Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 Mar 1.
Published in final edited form as: Curr Pathobiol Rep. 2016 Feb 19;4(1):27–35. doi: 10.1007/s40139-016-0097-3

The role of IL-17 signaling in regulation of the liver-brain axis and intestinal permeability in Alcoholic Liver Disease

Hsiao-Yen Ma 1,2, Jun Xu 1,2, Xiao Liu 1,2, Yunheng Zhu 1,2, Bin Gao 8, Michael Karin 4, Hidekazu Tsukamoto 9, Dilip V Jeste 5,7, Igor Grant 5, Amanda J Roberts 10, Candice Contet 11, Cedric Geoffroy 6, Binhai Zheng 6, David Brenner 1, Tatiana Kisseleva 2
PMCID: PMC4878828  NIHMSID: NIHMS761714  PMID: 27239399

Abstract

Alcoholic liver disease (ALD) progresses from a normal liver, to steatosis, steatohepatitis, fibrosis and hepatocellular carcinoma (HCC). Despite intensive studies, the pathogenesis of ALD is poorly understood, in part due to a lack of suitable animal models which mimic the stages of ALD progression. Furthermore, the role of IL-17 in ALD has not been evaluated. We and others have recently demonstrated that IL-17 signaling plays a critical role in development of liver fibrosis and cancer. Here we summarize the most recent evidence supporting the role of IL-17 in ALD. As a result of a collaborative effort of Drs. Karin, Gao, Tsukamoto and Kisseleva, we developed several improved models of ALD in mice: 1) chronic-plus-binge model that mimics early stages of steatohepatitis, 2) intragastric ethanol feeding model that mimics alcoholic steatohepatitis and fibrosis, and 3) diethylnitrosamine (DEN)+alcohol model that mimics alcoholic liver cancer. These models might provide new insights into the mechanism of IL-17 signaling in ALD and help identify novel therapeutic targets.

Keywords: alcoholic liver disease, ethanol metabolism, activated myofibroblasts, regression of liver fibrosis, hepatocellular carcinoma, Innate Immunity, Adaptive Immunity, Cytokines, Inflammation

I. Introduction

Alcoholic liver disease (ALD) is a major cause of cirrhosis and liver failure, which is the 12th leading cause of death in patients in the United States1. ALD progresses from steatosis, to steatohepatitis, fibrosis, cirrhosis, and finally hepatocellular carcinoma (HCC)1. Several injury-triggered events (see below) play a critical role in the pathogenesis of ALD. To-date there is no effective treatment of ALD, in part because there are no pre-clinical models available to study ALD progression. Furthermore, the majority of preclinical models focuses on the effect of chronic alcohol consumption on pathology of a single organ, such as liver, brain, heart or kidneys. In reality, alcohol-induced injury produces a systemic effect, and the failure of the damaged liver to perform detoxifying function also has a profound effect on the brain, and other organs. Here we summarize the recent evidence for the role of IL-17 signaling pathway in alcohol-induced injury of the liver and the brain, and regulation of the intestinal permeability, the critical factors that drive development of alcoholic liver disease, .

II. ALD progression in patients

Progression of ALD from steatohepatitis to fibrosis

ALD studies have been hampered by the absence of suitable animal models. In patients ALD progresses from fatty liver to steatohepatitis and fibrosis, and often leads to development of HCC. Each stage is characterized by specific morphological changes and upregulation of specific sets of cytokines. Recently, we developed a chronic-plus-binge ethanol feeding model, which induces significant liver inflammation and neutrophil infiltration but not fibrosis 2,3, and reflects early stages of steatohepatitis. Alcohol-induced damage to hepatocytes is induced via up-regulation of cytochrome P4502E1, SREBP-1c causing accumulation of fat droplets (mainly triglycerides and phospholipids), centrilobular ballooning of hepatocytes, and formation of Mallory–Denk hyaline inclusions4. Serum levels of about 250IU/L ALT and 420 IU/L AST were found post single binge gavage, and correlated with increased expression of inflammatory cytokines IL-8, IL-6, IL-1β and development of hepatic oxidative stress 1,4. Neutrophilic infiltration is the major feature of alcoholic steatohepatitis. Apoptotic hepatocytes release TGF-β1 and factors, including IL-8, CXCL1 (Gro-α), and IL-17, that facilitate recruitment of inflammatory cells to the fatty liver. Infiltrating BM-derived neutrophils kill sensitized hepatocytes, and further exacerbate alcohol-induced liver injury5. A rodent model of ASH has demonstrated a pivotal role of neutrophils in pathogenesis of ALD 1,2. Recruited T and B lymphocytes also contribute to liver damage causing activation liver resident Kuppfer cells, which secrete TGF-β1 and activate hepatic myofibroblasts. Myofibroblasts are the primary source of extracellular matrix (ECM) in fibrotic liver611. Activated Hepatic Stellate Cells (aHSCs) have been recently demonstrated to serve as a major source of myofibroblasts in alcohol-damaged liver Under physiological conditions HSCs store Vitamin A and function as liver pericytes, but in response to sustained exposure to alcohol, HSCs rapidly differentiate into fibrogenic myofibroblasts, start producing Collagen Type I, the major component of extracellular matrix, and make liver fibrotic. Up-to-date the intragastric model of ethanol feeding (Tsukamoto-French model)12 is the best rodent model of alcohol-induced liver fibrosis mimics this stage of alcoholic fibrosis in patients, and these mice develop significant level of liver fibrosis after 2 months of alcohol12,13. This stage is characterized by release of TGF-β1, mostly by Kupffer cells4, and activation of Hepatic stellate cells (HSCs)13,14. Furthermore, a recent study has demonstrated that addition of ethanol to drinking water increased tumor incidence in DEN-injected male mice15, suggesting that this model can be used to study the effects of ethanol on HCC progression.

Hepatocellular carcinoma (HCC)

HCC is the firth most common cancer worldwide and the third most common cause of cancer death16. HCC is a malignant tumor made of cells resembling hepatocytes with a plate-like organization around sinusoids17, usually arises in a cirrhotic liver of patients with ALD16,18, and is identified by expression of alpha-fetoprotein (AFP), CD90, CD133, YAP and EpCAM expression19. Several mechanisms contribute to development of HCC in patients with alcoholic cirrhosis, including sustained inflammation, immunosuppressive effect of alcohol, impaired hepatocyte proliferation, loss of cell cycle checkpoints and increased tumor cell survival, telomere shortening and chromosomal instability1,2. Three potential cellular sources of HCC have been suggested: 1) mature hepatocytes as unipotential stem cells which rapidly regenerate to restore the liver mass in response to acute injury, 2) oval cells as bipotential stem cells which are activated and proliferate in response to chronic injury when proliferation of hepatocytes is exhausted or inhibited, 3) BM-derived stem cells20,21. Accumulating evidence suggests that HCC originates from dedifferentiation and transformation of mature hepatocytes, or maturation arrest of oval cells18,19. Progression of HCC in patients with ALD is associated with upregulation of IL-622, IL-1723,24 and IL-2225 and constitutive activation of Stat326. Consistent with this, IL-22-/- mice are less susceptible to DEN-induced HCC than wild type mice. In addition to Stat327, NFκB, Wnt/β-catenin, and Hedgehog signaling pathways were implicated in HCC development23,2831.

After injury and loss of hepatic mass, the liver regenerates mainly via proliferation of remaining adult hepatocytes. Oval cells (ductular reaction) activate when proliferation of hepatocytes is inhibited or exhasted1. Oval cells are bipotential liver progenitor cells, which reside in the Canal of Herring32, and give rise to hepatocytes and cholangiocytes20,33. Oval cells exhibit a CD45/11b/31/MIC1-1C3+/133+/26 phenotype34. Several studies indicate that these may originate from Sox9-expressing clonogenic progenitors3336. Therefore, recently generated Sox9creERT2-R26RYFP mice34 may be useful for lineage tracing of oval cells. The oval cell reaction includes a broader progenitor population which can be identified by expression of A6, AFP, FoxJ134,36 and other markers37 in mice. Chronic alcohol consumption inhibits hepatocyte proliferation, increasing the number of oval cells in patients with ALD. Proliferation of oval cells correlates with the severity of ALD and risk of alcoholic HCC. It has been suggested that tumor progenitors may originate from the oval cell reaction emerging in response chronic alcohol exposure38. Recent studies have implicated IL-22 in the regulation of alcohol-induced oval cell response and HCC progression. Thus, overexpression of IL-22 in the liver (IL-22TG mice) drives exaggerated oval cell proliferation via Stat3 activation26,39, suggesting that IL-22/Stat3 signaling may be critical in HCC40.

Systemic effect of alcohol on liver-brain axis and intestinal permeability

Liver metabolizes alcohol, and therefore, is directly affected by chronic alcohol consumption. In turn, liver dysfunction contributes to systemic release of proinflammatory microbial products, toxic lipids (such as ceramides) and cytokines into the circulation, and exacerbates cytotoxic effect of alcohol on other organs, including development of insulin resistance and oxidative stress. The central nervous system (CNS) is the other major target of alcohol toxicity and degeneration. In addition to its direct neurotoxic effects, alcohol misuse establishes a liver-brain axis of neurodegeneration mediated by toxic lipid trafficking across the blood-brain barrier, leading to a range of complications that begin with mild neurocognitive impairment but can progress to more severe dementing disorder. The neuroanatomic underpinnings of these neurocognitive disorders include disruption of white matter integrity as evidenced by reduction in fractional anisotropy and increase in diffusivity measures on diffusion tensor imaging; and loss of volume in hippocampus, frontal cortex, subcortical structures and cerebellum. On structural brain imaging brain volume loss may be manifested by cortical thinning, white matter loss, and corresponding enlargement of sulci and ventricles. These changes may be accompanied by neuropathologic findings of astrogliosis, loss of synaptodendritic complexity, loss of cytoskeleton, and ultimately neuronal loss. When complicated by thiamine deficiency there may be additional damage to thalamus, and mammillary bodies with clinical presentation of Wernicke Korsakoff syndrome [amnestic-confabulatory syndrome]4145.

III. Evidence of the role of IL-17 signaling pathway in ALD

Interleukin 17 (IL-17)

Interleukin-17 (IL-17)-producing effector CD4+ T (Th17) cells46,47 originate from naïve T cells via activation of lineage specific transcription factors48,49, regulated by TGF-β1 and IL-6, and other cytokines50,51. IL-17 is mainly produced by CD4+ Th17 cells, but also by a variety of cells, including γδ T cells, CD8+ T cells, NKT cells, NK cells, innate lymphoid cells, eosinophils, neutrophils, and monocytes52. Th17 cells secrete IL-17 cytokines, a family of cytokines comprised of IL-17A, IL-17F, IL-17B, IL-17C and IL-17E53. IL-17A homodimers (also known as IL-17) is the most abundant in Th17 cells, exhibit higher biological activity, and signal through their cognate receptors IL-17RA and IL-17RC52. IL-17RA is ubiquitously expressed, but is strongly induced in hematopoietic cells54 and fibroblasts55 in response to stress. IL-17A signaling activates inflammatory signaling in target cells, including Stat3, TRAF6, Act1, JNK, ERK, NF-κB54,56. IL-17 mediates autoimmunity, and the autoimmune inflammatory diseases psoriasis and rheumatoid arthritis respond to anti-IL-17 biological therapies57. Most recently, IL-17 has been implicated in liver, lung, and skin fibrosis, and in tumorigenesis52,53,5860,5,6164. Although anti-TNF-α therapy has been ineffective in patients with ALD1,65, the corollary of our underlying hypothesis is that anti-IL-17 therapy is a potential novel therapy for ALD. The autoimmune inflammatory diseases psoriasis and rheumatoid arthritis respond to anti-IL-17 biological therapies57. Most recently, IL-17 has been implicated in liver, lung, and skin fibrosis, and in tumorigenesis 52,66. We have demonstrated that IL-17 is a critical mediator of liver fibrosis of different etiologies2,67.

IL-17 in liver fibrosis

Patients with ALD have a high serum level of IL-17. Accumulation of Th17 cells was significantly increased in the livers of patients with ALD, and the numbers of Th17 cells correlated with fibrosis score5. Several events play a critical role in progression of alcohol-related liver fibrosis. Hepatocyte apoptosis causes recruitment of inflammatory cells to the damaged liver and release of pro-fibrogenic cytokines (TGF-β1, IL-6, IL-1β, TNF-α). Our group has recently demonstrated that IL-17A and its receptor IL-17RA are highly upregulated in injured livers, and IL-17 signaling plays a critical role in the pathogenesis of liver fibrosis. IL-17 regulates production of TGF-β1 by activated Kupffer cells, and can directly activate Collagen Type I production by HSCs, the major source of fibrogenic myofibroblasts in fibrotic liver. Deletion of IL-17 signaling in mice resulted in inhibition of liver fibrosis by 75%. Abrogation of IL-17 signaling in hematopoietic cells (as demonstrated by deletion of either IL-17A or IL-17RA in BM) decreases liver fibrosis by 50%. Kupffer cells are the primary targets of IL-17, IL-17 regulates approximately 30% of TGF-β1 production by Kupffer cells. Meanwhile, deletion of IL-17RA in non-immune liver resident cells decreases liver fibrosis by 25%67. In this case, HSCs are the primary non-parenchymal targets of IL-17 in fibrotic liver, and IL-17A can directly stimulate activation of HSCs or induce IL-6 production, which stimulates Collagen Type I production in HSCs67. Increased expression of IL-17A was detected in livers from patients with liver fibrosis and cirrhosis of different etiologies (compared to patients with no fibrosis), and correlated with the severity of the disease23.

Regulation of Th17 differentiation in liver fibrosis

TGF-β1, IL-6 are strongly upregulated during development of ALD-induced fibrosis. In the mean time, TGF-β1, IL-6 and IL-21 drive differentiation of Th17 cells from naïve Th0 cells50 via activation of retinoid-related orphan receptor γt (ROR γt)48. IL-23 is required for Th17 proliferation52. IL-23 is expressed by the macrophages and DCs, signals through IL-12Rbeta1 and IL-23R receptors, and activate Jak2/STAT3 signaling pathway68. Mice deficient of IL-23p19, have very few Th17 cells69,70, suggesting that and the main biological functions of IL-23 is regulation of Th17 cell expansion. IL-23 is upregulated along with IL-17 in fibrotic liver, and IL-23−/− deficient mice develop less fibrosis in response to cholestatic and toxic liver injury67, indicating that the IL-23/Th17 axis may become a promising target for suppressing liver inflammation during ALD71,72. Furthermore, IL-23 is upregulated in multiple human cancers, and ablation of IL-23p19 gene resulted in reduced tumorigenesis in a mouse model of skin cancer73, colitis-associated cancer (CAC)66. There is emerging evidence that IL-23 also promotes HCC development7476. IL-27 antagonizes expansion of Th17 via inhibition of IL-23-producing cells. is formed from IL-27p28 and Ebi3 subunits 77 and IL-27p28−/−78 and Ebi3−/ −79 knockout mice have been generated. IL-27 signals via IL-27RA and common receptor chain gp130, activating STAT3 and Stat1 in target cells77,80. IL-17RA−/ − mice81 exhibit a dramatic increase in Th17 activity, demonstrating that IL-27 suppresses de novo Th17 cell differentiation driven by IL-6 and TGF-β178.

IL-25 also blocks Th17 cell proliferation via inhibition of IL-23, IL-1β1 and IL-6 secretion by dendritic (and other) cells 52. IL-25 propagates allergic responses8284. IL-25 binds to IL-17RA and IL-17RB heterodimers (of which IL-17RB represents an IL-25 specific moiety85,86), and induces Act1-dependent activation of NFκB signaling pathway in target cells87. IL-25 drives the expression of IL-1388, which is required for suppression of Th17 responses8284,89,90. We have demonstrated that IL-25 attenuates liver fibrosis in mice, suggesting that IL-25 agonists may become targets for ALD treatment67.

IL-17 in brain and spinal cord

In addition to immune cells, glial cells in the CNS also express IL-17 under physiological conditions91. IL-17R is widely expressed within the CNS and upregulated under inflammatory conditions92. Genetic deletion of IL-17 increased the number of adult-born neurons. Furthermore, IL-17 deletion altered the network of the cytokines, facilitated basal excitatory synaptic transmission, enhanced intrinsic neuronal excitability, and increased expression of proneuronal genes in neuronal progenitor cells (NPCs), suggesting a profound role of IL-17 in the negative regulation of adult hippocampal neurogenesis under physiology conditions93. In an ischemic brain injury model, IL-17, highly expressed by γδ T lymphocytes, has been shown to play an important role in mediating the evolution of brain infarction and accompanying neurological deficits in the delayed phase of injury94. In a spinal cord injury model, IL-17 deletion improved tissue sparing and locomotor recovery without significantly affecting microglial activation and astroglial reactivity95.

IL-17 in blood-brain barrier

In addition, Th17 lymphocytes promote blood-brain barrier disruption and central nervous system inflammation92,96. Aging augments T cell-derived release of IL-17 and granzyme B that mediate neuronal cell death. IL-17 and IL-22 receptors are expressed on blood-brain barrier endothelial cells (BBB-ECs), and elevated levels of IL-17 and IL-22 disrupt BBB tight junctions in vitro and in vivo. Furthermore, Th17 lymphocytes transmigrate efficiently across BBB-ECs, highly express granzyme B, and kill neurons and promote CNS inflammation through recruitment of CD4+ lymphocytes.

IL-17 and intestinal permeability

The translocation of bacteria and bacterial products into the circulation, and subsequent changes in the microbiome composition are associated with chronic alcohol consumption. Thus, overgrowth of Bacteroidetes and Verrucomicrobia bacteria was observed in alcohol-fed mice (compared with a predominance of Firmicutes bacteria in control mice), and was associated with downregulation in gene and protein expression of bactericidal c-type lectins Reg3b and Reg3g in the small intestine97. Commensal bacteria regulate efficiency of immune response, and vice versa. For example, mono-colonization of mice with segmented filamentous bacteria (SFB) results in induction of proinflammatory factors that favor expansion and accumulation of Th17 cells in the small intestine, and elicits a systemic Th17 response. Intestinal microbiota have also been shown to play a critical role in the absorption of lipopolysaccharide (LPS)98.

IL-17 and aging

Aging is associated with change of liver function caused by increased steatosis, inflammation, and fibrosis99. Changes in hypothalamic-pituitary-adrenal (HPA) activity are one of several proposed mechanisms involved in brain aging100. Recent studies have also implicated IL-17 in the process of aging in humans and mice101. For example, it has been suggested that aging is associated with changes in the immune system that affect specific T cell functions. The immune response to infection, immunization, and tumors in aged individuals is quite different from that found in the young. Specifically, aged naive CD4 T cells do not differentiate well to Th1 and Th2 effector subsets, but exhibit an increased ability to generate functional Th17 effectors, which can be found readily in older individuals. Therefore, the levels of IL-17 are highly increased in older individuals. Th17 effectors produce high levels of IL-17 family cytokines (IL-17, IL-21, and IL-22). In addition to the greater prevalence of Th17 effectors, aging is associated with expansion of regulatory T cells (Treg)101. Since IL-2 was shown to inhibit the expression of IL-17, and blocking IL-2 promotes the differentiation of Th17 effectors102. Therefore, it has been suggested that the presence of regulatory T cells during an immune response may favor the development of a Th17 polarized response because the regulatory cells consume IL-2, which is needed for the development of Th1 and Th2 but not Th17 effectors. These observations also suggest that aging has very specific effects on CD4 T cell populations and does not lead simply to an overall downregulation of T cell function101.

IL-17 and HIV

Th17 cells play a crucial role in protection against infections. Therefore, it is not surprising that IL-17-producing T cells play an important role in pathogenesis of HIV and opportunistic infections observed in AIDS patients103. Specifically, the loss of balance between Th17 cells and Tregs was linked to increased immune activation and HIV progression. Although the numbers of Th17 cells in the peripheral blood often vary in AIDS patients, Th17 cells are substantially depleted from the gastrointestinal tract, leading to a loss of mucosal integrity, increased microbial translocation, and further impairment of systemic immune responses103. Furthermore, excessive alcohol use is common among AIDS patients, and greatly augments HIV-associated neurocognitive deficits104. However, the role of IL-17 signaling in HIV progression complicated by chronic alcohol abuse has not been evaluated. A longitudinal assessment of functional changes in circulating and tissue Th17 cells is urgently needed in order to better determine the dynamic of Th17 cells in peripheral blood, and IL-17-specific regulation of liver-brain axis and intestinal permeability in AIDS patients with a history of chronic alcohol abuse.

Regulation of Th17 differentiation by gut microbiota

The composition of microbiota has been linked to the differentiation of Th17 cells in the gut, specifically in the small intestine lamina propria. In vitro, IL-17 expressing T-helper cells are induced by the interactions of cytokines TGF-β, IL-6, IL-21 & IL-23; these cytokines also play an important role in Th17 differentiation in vivo and regulation of inflammatory immune response. Recent studies have demonstrated the correlation between Th17 in vivo differentiation and induction in the small intestine lamina propria with the presence of intestinal Cytophaga-Flavobacter-Bacteriodes bacteria105. Here, Th17 differentiation was observed independent of IL-21 & IL-23 signaling, the cytokines typically associated with regulation of Th17 expansion. Additionally, the abrogation of Th17 inducing bacteria in the gut microbiota was linked to increased Foxp3+ T regulatory cells in the lamina propria. These findings implicate gut microbiota composition in the induction of Th17 cells and the regulation of Th17:Treg balance in the lamina propria; this in turn suggests that certain populations of bacteria influence host defense and predisposition to inflammatory bowel diseases105. A subsequent investigation narrowed the search for Th17 inducing bacteria down to segmented filamentous bacteria (SFB). Germ-free mice were used as a model for Th17 deficient mice; the colonization of SFB in these mice led to the expression of IL-17 and IL-22 in the CD4+ T cells found in the intestine lamina propria. SFB colonization was also associated with a more potent host defense against Citrobacter rodentium, an intestinal pathogen. SFB is the first specific microbiota component that has been linked to Th17 cell differentiation, an important step in the still-growing understanding of the commensal mechanisms that shape host immunity106. The discovery that microbiota induce CD4+ T cells expressing IL-17 arouses speculation that alcoholic liver disease can be curbed through antibiotics that target specific microbiota components. However, studies have found that germ free mice associated with immune deficiency exhibit elevated levels of cirrhosis compared to those with active microbiota. Given this, the implications commensal bacteria carry for alcoholic liver disease97, as well as any roles they may hold in its treatment107, have yet to be conclusively defined.

IV. Conclusion

Considerable progress has been made in our understanding of the effects of alcohol on liver function, brain function, intestinal permeability, composition of the gut microbiota, and dysregulation of immune responses. However, we are still far from achieving a comprehensive understanding of the systemic interactions between affected organs, and mechanisms underlying pathological changes associated with chronic alcohol abuse. Therefore, further interdisciplinary collaborative studies are required to identify targets which mediate a crosstalk among injured organs, and that can either protect from or exacerbate alcohol-induced systemic multi-organ damage. IL-17 signaling may function as one of these potential targets, and more studies are required to address this question. The new animal models described above might provide new insights into the mechanism of IL-17 signaling in ALD and identify novel therapeutic targets.

Acknowledgments

Grant support: Supported by the National Institutes of Health (DK088837, U01AA022614, AI0777802 P50 AA011999)

Abbreviations

Col

collagen α1(I)

α-SMA

α-smooth muscle actin

qHSCs

quiescent Hepatic Stellate Cells

aHSCs

activated Hepatic Stellate Cells

iHSCs

inactivated Hepatic Stellate Cells

Footnotes

Conflict of Interest

Hsiao-Yen Ma, Jun Xu, Xiao Liu, Yunheng Zhu, Bin Gao, Michael Karin, Hidekazu Tsukamoto, Dilip V. Jeste, Igor Grant, Amanda J Roberts, Candice Contet, Cedric Geoffroy, Binhai Zheng, David Brenner and Tatiana Kisseleva declare that they have no conflict of interest.

Human and Animal Rights and Informed Consent

Animal studies are approved by IACUC Protocol S07088 at UCSD

References

Papers of particular interest, published recently, have been highlighted as:

*Of importance

**Of major importance

  • 1**.Gao B, Bataller R. Alcoholic liver disease: pathogenesis and new therapeutic targets. Gastroenterology. 2011;141:1572–1585. doi: 10.1053/j.gastro.2011.09.002. This review described the pathological process of alcoholic liver diseases from steatosis, alcoholic hepatitis, alcoholic fibrosis to the end stage hepatocellular carcinoma. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2**.Xu J, et al. New Approaches for Studying Alcoholic Liver Disease. Current pathobiology reports. 2014;2:171–183. doi: 10.1007/s40139-014-0053-z. This review summarized the progression of alcoholic liver diseases and proposed several potential targets for ALD treatment. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3**.Bertola AMS, Ki S, Wang H, Gao B. Mouse chronic plus binge ethanol feeding model (the NIAAA model) Nature Protocol. 2013 doi: 10.1038/nprot.2013.032. in press. This paper described the method of chronic alcohol consumption and single high dose of alcohol binge, causing more severe steatosis and neutrophil infiltration than chronic alcohol feeding alone. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.O’Shea RS, Dasarathy S, McCullough AJ Practice Guideline Committee of the American Association for the Study of Liver D, Practice Parameters Committee of the American College of G. Alcoholic liver disease. Hepatology. 2010;51:307–328. doi: 10.1002/hep.23258. [DOI] [PubMed] [Google Scholar]
  • 5.Lemmers A, et al. The interleukin-17 pathway is involved in human alcoholic liver disease. Hepatology. 2009;49:646–657. doi: 10.1002/hep.22680. [DOI] [PubMed] [Google Scholar]
  • 6.Kisseleva T, Brenner DA. Mechanisms of fibrogenesis. Experimental biology and medicine. 2008;233:109–122. doi: 10.3181/0707-MR-190. [DOI] [PubMed] [Google Scholar]
  • 7.Bataller R, Brenner DA. Liver fibrosis. J Clin Invest. 2005;115:209–218. doi: 10.1172/JCI24282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Kisseleva T, Brenner DA. Hepatic stellate cells and the reversal of fibrosis. Journal of gastroenterology and hepatology. 2006;21(Suppl 3):S84–87. doi: 10.1111/j.1440-1746.2006.04584.x. [DOI] [PubMed] [Google Scholar]
  • 9.Kalluri R, Neilson EG. Epithelial-mesenchymal transition and its implications for fibrosis. J Clin Invest. 2003;112:1776–1784. doi: 10.1172/JCI20530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Gomperts BN, Strieter RM. Fibrocytes in lung disease. Journal of leukocyte biology. 2007;82:449–456. doi: 10.1189/jlb.0906587. [DOI] [PubMed] [Google Scholar]
  • 11.Fallowfield JA, et al. Scar-associated macrophages are a major source of hepatic matrix metalloproteinase-13 and facilitate the resolution of murine hepatic fibrosis. J Immunol. 2007;178:5288–5295. doi: 10.4049/jimmunol.178.8.5288. [DOI] [PubMed] [Google Scholar]
  • 12.Ueno A, et al. Mouse intragastric infusion (iG) model. Nat Protoc. 2012;7:771–781. doi: 10.1038/nprot.2012.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13**.Kisseleva T, et al. Myofibroblasts revert to an inactive phenotype during regression of liver fibrosis. Proc Natl Acad Sci U S A. 2012;109:9448–9453. doi: 10.1073/pnas.1201840109. This paper reported a novel alcohol feeding protocol by implanting gastrostomy catheter into gastrointestinal tract to create the alcoholic liver disease model. This ALD model is characterized by elevated alanine aminotranferase levels and severe hepatic steatosis. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Mello T, Ceni E, Surrenti C, Galli A. Alcohol induced hepatic fibrosis: role of acetaldehyde. Molecular aspects of medicine. 2008;29:17–21. doi: 10.1016/j.mam.2007.10.001. [DOI] [PubMed] [Google Scholar]
  • 15*.Brandon-Warner E, Walling TL, Schrum LW, McKillop IH. Chronic ethanol feeding accelerates hepatocellular carcinoma progression in a sex-dependent manner in a mouse model of hepatocarcinogenesis. Alcohol Clin Exp Res. 2012;36:641–653. doi: 10.1111/j.1530-0277.2011.01660.x. This paper described the mouse model with alcohol induced hepatocellular carcinoma development. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Alison MR. Liver stem cells: implications for hepatocarcinogenesis. Stem cell reviews. 2005;1:253–260. doi: 10.1385/SCR:1:3:253. [DOI] [PubMed] [Google Scholar]
  • 17.Sell S. Is there a liver stem cell? Cancer Res. 1990;50:3811–3815. [PubMed] [Google Scholar]
  • 18.Wu XZ, Chen D. Origin of hepatocellular carcinoma: role of stem cells. Journal of gastroenterology and hepatology. 2006;21:1093–1098. doi: 10.1111/j.1440-1746.2006.04485.x. [DOI] [PubMed] [Google Scholar]
  • 19.Shen Y, Cao D. Hepatocellular carcinoma stem cells: origins and roles in hepatocarcinogenesis and disease progression. Frontiers in bioscience. 2012;4:1157–1169. doi: 10.2741/e448. [DOI] [PubMed] [Google Scholar]
  • 20.Duncan AW, Dorrell C, Grompe M. Stem cells and liver regeneration. Gastroenterology. 2009;137:466–481. doi: 10.1053/j.gastro.2009.05.044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Sell S. Heterogeneity and plasticity of hepatocyte lineage cells. Hepatology. 2001;33:738–750. doi: 10.1053/jhep.2001.21900. [DOI] [PubMed] [Google Scholar]
  • 22**.Naugler WE, et al. Gender disparity in liver cancer due to sex differences in MyD88-dependent IL-6 production. Science. 2007;317:121–124. doi: 10.1126/science.1140485. The paper discovered the differential expression of IL-6 between male and female mice, when they were under diethylnitrosamine (DEN) challenging. It explained the gender disparity in liver cancer. [DOI] [PubMed] [Google Scholar]
  • 23.Gu FM, et al. IL-17 induces AKT-dependent IL-6/JAK2/STAT3 activation and tumor progression in hepatocellular carcinoma. Molecular cancer. 2011;10:150. doi: 10.1186/1476-4598-10-150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Li J, et al. Interleukin 17A promotes hepatocellular carcinoma metastasis via NF-kB induced matrix metalloproteinases 2 and 9 expression. PLoS One. 2011;6:e21816. doi: 10.1371/journal.pone.0021816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Jiang R, et al. Interleukin-22 promotes human hepatocellular carcinoma by activation of STAT3. Hepatology. 2011;54:900–909. doi: 10.1002/hep.24486. This paper highlight the alteration of STAT3 signaling during liver disease progression, and propose that small molecues activate STAT3 can be potential therapeutic targets for liver diseases. [DOI] [PubMed] [Google Scholar]
  • 26*.Wang H, Lafdil F, Kong X, Gao B. Signal transducer and activator of transcription 3 in liver diseases: a novel therapeutic target. Int J Biol Sci. 2011;7:536–550. doi: 10.7150/ijbs.7.536. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Yu H, Pardoll D, Jove R. STATs in cancer inflammation and immunity: a leading role for STAT3. Nature reviews. Cancer. 2009;9:798–809. doi: 10.1038/nrc2734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Sicklick JK, et al. Dysregulation of the Hedgehog pathway in human hepatocarcinogenesis. Carcinogenesis. 2006;27:748–757. doi: 10.1093/carcin/bgi292. [DOI] [PubMed] [Google Scholar]
  • 29**.Karin M. NF-kappaB as a critical link between inflammation and cancer. Cold Spring Harbor perspectives in biology. 2009;1:a000141. doi: 10.1101/cshperspect.a000141. NF-kappaB is activated by pro-inflammatory cytokines IL-17 and TNFs from activated macrophages and lymphocytes. Downstream genes of NF-kappaB promote cancer cell proliferation and survival. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30*.Sakurai T, et al. Hepatocyte necrosis induced by oxidative stress and IL-1 alpha release mediate carcinogen-induced compensatory proliferation and liver tumorigenesis. Cancer Cell. 2008;14:156–165. doi: 10.1016/j.ccr.2008.06.016. Carcinogenesis effect of Ikk-beta is mediated by IL-1α releasing and ROS acuumulation. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.White BD, Chien AJ, Dawson DW. Dysregulation of Wnt/beta-catenin signaling in gastrointestinal cancers. Gastroenterology. 2012;142:219–232. doi: 10.1053/j.gastro.2011.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Sell S, Osborn K, Leffert HL. Autoradiography of “oval cells” appearing rapidly in the livers of rats fed N-2-fluorenylacetamide in a choline devoid diet. Carcinogenesis. 1981;2:7–14. doi: 10.1093/carcin/2.1.7. [DOI] [PubMed] [Google Scholar]
  • 33.Carpentier R, et al. Embryonic ductal plate cells give rise to cholangiocytes, periportal hepatocytes, and adult liver progenitor cells. Gastroenterology. 2011;141:1432–1438. 1438 e1431–1434. doi: 10.1053/j.gastro.2011.06.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Dorrell C, et al. Prospective isolation of a bipotential clonogenic liver progenitor cell in adult mice. Genes Dev. 2011;25:1193–1203. doi: 10.1101/gad.2029411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Kopp JL, et al. Sox9+ ductal cells are multipotent progenitors throughout development but do not produce new endocrine cells in the normal or injured adult pancreas. Development. 2011;138:653–665. doi: 10.1242/dev.056499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Shin S, et al. Foxl1-Cre-marked adult hepatic progenitors have clonogenic and bilineage differentiation potential. Genes Dev. 2011;25:1185–1192. doi: 10.1101/gad.2027811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Dorrell C, et al. Surface markers for the murine oval cell response. Hepatology. 2008;48:1282–1291. doi: 10.1002/hep.22468. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Lemaigre FP. Mechanisms of liver development: concepts for understanding liver disorders and design of novel therapies. Gastroenterology. 2009;137:62–79. doi: 10.1053/j.gastro.2009.03.035. [DOI] [PubMed] [Google Scholar]
  • 39.Feng D, et al. Interleukin-22 promotes proliferation of liver stem/progenitor cells in mice and patients with chronic hepatitis B virus infection. Gastroenterology. 2012;143:188–198 e187. doi: 10.1053/j.gastro.2012.03.044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Zhao X, et al. Zhongguo Zhong yao za zhi = Zhongguo zhongyao zazhi = China journal of Chinese materia medica. 2012;37:515–519. [PubMed] [Google Scholar]
  • 41.de la Monte SM, Longato L, Tong M, DeNucci S, Wands JR. The liver-brain axis of alcohol-mediated neurodegeneration: role of toxic lipids. International journal of environmental research and public health. 2009;6:2055–2075. doi: 10.3390/ijerph6072055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Crews FT, Vetreno RP. Neuroimmune basis of alcoholic brain damage. International review of neurobiology. 2014;118:315–357. doi: 10.1016/B978-0-12-801284-0.00010-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Sutherland GT, Sheedy D, Kril JJ. Neuropathology of alcoholism. Handbook of clinical neurology. 2014;125:603–615. doi: 10.1016/B978-0-444-62619-6.00035-5. [DOI] [PubMed] [Google Scholar]
  • 44.Erdozain AM, et al. Alcohol-related brain damage in humans. PLoS One. 2014;9:e93586. doi: 10.1371/journal.pone.0093586. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Szabo G, Lippai D. Converging actions of alcohol on liver and brain immune signaling. International review of neurobiology. 2014;118:359–380. doi: 10.1016/B978-0-12-801284-0.00011-7. [DOI] [PubMed] [Google Scholar]
  • 46.Weaver CT, Harrington LE, Mangan PR, Gavrieli M, Murphy KM. Th17: an effector CD4 T cell lineage with regulatory T cell ties. Immunity. 2006;24:677–688. doi: 10.1016/j.immuni.2006.06.002. [DOI] [PubMed] [Google Scholar]
  • 47.Steinman L. A brief history of T(H)17, the first major revision in the T(H)1/T(H)2 hypothesis of T cell-mediated tissue damage. Nat Med. 2007;13:139–145. doi: 10.1038/nm1551. [DOI] [PubMed] [Google Scholar]
  • 48.Ivanov II, et al. Cell. 2006;126:1121–1133. doi: 10.1016/j.cell.2006.07.035. [DOI] [PubMed] [Google Scholar]
  • 49.Yang XO, et al. T helper 17 lineage differentiation is programmed by orphan nuclear receptors ROR alpha and ROR gamma. Immunity. 2008;28:29–39. doi: 10.1016/j.immuni.2007.11.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Bettelli E, et al. Reciprocal developmental pathways for the generation of pathogenic effector TH17 and regulatory T cells. Nature. 2006;441:235–238. doi: 10.1038/nature04753. [DOI] [PubMed] [Google Scholar]
  • 51.Mangan PR, et al. Transforming growth factor-beta induces development of the T(H)17 lineage. Nature. 2006;441:231–234. doi: 10.1038/nature04754. [DOI] [PubMed] [Google Scholar]
  • 52.Kolls JK, Linden A. Interleukin-17 family members and inflammation. Immunity. 2004;21:467–476. doi: 10.1016/j.immuni.2004.08.018. [DOI] [PubMed] [Google Scholar]
  • 53.Iwakura Y, Ishigame H, Saijo S, Nakae S. Functional specialization of interleukin-17 family members. Immunity. 34:149–162. doi: 10.1016/j.immuni.2011.02.012. [DOI] [PubMed] [Google Scholar]
  • 54.Yao Z, et al. Herpesvirus Saimiri encodes a new cytokine, IL-17, which binds to a novel cytokine receptor. Immunity. 1995;3:811–821. doi: 10.1016/1074-7613(95)90070-5. [DOI] [PubMed] [Google Scholar]
  • 55.Andoh A, et al. IL-17 selectively down-regulates TNF-alpha-induced RANTES gene expression in human colonic subepithelial myofibroblasts. J Immunol. 2002;169:1683–1687. doi: 10.4049/jimmunol.169.4.1683. [DOI] [PubMed] [Google Scholar]
  • 56.Subramaniam SV, Cooper RS, Adunyah SE. Evidence for the involvement of JAK/STAT pathway in the signaling mechanism of interleukin-17. Biochem Biophys Res Commun. 1999;262:14–19. doi: 10.1006/bbrc.1999.1156. [DOI] [PubMed] [Google Scholar]
  • 57.Miossec P, Kolls JK. Targeting IL-17 and TH17 cells in chronic inflammation. Nature reviews. Drug discovery. 2012;11:763–776. doi: 10.1038/nrd3794. [DOI] [PubMed] [Google Scholar]
  • 58.Ye P, et al. Requirement of interleukin 17 receptor signaling for lung CXC chemokine and granulocyte colony-stimulating factor expression, neutrophil recruitment, and host defense. J Exp Med. 2001;194:519–527. doi: 10.1084/jem.194.4.519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Shen F, Gaffen SL. Structure-function relationships in the IL-17 receptor: implications for signal transduction and therapy. Cytokine. 2008;41:92–104. doi: 10.1016/j.cyto.2007.11.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Gaffen SL. Structure and signalling in the IL-17 receptor family. Nat Rev Immunol. 2009;9:556–567. doi: 10.1038/nri2586. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Faust SM, et al. Role of T cell TGFbeta signaling and IL-17 in allograft acceptance and fibrosis associated with chronic rejection. J Immunol. 2009;183:7297–7306. doi: 10.4049/jimmunol.0902446. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Wilson MS, et al. Bleomycin and IL-1beta-mediated pulmonary fibrosis is IL-17A dependent. J Exp Med. 207:535–552. doi: 10.1084/jem.20092121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Longhi MS, et al. Impairment of CD4(+)CD25(+) regulatory T-cells in autoimmune liver disease. J Hepatol. 2004;41:31–37. doi: 10.1016/j.jhep.2004.03.008. [DOI] [PubMed] [Google Scholar]
  • 64.Ge J, et al. Implication of Th17 and Th1 cells in patients with chronic active hepatitis B. J Clin Immunol. 30:60–67. doi: 10.1007/s10875-009-9328-2. [DOI] [PubMed] [Google Scholar]
  • 65.Affo S, et al. Transcriptome analysis identifies TNF superfamily receptors as potential therapeutic targets in alcoholic hepatitis. Gut. 2012 doi: 10.1136/gutjnl-2011-301146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Grivennikov SI, et al. Adenoma-linked barrier defects and microbial products drive IL-23/IL-17-mediated tumour growth. Nature. 2012;491:254–258. doi: 10.1038/nature11465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67**.Meng F, et al. Interleukin-17 signaling in inflammatory, Kupffer cells, and hepatic stellate cells exacerbates liver fibrosis in mice. Gastroenterology. 2012;143:765–776. e761–763. doi: 10.1053/j.gastro.2012.05.049. In response to liver injury (hepatotoxicity and billiary obstruction), IL-17A expression is increased and IL-17A directly activates hepatic stellate cells by STAT3 signaling. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Parham C, et al. A receptor for the heterodimeric cytokine IL-23 is composed of IL-12Rbeta1 and a novel cytokine receptor subunit, IL-23R. J Immunol. 2002;168:5699–5708. doi: 10.4049/jimmunol.168.11.5699. [DOI] [PubMed] [Google Scholar]
  • 69.Cua DJ, et al. Interleukin-23 rather than interleukin-12 is the critical cytokine for autoimmune inflammation of the brain. Nature. 2003;421:744–748. doi: 10.1038/nature01355. [DOI] [PubMed] [Google Scholar]
  • 70.Langrish CL, et al. IL-23 drives a pathogenic T cell population that induces autoimmune inflammation. J Exp Med. 2005;201:233–240. doi: 10.1084/jem.20041257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Zhang JY, et al. Interleukin-17-producing CD4(+) T cells increase with severity of liver damage in patients with chronic hepatitis B. Hepatology. 2010;51:81–91. doi: 10.1002/hep.23273. [DOI] [PubMed] [Google Scholar]
  • 72**.Gao B. Hepatoprotective and anti-inflammatory cytokines in alcoholic liver disease. Journal of gastroenterology and hepatology. 2012;27(Suppl 2):89–93. doi: 10.1111/j.1440-1746.2011.07003.x. IL-22 treatment is a potential therapeutic option for treating severe forms of alcoholic liver disease because of its antioxidant, antiapoptotic, antisteatotic, proliferative, and antimicrobial effects, as well as the potential added benefit of few side effects. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Langowski JL, et al. IL-23 promotes tumour incidence and growth. Nature. 2006;442:461–465. doi: 10.1038/nature04808. [DOI] [PubMed] [Google Scholar]
  • 74.Li J, et al. Interleukin 23 promotes hepatocellular carcinoma metastasis via NF-kappa B induced matrix metalloproteinase 9 expression. PLoS One. 2012;7:e46264. doi: 10.1371/journal.pone.0046264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Schaalan MF, Mohamed WA, Amin HH. Vitamin D deficiency: correlation to interleukin-17, interleukin-23 and PIIINP in hepatitis C virus genotype 4. World J Gastroenterol. 2012;18:3738–3744. doi: 10.3748/wjg.v18.i28.3738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Xu Y, et al. IL-23R polymorphisms, HBV infection, and risk of hepatocellular carcinoma in a high-risk Chinese population. J Gastroenterol. 2012 doi: 10.1007/s00535-012-0620-1. [DOI] [PubMed] [Google Scholar]
  • 77.Hall AO, Silver JS, Hunter CA. The immunobiology of IL-27. Advances in immunology. 2012;115:1–44. doi: 10.1016/B978-0-12-394299-9.00001-1. [DOI] [PubMed] [Google Scholar]
  • 78.Diveu C, et al. IL-27 blocks RORc expression to inhibit lineage commitment of Th17 cells. J Immunol. 2009;182:5748–5756. doi: 10.4049/jimmunol.0801162. [DOI] [PubMed] [Google Scholar]
  • 79.Wirtz S, et al. Protection from lethal septic peritonitis by neutralizing the biological function of interleukin 27. J Exp Med. 2006;203:1875–1881. doi: 10.1084/jem.20060471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Hibbert L, Pflanz S, De Waal Malefyt R, Kastelein RA. IL-27 and IFN-alpha signal via Stat1 and Stat3 and induce T-Bet and IL-12Rbeta2 in naive T cells. J Interferon Cytokine Res. 2003;23:513–522. doi: 10.1089/10799900360708632. [DOI] [PubMed] [Google Scholar]
  • 81.Villarino A, et al. The IL-27R (WSX-1) is required to suppress T cell hyperactivity during infection. Immunity. 2003;19:645–655. doi: 10.1016/s1074-7613(03)00300-5. [DOI] [PubMed] [Google Scholar]
  • 82.Fort MM, et al. IL-25 induces IL-4, IL-5, and IL-13 and Th2-associated pathologies in vivo. Immunity. 2001;15:985–995. doi: 10.1016/s1074-7613(01)00243-6. [DOI] [PubMed] [Google Scholar]
  • 83.Hurst SD, et al. New IL-17 family members promote Th1 or Th2 responses in the lung: in vivo function of the novel cytokine IL-25. J Immunol. 2002;169:443–453. doi: 10.4049/jimmunol.169.1.443. [DOI] [PubMed] [Google Scholar]
  • 84.Petersen BC, Lukacs NW. IL-17A and IL-25: therapeutic targets for allergic and exacerbated asthmatic disease. Future medicinal chemistry. 2012;4:833–836. doi: 10.4155/fmc.12.39. [DOI] [PubMed] [Google Scholar]
  • 85.Rickel EA, et al. Identification of functional roles for both IL-17RB and IL-17RA in mediating IL-25-induced activities. J Immunol. 2008;181:4299–4310. doi: 10.4049/jimmunol.181.6.4299. [DOI] [PubMed] [Google Scholar]
  • 86.Lee J, et al. IL-17E, a novel proinflammatory ligand for the IL-17 receptor homolog IL-17Rh1. J Biol Chem. 2001;276:1660–1664. doi: 10.1074/jbc.M008289200. [DOI] [PubMed] [Google Scholar]
  • 87.Liu C, et al. A CC’ loop decoy peptide blocks the interaction between Act1 and IL-17RA to attenuate IL-17- and IL-25-induced inflammation. Science signaling. 2011;4:ra72. doi: 10.1126/scisignal.2001843. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Liu Y, Munker S, Mullenbach R, Weng HL. IL-13 Signaling in Liver Fibrogenesis. Frontiers in immunology. 2012;3:116. doi: 10.3389/fimmu.2012.00116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Kleinschek MA, et al. IL-25 regulates Th17 function in autoimmune inflammation. J Exp Med. 2007;204:161–170. doi: 10.1084/jem.20061738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Zaph C, et al. Commensal-dependent expression of IL-25 regulates the IL-23-IL-17 axis in the intestine. J Exp Med. 2008;205:2191–2198. doi: 10.1084/jem.20080720. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Kawanokuchi J, et al. Production and functions of IL-17 in microglia. Journal of neuroimmunology. 2008;194:54–61. doi: 10.1016/j.jneuroim.2007.11.006. [DOI] [PubMed] [Google Scholar]
  • 92**.Kebir H, et al. Human TH17 lymphocytes promote blood-brain barrier disruption and central nervous system inflammation. Nat Med. 2007;13:1173–1175. doi: 10.1038/nm1651. This paper reported the potential importance of TH17 lymphocyte infiltration into the CNS and these lymphocytes’ consequent involvement in lesion formation in multiple sclerosis and experimental autoimmune encephalomyelitis. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Liu Q, et al. Interleukin-17 inhibits Adult Hippocampal Neurogenesis. Scientific reports. 2014;4:7554. doi: 10.1038/srep07554. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94*.Shichita T, et al. Pivotal role of cerebral interleukin-17-producing gammadeltaT cells in the delayed phase of ischemic brain injury. Nat Med. 2009;15:946–950. doi: 10.1038/nm.1999. IL-17A producing lymphocytes infiltrated into ischemia-reperfusion injured brain. The IL-17A producing gammadelta T cells played a pivotal role in the late stage of ischemic brain injury. [DOI] [PubMed] [Google Scholar]
  • 95.Hill F, Kim CF, Gorrie CA, Moalem-Taylor G. Interleukin-17 deficiency improves locomotor recovery and tissue sparing after spinal cord contusion injury in mice. Neuroscience letters. 2011;487:363–367. doi: 10.1016/j.neulet.2010.10.057. [DOI] [PubMed] [Google Scholar]
  • 96.Matsui T, Yoshida Y, Yanagihara M, Suenaga H. Hypothermia at 35 degrees C reduces the time-dependent microglial production of pro-inflammatory and anti-inflammatory factors that mediate neuronal cell death. Neurocritical care. 2014;20:301–310. doi: 10.1007/s12028-013-9911-5. [DOI] [PubMed] [Google Scholar]
  • 97.Yan AW, et al. Enteric dysbiosis associated with a mouse model of alcoholic liver disease. Hepatology. 2011;53:96–105. doi: 10.1002/hep.24018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Kisseleva T. Does interleukin-17 play the villain in nonalcoholic steatohepatitis? Hepatology. 2014;59:1671–1672. doi: 10.1002/hep.26955. [DOI] [PubMed] [Google Scholar]
  • 99.Scholten D, et al. Migration of fibrocytes in fibrogenic liver injury. Am J Pathol. 2011;179:189–198. doi: 10.1016/j.ajpath.2011.03.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Jeste DV, Depp CA. Positive mental aging. The American journal of geriatric psychiatry : official journal of the American Association for Geriatric Psychiatry. 2010;18:1–3. doi: 10.1097/JGP.0b013e3181c3ef09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Haynes L, Maue AC. Effects of aging on T cell function. Current opinion in immunology. 2009;21:414–417. doi: 10.1016/j.coi.2009.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Lim MA, et al. Increased Th17 differentiation in aged mice is significantly associated with high IL-1beta level and low IL-2 expression. Experimental gerontology. 2014;49:55–62. doi: 10.1016/j.exger.2013.10.006. [DOI] [PubMed] [Google Scholar]
  • 103.Hartigan-O’Connor DJ, Hirao LA, McCune JM, Dandekar S. Th17 cells and regulatory T cells in elite control over HIV and SIV. Current opinion in HIV and AIDS. 2011;6:221–227. doi: 10.1097/COH.0b013e32834577b3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Gongvatana A, et al. A history of alcohol dependence augments HIV-associated neurocognitive deficits in persons aged 60 and older. Journal of neurovirology. 2014;20:505–513. doi: 10.1007/s13365-014-0277-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Ivanov II, et al. Cell Host Microbe. 2008;4:337–349. doi: 10.1016/j.chom.2008.09.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Ivanov II, et al. Cell. 2009;139:485–498. doi: 10.1016/j.cell.2009.09.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Fouts DE, Torralba M, Nelson KE, Brenner DA, Schnabl B. Bacterial translocation and changes in the intestinal microbiome in mouse models of liver disease. J Hepatol. 2012;56:1283–1292. doi: 10.1016/j.jhep.2012.01.019. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES