [Google Scholar] [CrossRef] [PubMed]
  135. Hausmann, J.; Kamtekar, S.; Christodoulou, E.; Day, J.E.; Wu, T.; Fulkerson, Z.; Albers, H.M.; van Meeteren, L.A.; Houben, A.J.; van Zeijl, L.; et al. Structural basis of substrate discrimination and integrin binding by autotaxin. Nat. Struct. Mol. Biol. 2011, 18, 198–204. [Google Scholar] [CrossRef] [PubMed]
  136. Liu, J.; Chen, S.; Wang, W.; Ning, B.F.; Chen, F.; Shen, W.; Ding, J.; Chen, W.; Xie, W.F.; Zhang, X. Cancer-associated fibroblasts promote hepatocellular carcinoma metastasis through chemokine-activated hedgehog and TGF-β pathways. Cancer Lett. 2016, 379, 49–59. [Google Scholar] [CrossRef] [PubMed]
  137. Benesch, M.G.K.; Tang, X.; Dewald, J.; Dong, W.F.; Mackey, J.R.; Hemmings, D.G.; McMullen, T.P.; Brindley, D.N. Tumor-induced inflammation in mammary adipose tissue stimulates a vicious cycle of autotaxin expression and breast cancer progression. FASEB J. 2015, 29, 3990–4000. [Google Scholar] [CrossRef] [PubMed]
  138. Wu, T.; Kooi, C.V.; Shah, P.; Charnigo, R.; Huang, C.; Smyth, S.S.; Morris, A.J. Integrin-mediated cell surface recruitment of autotaxin promotes persistent directional cell migration. FASEB J. 2014, 28, 861–870. [Google Scholar] [CrossRef] [PubMed]
  139. Barbayianni, E.; Kaffe, E.; Aidinis, V.; Kokotos, G. Autotaxin, a secreted lysophospholipase D, as a promising therapeutic target in chronic inflammation and cancer. Prog. Lipid Res. 2015, 58, 76–96. [Google Scholar] [CrossRef] [PubMed]
  140. Zheng, H.; Hogberg, J.; Stenius, U. ATM-activated autotaxin (ATX) propagates inflammation and DNA damage in lung epithelial cells; a new mode of action for silica-induced DNA damage? Carcinogenesis 2017, 38, 1196–1206. [Google Scholar] [CrossRef] [PubMed]
  141. Sun, S.; Zhang, X.; Lyu, L.; Li, X.; Yao, S.; Zhang, J. Autotaxin expression is regulated at the post-transcriptional level by the RNA-binding proteins HuR and AUF1. J. Biol. Chem. 2016, 291, 25823–25836. [Google Scholar] [CrossRef] [PubMed]
  142. Yi, X.; Zhou, Y.; Zheng, W.; Chambers, S.K. Hur expression in the nucleus correlates with high histological grade and poor disease-free survival in ovarian cancer. Aust. N. Z. J. Obstet. Gynaecol. 2009, 49, 93–98. [Google Scholar] [CrossRef] [PubMed]
  143. Zhu, Z.; Wang, B.; Bi, J.; Zhang, C.; Guo, Y.; Chu, H.; Liang, X.; Zhong, C.; Wang, J. Cytoplasmic HuR expression correlates with P-gp, HER-2 positivity, and poor outcome in breast cancer. Tumour Biol. 2013, 34, 2299–2308. [Google Scholar] [CrossRef] [PubMed]
  144. Benesch, M.G.K.; Ko, Y.M.; Tang, X.; Dewald, J.; Lopez-Campistrous, A.; Zhao, Y.Y.; Lai, R.; Curtis, J.M.; Brindley, D.N.; McMullen, T.P. Autotaxin is an inflammatory mediator and therapeutic target in thyroid cancer. Endocr. Relat. Cancer 2015, 22, 593–607. [Google Scholar] [CrossRef] [PubMed]
  145. Baldan, F.; Mio, C.; Allegri, L.; Conzatti, K.; Toffoletto, B.; Puppin, C.; Radovic, S.; Vascotto, C.; Russo, D.; Di Loreto, C.; et al. Identification of tumorigenesis-related mRNAs associated with RNA-binding protein HuR in thyroid cancer cells. Oncotarget 2016, 7, 63388–63407. [Google Scholar] [CrossRef] [PubMed]
  146. Liu, S.; Umezu-Goto, M.; Murph, M.; Lu, Y.; Liu, W.; Zhang, F.; Yu, S.; Stephens, L.C.; Cui, X.; Murrow, G.; et al. Expression of autotaxin and lysophosphatidic acid receptors increases mammary tumorigenesis, invasion, and metastases. Cancer Cell 2009, 15, 539–550. [Google Scholar] [CrossRef] [PubMed]
  147. Benesch, M.G.K.; Tang, X.; Maeda, T.; Ohhata, A.; Zhao, Y.Y.; Kok, B.P.C.; Dewald, J.; Hitt, M.; Curtis, J.M.; McMullen, T.P.W.; et al. Inhibition of autotaxin delays breast tumor growth and lung metastasis in mice. FASEB J. 2014, 28, 2655–2666. [Google Scholar] [CrossRef] [PubMed]
  148. Volden, P.A.; Skor, M.N.; Johnson, M.B.; Singh, P.; Patel, F.N.; McClintock, M.K.; Brady, M.J.; Conzen, S.D. Mammary adipose tissue-derived lysophospholipids promote estrogen receptor-negative mammary epithelial cell proliferation. Cancer Prev. Res. 2016, 9, 367–378. [Google Scholar] [CrossRef] [PubMed]
  149. Sundaram, S.; Johnson, A.R.; Makowski, L. Obesity, metabolism and the microenvironment: Links to cancer. J. Carcinog. 2013, 12, 19. [Google Scholar] [CrossRef] [PubMed]
  150. De Pergola, G.; Silvestris, F. Obesity as a major risk factor for cancer. J. Obes. 2013, 2013, 291546. [Google Scholar] [CrossRef] [PubMed]
  151. Memet, I.; Tsalkidou, E.; Tsaroucha, A.K.; Lambropoulou, M.; Chatzaki, E.; Trypsianis, G.; Schizas, D.; Pitiakoudis, M.; Simopoulos, C. Autotaxin expression in hepatocellular carcinoma. J. Investig. Surg. 2017, 1–7. [Google Scholar] [CrossRef] [PubMed]
  152. Enooku, K.; Uranbileg, B.; Ikeda, H.; Kurano, M.; Sato, M.; Kudo, H.; Maki, H.; Koike, K.; Hasegawa, K.; Kokudo, N.; et al. Higher LPA2 and LPA6 mRNA levels in hepatocellular carcinoma are associated with poorer differentiation, microvascular invasion and earlier recurrence with higher serum autotaxin levels. PLoS ONE 2016, 11, e0161825. [Google Scholar] [CrossRef] [PubMed]
  153. Lopane, C.; Agosti, P.; Gigante, I.; Sabbà, C.; Mazzocca, A. Implications of the lysophosphatidic acid signaling axis in liver cancer. Biochim. Biophys. Acta 2017, 1868, 277–282. [Google Scholar] [CrossRef] [PubMed]
  154. Nakai, Y.; Ikeda, H.; Nakamura, K.; Kume, Y.; Fujishiro, M.; Sasahira, N.; Hirano, K.; Isayama, H.; Tada, M.; Kawabe, T.; et al. Specific increase in serum autotaxin activity in patients with pancreatic cancer. Clin. Biochem. 2011, 44, 576–581. [Google Scholar] [CrossRef] [PubMed]
  155. Kadekar, S.; Silins, I.; Korhonen, A.; Dreij, K.; Al-Anati, L.; Hogberg, J.; Stenius, U. Exocrine pancreatic carcinogenesis and autotaxin expression. PLoS ONE 2012, 7, e43209. [Google Scholar] [CrossRef] [PubMed]
  156. Emoto, S.; Kurano, M.; Kano, K.; Matsusaki, K.; Yamashita, H.; Nishikawa, M.; Igarashi, K.; Ikeda, H.; Aoki, J.; Kitayama, J.; et al. Analysis of glycero-lysophospholipids in gastric cancerous ascites. J. Lipid Res. 2017, 58, 763–771. [Google Scholar] [CrossRef] [PubMed]
  157. Muinao, T.; Deka Boruah, H.P.; Pal, M. Diagnostic and prognostic biomarkers in ovarian cancer and the potential roles of cancer stem cells—An updated review. Exp. Cell Res. 2018, 362, 1–10. [Google Scholar] [CrossRef] [PubMed]
  158. Seo, E.J.; Kwon, Y.W.; Jang, I.H.; Kim, D.K.; Lee, S.I.; Choi, E.J.; Kim, K.H.; Suh, D.S.; Lee, J.H.; Choi, K.U.; et al. Autotaxin regulates maintenance of ovarian cancer stem cells through lysophosphatidic acid-mediated autocrine mechanism. Stem Cells 2016, 34, 551–564. [Google Scholar] [CrossRef] [PubMed]
  159. Wasniewski, T.; Woclawek-Potocka, I.; Boruszewska, D.; Kowalczyk-Zieba, I.; Sinderewicz, E.; Grycmacher, K. The significance of the altered expression of lysophosphatidic acid receptors, autotaxin and phospholipase A2 as the potential biomarkers in type 1 endometrial cancer biology. Oncol. Rep. 2015, 34, 2760–2767. [Google Scholar] [CrossRef] [PubMed]
  160. Takahashi, K.; Fukushima, K.; Fukushima, N.; Honoki, K.; Tsujiuchi, T. Enhanced cellular functions through induction of LPA2 by cisplatin in fibrosarcoma HT1080 cells. Mol. Cell. Biochem. 2017, 431, 29–35. [Google Scholar] [CrossRef] [PubMed]
  161. Kanehira, M.; Fujiwara, T.; Nakajima, S.; Okitsu, Y.; Onishi, Y.; Fukuhara, N.; Ichinohasama, R.; Okada, Y.; Harigae, H. A lysophosphatidic acid receptors 1 and 3 axis governs cellular senescence of mesenchymal stromal cells and promotes growth and vascularization of multiple myeloma. Stem Cells 2017, 35, 739–753. [Google Scholar] [CrossRef] [PubMed]
  162. Hashimoto, S.; Mikami, S.; Sugino, H.; Yoshikawa, A.; Hashimoto, A.; Onodera, Y.; Furukawa, S.; Handa, H.; Oikawa, T.; Okada, Y.; et al. Lysophosphatidic acid activates Arf6 to promote the mesenchymal malignancy of renal cancer. Nat. Commun. 2016, 7, 10656. [Google Scholar] [CrossRef] [PubMed]
  163. Braeuer, R.R.; Zigler, M.; Kamiya, T.; Dobroff, A.S.; Huang, L.; Choi, W.; McConkey, D.J.; Shoshan, E.; Mobley, A.K.; Song, R.; et al. Galectin-3 contributes to melanoma growth and metastasis via regulation of NFAT1 and autotaxin. Cancer Res. 2012, 72, 5757–5766. [Google Scholar] [CrossRef] [PubMed]
  164. Shoshan, E.; Braeuer, R.R.; Kamiya, T.; Mobley, A.K.; Huang, L.; Vasquez, M.E.; Velazquez-Torres, G.; Chakravarti, N.; Ivan, C.; Prieto, V.; et al. NFAT1 directly regulates IL8 and MMP3 to promote melanoma tumor growth and metastasis. Cancer Res. 2016, 76, 3145–3155. [Google Scholar] [CrossRef] [PubMed]
  165. Lee, S.C.; Fujiwara, Y.; Liu, J.; Yue, J.; Shimizu, Y.; Norman, D.D.; Wang, Y.; Tsukahara, R.; Szabo, E.; Patil, R.; et al. Autotaxin and LPA1 and LPA5 receptors exert disparate functions in tumor cells versus the host tissue microenvironment in melanoma invasion and metastasis. Mol. Cancer Res. 2015, 13, 174–185. [Google Scholar] [CrossRef] [PubMed]
  166. Lee, S.C.; Fujiwara, Y.; Tigyi, G.J. Uncovering unique roles of LPA receptors in the tumor microenvironment. Recept. Clin. Investig. 2015, 2. [Google Scholar] [CrossRef]
  167. Tang, X.; Benesch, M.G.; Dewald, J.; Zhao, Y.Y.; Patwardhan, N.; Santos, W.L.; Curtis, J.M.; McMullen, T.P.; Brindley, D.N. Lipid phosphate phosphatase-1 expression in cancer cells attenuates tumor growth and metastasis in mice. J. Lipid Res. 2014, 55, 2389–2400. [Google Scholar] [CrossRef] [PubMed]
  168. Tanyi, J.L.; Morris, A.J.; Wolf, J.K.; Fang, X.; Hasegawa, Y.; Lapushin, R.; Auersperg, N.; Sigal, Y.J.; Newman, R.A.; Felix, E.A.; et al. The human lipid phosphate phosphatase-3 decreases the growth, survival, and tumorigenesis of ovarian cancer cells: Validation of the lysophosphatidic acid signaling cascade as a target for therapy in ovarian cancer. Cancer Res. 2003, 63, 1073–1082. [Google Scholar] [PubMed]
  169. Mao, X.Y.; Lee, M.J.; Zhu, J.; Zhu, C.; Law, S.M.; Snijders, A.M. Genome-wide screen identifies a novel prognostic signature for breast cancer survival. Oncotarget 2017, 8, 14003–14016. [Google Scholar] [CrossRef] [PubMed]
  170. Morris, K.E.; Schang, L.M.; Brindley, D.N. Lipid phosphate phosphatase-2 activity regulates S-phase entry of the cell cycle in Rat2 fibroblasts. J. Biol. Chem. 2006, 281, 9297–9306. [Google Scholar] [CrossRef] [PubMed]
  171. Flanagan, J.M.; Funes, J.M.; Henderson, S.; Wild, L.; Carey, N.; Boshoff, C. Genomics screen in transformed stem cells reveals RNASEH2A, PPAP2C, and ADARB1 as putative anticancer drug targets. Mol. Cancer Ther. 2009, 8, 249–260. [Google Scholar] [CrossRef] [PubMed]
  172. Tang, X.; Wang, X.; Zhao, Y.Y.; Curtis, J.M.; Brindley, D.N. Doxycycline attenuates breast cancer related inflammation by decreasing plasma lysophosphatidate concentrations and inhibiting NF-κb activation. Mol. Cancer 2017, 16, 36. [Google Scholar] [CrossRef] [PubMed]
  173. Nikolaou, A.; Kokotou, M.G.; Limnios, D.; Psarra, A.; Kokotos, G. Autotaxin inhibitors: A patent review (2012–2016). Expert Opin. Ther. Pat. 2017, 27, 815–829. [Google Scholar] [CrossRef] [PubMed]
  174. Shah, P.; Cheasty, A.; Foxton, C.; Raynham, T.; Farooq, M.; Gutierrez, I.F.; Lejeune, A.; Pritchard, M.; Turnbull, A.; Pang, L.; et al. Discovery of potent inhibitors of the lysophospholipase autotaxin. Bioorg. Med. Chem. Lett. 2016, 26, 5403–5410. [Google Scholar] [CrossRef] [PubMed]
  175. Llona-Minguez, S.; Ghassemian, A.; Helleday, T. Lysophosphatidic acid receptor (LPAR) modulators: The current pharmacological toolbox. Prog. Lipid Res. 2015, 58, 51–75. [Google Scholar] [CrossRef] [PubMed]
  176. Katsifa, A.; Kaffe, E.; Nikolaidou-Katsaridou, N.; Economides, A.N.; Newbigging, S.; McKerlie, C.; Aidinis, V. The bulk of autotaxin activity is dispensable for adult mouse life. PLoS ONE 2015, 10, e0143083. [Google Scholar] [CrossRef] [PubMed]
  177. Castelino, F.V.; Bain, G.; Pace, V.A.; Black, K.E.; George, L.; Probst, C.K.; Goulet, L.; Lafyatis, R.; Tager, A.M. An autotaxin/lysophosphatidic acid/interleukin-6 amplification loop drives scleroderma fibrosis. Arthritis Rheumatol. 2016, 68, 2964–2974. [Google Scholar] [CrossRef] [PubMed]
  178. Castelino, F.V.; Seiders, J.; Bain, G.; Brooks, S.F.; King, C.D.; Swaney, J.S.; Lorrain, D.S.; Chun, J.; Luster, A.D.; Tager, A.M. Amelioration of dermal fibrosis by genetic deletion or pharmacologic antagonism of lysophosphatidic acid receptor 1 in a mouse model of scleroderma. Arthritis Rheum. 2011, 63, 1405–1415. [Google Scholar] [CrossRef] [PubMed]
  179. Sakkas, L.I. Spotlight on tocilizumab and its potential in the treatment of systemic sclerosis. Drug Des. Dev. Ther. 2016, 10, 2723–2728. [Google Scholar] [CrossRef] [PubMed]
  180. Bain, G.; Shannon, K.E.; Huang, F.; Darlington, J.; Goulet, L.; Prodanovich, P.; Ma, G.L.; Santini, A.M.; Stein, A.J.; Lonergan, D.; et al. Selective inhibition of autotaxin is efficacious in mouse models of liver fibrosis. J. Pharmacol. Exp. Ther. 2017, 360, 1–13. [Google Scholar] [CrossRef] [PubMed]
  181. Nakagawa, S.; Wei, L.; Song, W.M.; Higashi, T.; Ghoshal, S.; Kim, R.S.; Bian, C.B.; Yamada, S.; Sun, X.; Venkatesh, A.; et al. Molecular liver cancer prevention in cirrhosis by organ transcriptome analysis and lysophosphatidic acid pathway inhibition. Cancer Cell 2016, 30, 879–890. [Google Scholar] [CrossRef] [PubMed]
  182. Mouratis, M.A.; Magkrioti, C.; Oikonomou, N.; Katsifa, A.; Prestwich, G.D.; Kaffe, E.; Aidinis, V. Autotaxin and endotoxin-induced acute lung injury. PLoS ONE 2015, 10, e0133619. [Google Scholar] [CrossRef] [PubMed]
  183. Thirunavukkarasu, K.; Swearingen, C.A.; Oskins, J.L.; Lin, C.; Bui, H.H.; Jones, S.B.; Pfeifer, L.A.; Norman, B.H.; Mitchell, P.G.; Chambers, M.G. Identification and pharmacological characterization of a novel inhibitor of autotaxin in rodent models of joint pain. Osteoarthr. Cartil. 2017, 25, 935–942. [Google Scholar] [CrossRef] [PubMed]
  184. Kato, K.; Ikeda, H.; Miyakawa, S.; Futakawa, S.; Nonaka, Y.; Fujiwara, M.; Okudaira, S.; Kano, K.; Aoki, J.; Morita, J.; et al. Structural basis for specific inhibition of autotaxin by a DNA aptamer. Nat. Struct. Mol. Biol. 2016, 23, 395–401. [Google Scholar] [CrossRef] [PubMed]
  185. Keune, W.J.; Hausmann, J.; Bolier, R.; Tolenaars, D.; Kremer, A.; Heidebrecht, T.; Joosten, R.P.; Sunkara, M.; Morris, A.J.; Matas-Rico, E.; et al. Steroid binding to autotaxin links bile salts and lysophosphatidic acid signalling. Nat. Commun. 2016, 7, 11248. [Google Scholar] [CrossRef] [PubMed]
  186. Venkatraman, G.; Benesch, M.G.; Tang, X.; Dewald, J.; McMullen, T.P.; Brindley, D.N. Lysophosphatidate signaling stabilizes Nrf2 and increases the expression of genes involved in drug resistance and oxidative stress responses: Implications for cancer treatment. FASEB J. 2015, 29, 772–785. [Google Scholar] [CrossRef] [PubMed]
  187. Bekele, R.T.; Venkatraman, G.; Liu, R.Z.; Tang, X.; Mi, S.; Benesch, M.G.; Mackey, J.R.; Godbout, R.; Curtis, J.M.; McMullen, T.P.; et al. Oxidative stress contributes to the tamoxifen-induced killing of breast cancer cells: Implications for tamoxifen therapy and resistance. Sci. Rep. 2016, 6, 21164. [Google Scholar] [CrossRef] [PubMed]
  188. Balogh, A.; Shimizu, Y.; Lee, S.C.; Norman, D.D.; Gangwar, R.; Bavaria, M.; Moon, C.; Shukla, P.; Rao, R.; Ray, R.; et al. The autotaxin-LPA2 GPCR axis is modulated by gamma-irradiation and facilitates DNA damage repair. Cell. Signal. 2015, 27, 1751–1762. [Google Scholar] [CrossRef] [PubMed]
  189. Jayakumar, S.; Pal, D.; Sandur, S.K. Nrf2 facilitates repair of radiation induced DNA damage through homologous recombination repair pathway in a ros independent manner in cancer cells. Mutat. Res. 2015, 779, 33–45. [Google Scholar] [CrossRef] [PubMed]
  190. Sekhar, K.R.; Freeman, M.L. Nrf2 promotes survival following exposure to ionizing radiation. Free Radic. Biol. Med. 2015, 88, 268–274. [Google Scholar] [CrossRef] [PubMed]
  191. Zhou, S.; Ye, W.; Shao, Q.; Zhang, M.; Liang, J. Nrf2 is a potential therapeutic target in radioresistance in human cancer. Crit. Rev. Oncol. Hematol. 2013, 88, 706–715. [Google Scholar] [CrossRef] [PubMed]
  192. Meng, G.; Tang, X.; Yang, Y.; Benesch, M.G.K.; Marshall, A.; Murray, D.; Hemmings, D.G.; Wuest, F.; McMullen, T.P.W.; Brindley, D.N. Gamma-radiation causes an inflammatory wound healing response in adipose tissue involving autotaxin and lysophosphatidate signaling. FASEB J. 2017. [Google Scholar] [CrossRef]
  193. Schleicher, S.M.; Thotala, D.K.; Linkous, A.G.; Hu, R.; Leahy, K.M.; Yazlovitskaya, E.M.; Hallahan, D.E. Autotaxin and LPA receptors represent potential molecular targets for the radiosensitization of murine glioma through effects on tumor vasculature. PLoS ONE 2011, 6, e22182. [Google Scholar] [CrossRef] [PubMed]
  194. Bhave, S.R.; Dadey, D.Y.; Karvas, R.M.; Ferraro, D.J.; Kotipatruni, R.P.; Jaboin, J.J.; Hallahan, A.N.; Dewees, T.A.; Linkous, A.G.; Hallahan, D.E.; et al. Autotaxin inhibition with PF-8380 enhances the radiosensitivity of human and murine glioblastoma cell lines. Front. Oncol. 2013, 3, 236. [Google Scholar] [CrossRef] [PubMed]
  195. Harris, J.R. Fifty years of progress in radiation therapy for breast cancer. Am. Soc. Clin. Oncol. Educ. Book 2014, 21–25. [Google Scholar] [CrossRef] [PubMed]
  196. Gan, L.; Xue, J.X.; Li, X.; Liu, D.S.; Ge, Y.; Ni, P.Y.; Deng, L.; Lu, Y.; Jiang, W. Blockade of lysophosphatidic acid receptors LPAR1/3 ameliorates lung fibrosis induced by irradiation. Biochem. Biophys. Res. Commun. 2011, 409, 7–13. [Google Scholar] [CrossRef] [PubMed]
  197. Xue, J.; Gan, L.; Li, X.; Li, J.; Qi, G.; Wu, Y.; Fu, X.; Mao, Q.; Ao, R.; Lang, J.; et al. Effects of lysophosphatidic acid and its receptors LPA(1/3) on radiation pneumonitis. Oncol. Rep. 2010, 24, 1515–1520. [Google Scholar] [PubMed]
  198. Desroy, N.; Housseman, C.; Bock, X.; Joncour, A.; Bienvenu, N.; Cherel, L.; Labeguere, V.; Rondet, E.; Peixoto, C.; Grassot, J.M.; et al. Discovery of 2-[[2-ethyl-6-[4-[2-(3-hydroxyazetidin-1-yl)-2-oxoethyl]piperazin-1-yl]-8-methyli midazo[1,2-a]pyridin-3-yl]methylamino]-4-(4-fluorophenyl)thiazole-5-carbonitrile (GLPG1690), a first-in-class autotaxin inhibitor undergoing clinical evaluation for the treatment of idiopathic pulmonary fibrosis. J. Med. Chem. 2017, 60, 3580–3590. [Google Scholar] [PubMed]
  199. Kihara, Y.; Mizuno, H.; Chun, J. Lysophospholipid receptors in drug discovery. Exp. Cell Res. 2015, 333, 171–177. [Google Scholar] [CrossRef] [PubMed]
  200. Crack, P.J.; Zhang, M.; Morganti-Kossmann, M.C.; Morris, A.J.; Wojciak, J.M.; Fleming, J.K.; Karve, I.; Wright, D.; Sashindranath, M.; Goldshmit, Y.; et al. Anti-lysophosphatidic acid antibodies improve traumatic brain injury outcomes. J. Neuroinflamm. 2014, 11, 37. [Google Scholar] [CrossRef] [PubMed]
  201. Brindley, D.N.; Lin, F.T.; Tigyi, G.J. Role of the autotaxin-lysophosphatidate axis in cancer resistance to chemotherapy and radiotherapy. Biochim. Biophys. Acta 2013, 1831, 74–85. [Google Scholar] [CrossRef] [PubMed]
  202. Rancoule, C.; Espenel, S.; Trone, J.-C.; Langrand-Escure, J.; Vallard, A.; Rehailia-Blanchard, A.; Hamrouni, A.E.M.; Xia, Y.; Guy, J.-B.; Ben-Mrad, M.; et al. Lysophosphatidic acid (LPA) as a pro-fibrotic and pro-oncogenic factor: A pivotal target to improve the radiotherapy therapeutic index. Oncotarget 2017, 8, 43543–43554. [Google Scholar] [CrossRef] [PubMed]
  203. Citrin, D.E.; Mitchell, J.B. Mechanisms of normal tissue injury from irradiation. Semin. Radiat. Oncol. 2017, 27, 316–324. [Google Scholar] [CrossRef] [PubMed]
  204. Gangadhar, T.C.; Vonderheide, R.H. Mitigating the toxic effects of anticancer immunotherapy. Nat. Rev. Clin. Oncol. 2014, 11, 91–99. [Google Scholar] [CrossRef] [PubMed]
  205. Yang, L.; Yu, H.; Dong, S.; Zhong, Y.; Hu, S. Recognizing and managing on toxicities in cancer immunotherapy. Tumour Biol. 2017, 39. [Google Scholar] [CrossRef] [PubMed]
Figure 1. General overview of cancer initiation and the role of autotaxin/lysophosphatidate (ATX/LPA). A convergence of extrinsic and intrinsic pathways leads to sustained inflammatory and survival signaling that involves upregulation of ATX/LPA signaling through both increases in ATX and LPA concentrations with concurrent decreases in eco-lipid phosphate phosphatase (LPP) activity. The establishment of this vicious cycle leads to cancer initiation and progression as often described by the hallmarks of cancer (sustained proliferative signaling, evasion of growth suppressors, replicative immortality, angiogenesis, resistance to cell death, deregulation of cellular genetics, avoidance of the immune system, and invasion and metastasis) [3,6,7].
Figure 1. General overview of cancer initiation and the role of autotaxin/lysophosphatidate (ATX/LPA). A convergence of extrinsic and intrinsic pathways leads to sustained inflammatory and survival signaling that involves upregulation of ATX/LPA signaling through both increases in ATX and LPA concentrations with concurrent decreases in eco-lipid phosphate phosphatase (LPP) activity. The establishment of this vicious cycle leads to cancer initiation and progression as often described by the hallmarks of cancer (sustained proliferative signaling, evasion of growth suppressors, replicative immortality, angiogenesis, resistance to cell death, deregulation of cellular genetics, avoidance of the immune system, and invasion and metastasis) [3,6,7].
Cancers 10 00073 g001
Figure 2. Overview of the LPA signaling axis. Extracellular LPA is produced from LPC by the lysophospholipase D activity of ATX. LPA then signals through at least six known G-protein coupled LPA receptors to mediate its host of physiological and pathological effects. LPA is rapidly turned over by the eco-activity of LPP1-3 into inorganic phosphate and monoacylglycerol (MAG), which apart from 2-arachidonoylglycerol, does not affect signaling.
Figure 2. Overview of the LPA signaling axis. Extracellular LPA is produced from LPC by the lysophospholipase D activity of ATX. LPA then signals through at least six known G-protein coupled LPA receptors to mediate its host of physiological and pathological effects. LPA is rapidly turned over by the eco-activity of LPP1-3 into inorganic phosphate and monoacylglycerol (MAG), which apart from 2-arachidonoylglycerol, does not affect signaling.
Cancers 10 00073 g002
Figure 3. Overview of γ-radiation-induced inflammation in adipose tissue. γ-radiation, upon inducing double strand DNA breaks, activates the proteins ATM, ATR, and PARP-1. These in turn activate NFκB, facilitating the expression of COX-2, ATX, LPA1–2 and numerous inflammatory mediators, which in concert lead to repair and survival of tissues.
Figure 3. Overview of γ-radiation-induced inflammation in adipose tissue. γ-radiation, upon inducing double strand DNA breaks, activates the proteins ATM, ATR, and PARP-1. These in turn activate NFκB, facilitating the expression of COX-2, ATX, LPA1–2 and numerous inflammatory mediators, which in concert lead to repair and survival of tissues.
Cancers 10 00073 g003
Table 1. Summary of novel targets of the ATX-LPA signaling axis currently in clinical trials.
Table 1. Summary of novel targets of the ATX-LPA signaling axis currently in clinical trials.
CompoundMechanism of ActionClinical StageClinical IndicationCompanyClinicaltrials.gov ID
GLPG1690ATX direct inhibitionPhase 2IPFGalapagos NV (Mechelen, Belgium)NCT02738801
BMS-986020LPA1 antagonistPhase 2IPFBristol-Myers Squibb (New York, NY, USA)NCT01766817
SAR100842LPA1/3 antagonistPhase 2Systemic SclerosisSanofi (Paris, France)NCT01651143
Lpathomab TMLPA monoclonal antibodyPhase 1___Lpath, Inc. (San Diego, CA, USA)NCT02341508
IPF, idiopathic pulmonary fibrosis.

Share and Cite

MDPI and ACS Style

Benesch, M.G.K.; MacIntyre, I.T.K.; McMullen, T.P.W.; Brindley, D.N. Coming of Age for Autotaxin and Lysophosphatidate Signaling: Clinical Applications for Preventing, Detecting and Targeting Tumor-Promoting Inflammation. Cancers 2018, 10, 73. https://doi.org/10.3390/cancers10030073

AMA Style

Benesch MGK, MacIntyre ITK, McMullen TPW, Brindley DN. Coming of Age for Autotaxin and Lysophosphatidate Signaling: Clinical Applications for Preventing, Detecting and Targeting Tumor-Promoting Inflammation. Cancers. 2018; 10(3):73. https://doi.org/10.3390/cancers10030073

Chicago/Turabian Style

Benesch, Matthew G.K., Iain T.K. MacIntyre, Todd P.W. McMullen, and David N. Brindley. 2018. "Coming of Age for Autotaxin and Lysophosphatidate Signaling: Clinical Applications for Preventing, Detecting and Targeting Tumor-Promoting Inflammation" Cancers 10, no. 3: 73. https://doi.org/10.3390/cancers10030073

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