CC BY-NC-ND 4.0 · Semin Liver Dis 2023; 43(01): 013-023
DOI: 10.1055/s-0042-1760306
Review Article

Update on Hepatobiliary Plasticity

Minwook Kim*
1   Department of Developmental Biology, University of Pittsburgh School of Medicine, Pittsburgh, Pennsylvania
,
Fatima Rizvi*
2   Department of Medicine, Gastroenterology Section, Center for Regenerative Medicine, Boston University and Boston Medical Center, Boston, Massachusetts
,
Donghun Shin
1   Department of Developmental Biology, University of Pittsburgh School of Medicine, Pittsburgh, Pennsylvania
,
Valerie Gouon-Evans
2   Department of Medicine, Gastroenterology Section, Center for Regenerative Medicine, Boston University and Boston Medical Center, Boston, Massachusetts
› Author Affiliations
Funding This work was supported by the NIH/NIDDK awards R01DK133404 to V.G-E. and D.S. and R01DK101426 and R01DK132014 to D.S.
 


Abstract

The liver field has been debating for decades the contribution of the plasticity of the two epithelial compartments in the liver, hepatocytes and biliary epithelial cells (BECs), to derive each other as a repair mechanism. The hepatobiliary plasticity has been first observed in diseased human livers by the presence of biphenotypic cells expressing hepatocyte and BEC markers within bile ducts and regenerative nodules or budding from strings of proliferative BECs in septa. These observations are not surprising as hepatocytes and BECs derive from a common fetal progenitor, the hepatoblast, and, as such, they are expected to compensate for each other's loss in adults. To investigate the cell origin of regenerated cell compartments and associated molecular mechanisms, numerous murine and zebrafish models with ability to trace cell fates have been extensively developed. This short review summarizes the clinical and preclinical studies illustrating the hepatobiliary plasticity and its potential therapeutic application.


#

During the last three decades, the liver field has been questioning “liver progenitor cell (LPC)” features and the plasticity of the two liver epithelial cell compartments, hepatocytes and biliary epithelial cells (BECs), for their therapeutic ability to regenerate the liver. The term “LPC” is used here to describe cells that have the potential to differentiate into hepatocytes and BECs and that usually express markers from both epithelial compartments. Early developmental studies undoubtedly support the existence of a common fetal precursor of BECs and hepatocytes, the hepatoblast. Although there is no evidence for a direct relationship between fetal hepatoblasts and adult LPCs, a growing literature reports some functional and phenotypic similarities between them.[1] Both can self-renew and differentiate into hepatocytes and BECs, and they both share cell surface markers.[2] [3] [4] [5] [6] [7] [8] [9] [10] [11] Given this critical question, it is not surprising to notice that studies on BEC-to-hepatocyte or hepatocyte-to-BEC conversion have garnered increased attention in the liver field. Yet, outstanding questions remain unanswered such as the following ([Fig. 1]): (1) Are there some resident LPCs ([Fig. 1A]) that have a potential to proliferate and differentiate into both BECs and hepatocytes? If so, as identified in intrahepatic small hepatic bile ductules (hepatic stem/progenitor cells, HpSCs) or large intra- and extra-hepatic bile ducts (biliary tree stem/progenitor cells, BTSCs),[12] are they phenotypically distinct from BECs and can we identify them with specific markers? (2) LPCs may be instead facultative ([Fig. 1B]) as they emerge only after liver injury. In this scenario, mature hepatocytes or BECs dedifferentiate into LPCs, reminiscent of fetal hepatoblasts. This process is succeeded by the proliferation of facultative LPCs and differentiation into the mature cell type in demand, hepatocytes when the hepatocyte pool is damaged or lost, or BECs in cholangiopathies. Yet, again, one can question whether all hepatocytes or BECs possess this capability. (3) Another possibility is that the mature parenchymal cells, hepatocytes and BECs, exhibit cell plasticity to directly give rise to each other by transdifferentiation ([Fig. 1C]) without transitioning into an intermediate LPC stage. Studies investigating adult liver regeneration in animal models as well as evidence from analyses of human diseased liver specimens may support one theory or the other; however, a careful assessment of the literature as well as continuous investigations will help reach some consensuses. In this short review, we have specifically focused our attention on the facultative LPCs, summarized clinical and preclinical studies to identify potential explanations of regenerative mechanisms in adult livers, and recognized gaps in the field that are limiting the therapeutic application of hepatobiliary plasticity to treat liver diseases. In an attempt to address some of these questions, we focus on the evidence of cell plasticity of hepatocytes and BECs in human liver diseases in the first section and compare them to animal study findings in the second and third sections to finally discuss the possible modulations of hepatobiliary plasticity as a therapeutic intervention and their limitations in the last two sections.

Zoom Image
Fig. 1 Mechanisms of hepatobiliary plasticity. Mature hepatocytes and BECs are known to divide to overcome minor injuries (no plasticity, left panel); however, when injuries are chronic or more severe, plasticity of the hepatobiliary compartment is observed (plasticity, right panel). In this case, various options have been proposed: (A) Resident LPCs self-renew and differentiate into both hepatocytes and BECs, (B) Facultative LPCs emerge following liver injury, self-renew and differentiate as the resident LPCs do, and (C) mature hepatocytes and BECs transdifferentiate into each other without transitioning through a LPC stage. BECs, biliary epithelial cells; LPC, liver progenitor cell.

Evidence of Hepatobiliary Plasticity in Humans

The process of ductular reactions (DRs), which involves the expansion of BECs, is a hallmark of all chronic and acute human liver diseases.[13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] This suggests an alternative BEC-driven liver regeneration to overcome an exhausted hepatocyte-driven repair, by which BECs proliferate and contribute to hepatocytes. The observation of hepatocytes expressing the BEC marker EpCAM within highly proliferative DR areas in advanced human cirrhotic livers[24] and that of hepatocyte-like cells expressing the central vein hepatocytic marker glutamine synthetase budding from BECs within the DRs[25] [26] [27] [28] [29] could represent the contribution of BECs to de novo hepatocytes. Further evidence is illustrated by the detection of “bi-phenotypic cells”[30] or “ductular hepatocytes”[31] that express both the hepatocyte marker HNF4a[30] or HepPar1[31] and the BEC marker KRT19. Human cirrhotic liver samples frequently harbor intermediate hepatocyte-like cells with morphology and size intermediate between hepatocytes and BECs.[32] Specifically, in human cirrhotic livers, quantification of immature hepatocytes with glutamine synthetase positivity budding from KRT19+ BECs demonstrated that they represented up to 70% of hepatocytes within the septa.[25] It has been suggested that glutamine synthetase re-expression away from the central vein areas is linked to a repair process, as a recent study demonstrated that aberrant glutamine synthetase positivity adjacent to portal tracts is associated with regressed cirrhosis in humans.[33] Lin and colleagues attempted to lineage trace LPCs among the DRs in human cirrhotic liver[34] [35] using mutational analysis in mitochondrial DNA encoding cytochrome c oxidase enzyme, demonstrating that hepatocytes within monoclonal regenerative nodules descend from adjacent LPC-associated DRs. This study supported the differentiation potential of BECs in humans, suggesting the clinical application of LPC-derived hepatocytes in resolving human cirrhosis.

As in disorders of hepatocyte degeneration, biliary degenerative diseases are also associated with prominent DRs along with occurrence of intermediate hepatobiliary cells (IHBCs).[12] Cholangiopathies are associated with genetic- or immune-mediated damage to the intrahepatic or extrahepatic biliary tree, fibrotic response, and subsequent liver damage. The need to replace deteriorating BECs that are impaired in their proliferative capabilities by chronic damage elicits an alternative regenerative mechanism facilitated by hepatocyte plasticity. Many histopathological examinations have reported the expression of the BEC marker KRT7 in hepatocytes during cholangiopathies.[36] [37] [38] [39] [40] [41] [42] In cases of Alagille syndrome as well as biliary atresia, the number of IHBCs co-expressing the BEC markers KRT7 or HNF6 and the hepatocyte markers LKM-1, BSEP, or HNF4a is significantly increased.[43] In both primary biliary cholangitis (PBC) and primary sclerosing cholangitis (PSC), the appearance of IHBCs increases with the stage of fibrotic damage to the tissue.[39] [44] [45] A significant number of hepatocytes express the BEC transcription factor FOXA2 in the late stage of PBC and biliary obstruction.[46] However, the DR phenotype in PSC differs from that in PBC.[44] In cirrhotic PSC, there are lesser reactive ductules due to their lower proliferative index, but more EpCAM +/Hep-Par1+ newly derived hepatocytes.[44] It seems, since BECs in biliary cholangiopathies often show senescent phenotypes[47] [48] [49] [50] [51] [52] and lower proliferative index,[44] [46] the transition of hepatocytes into BEC-like cells or IHBCs could be a way to compensate for deteriorating biliary function. The prominent expression of OV6, known to characterize ductal plates, bile ducts, and ductules in fetal tissue, in periseptal hepatocytes in the liver of biliary atresia patients,[53] as observed in the liver of PBC and PSC patients,[54] further supports hepatocyte metaplasia,[55] transitioning from a mature cell to an immature LPC stage in cholangiopathies.

The clinical studies illustrate both cell plasticity processes: hepatocytes give rise to BECs when the BEC compartment is compromised, and BECs give rise to hepatocytes when hepatocyte proliferation is exhausted. However, given the lack of lineage tracing strategy in humans, the question of the true identity of the origin of the progeny still stands. Recent studies using lineage tracing animal models have proven to be instrumental to specifically identify the contribution of hepatocytes and BECs to each other's compartments during various liver diseases.


#

BEC-to-Hepatocyte Conversion in Animal Models

Rodents

BEC-to-hepatocyte conversion was first examined in rats followed by mice and zebrafish. In the Solt-Farber protocol, 2-acetylaminofluorene (2-AAF) is given to rats 1 week before partial hepatectomy (PHx); PHx stimulates liver regeneration and 2-AAF suppresses hepatocyte proliferation, thereby permitting BECs to contribute to hepatocytes.[56] [57] Following the PHx, LPCs expressing BEC markers and the fetal marker AFP appeared in the portal regions and expanded. Subsequently, they lost BEC features and acquired hepatocyte features,[56] suggesting BEC-to-hepatocyte conversion. Given the lack of lineage tracing tools in rats, mice have been extensively used to prove BEC-to-hepatocyte conversion. Using both BEC- and hepatocyte-specific lineage tracing approaches ([Table 1]), it was initially reported that in mice, BECs barely contribute to hepatocytes during homeostasis and even in diverse liver injury models, including CCl4, DDC, and CDE.[58] [59] [60] [61] [62] [63] [64] [65] [66] This minimal contribution of BECs to hepatocytes raised skepticism about the significance of BEC-to-hepatocyte conversion in liver regeneration. Given the difference in the proliferation capacity of hepatocytes between the rat Solt-Farber model, in which hepatocyte proliferation was blocked, and the commonly used mouse liver injury models, novel mouse models were established in which hepatocyte proliferation was suppressed either by deleting Mdm2,[67] β1-integrin,[68] or β-catenin[69] or by overexpressing p21[68] in hepatocytes. These improved models revealed a significant contribution of BECs to hepatocytes ranging from 15 to 70%, depending on the use of either direct BEC- or indirect hepatocyte-lineage tracing mouse models. In addition to these genetic blocks of hepatocyte proliferation, long-term liver injury with 6- or 12-month administrations of DDC or thioacetamide (TAA), leading to natural impairment of hepatocyte proliferation, also induced a significant BEC-to-hepatocyte conversion.[30] These lineage tracing studies have been instrumental to demonstrate the potential of BECs to generate healthy de novo hepatocytes when hepatocyte proliferation is compromised, reflecting most human chronic liver diseases.[32] [70] [71] [72] [73] [74] [75] However, very few studies have started to elucidate the molecular mechanisms underlying BEC-to-hepatocyte conversion. The Notch–IGF1 axis[76] and Tet1[77] have been reported to control LPC proliferation during the conversion, thereby affecting the number of BEC-derived hepatocytes. We recently showed that vascular endothelial growth factor A (VEGFA) delivered to the liver via nucleoside-modified mRNA encapsulated into lipid nanoparticles[78] induced a fivefold increase in BEC-to-hepatocyte conversion using tamoxifen-inducible KRT19 and VEGFA receptor KDR lineage mouse models during acute and chronic liver injuries (unpublished data, Rizvi and Gouon-Evans, in preparation). VEGFA-mediated cell conversion may be mediated through the activation of a VEGFA receptor KDR expressed on a subset of BECs after liver injury (unpublished data, Rizvi and Gouon-Evans, in preparation). Further investigation is needed to understand the molecular mechanisms driving the cell conversion and to leverage them for therapeutic intervention.

Table 1

Summary of mouse experiments showing liver parenchymal cell plasticity

Cre lines used for lineage tracing

Injury

Genetic modulation

Contribution (○) and limitations (●)

BEC-to-HC

Foxl1-Cre

BDL,[66] DDC,[61] [63] [66] CDE[61]

○ ∼ 29% of HCs derived from Foxl1-Cre+ cells after CDE diet[61] (5% of HCs were labeled during CDE injury[61])

● Noninducible Cre line

Hnf1b-CreER

CDE[140]

○ 0.22% of HCs derived from BECs

● Rare contribution of BECs to HCs

Sox9-CreERT2

CCl4,[141] BDL,[141] MCDE,[141] DDC,[65] [141] APAP[141]

○ 1% of HCs derived from BECs[65]

● Sox9 is also expressed in a subset of periportal hepatocytes in a normal condition[60] [142]

OPN-CreERT2

CDE,[62] CCl4,[59] MCD[76]

○ 2.45% of HCs derived from BECs[62]

○ ∼ 13% of HCs derived from BECs[59]

● OPN is also expressed in other cell types[143] [144] [145]

CK19-CreERT

DDC,[30] [68] MCD,[68] TAA,[30] [68] CDE[69]

Δβ1-integrin [68]

p21 overexpression[68]

Ctnnb1-siRNA[69]

○ 9.1–10% of HCs derived from BECs[30]

○ ∼ 6.12% of HCs derived from BECs[68] [69]

● Low labeling efficiency[143] [146]

HC-to-BEC

AAV8-TBG-Cre

or AAV8-CMV-Cre

No injury

R26-LSL-NICD1 [113]

○ 23% of BECs derived from HCs assessed by KRT19 and BEC apical markers (PAR6, PKCζ, and Ac-tub)

DDC,[113] BDL[113]

○ 4.4–14.3% of BECs derived from HCs assessed by KRT19 and BEC apical markers (PAR6, PKCζ, and Ac-tub)

Alb-Cre

Retrorsine + PHx + DDC or CCl4 [112]

● Transplanted HCs converted to BECs assessed by KRT19

Mx1-Cre (induced by poly(I:C))

DDC,[112] DAPM,[112] BDL,[112] TAA,[112] CCl4 [112]

○ 1.9–20.6% of BECs derived from HCs assessed by KRT19

Alb-CreER

or Alb-CreERT2

DDC,[60] BDL,[60] TAA[147]

○ 10–11.31% of BECs derived from HCs assessed by KRT19[60]

DDC,[114] CCl4 [114]

DDC[114]

R26R-LSL-NICD1

Hes1f/f

BEC differentiation is assessed by KRT19, EpCAM, and keratin

AAV-TBG-Cre; R26-LSL-rtTA

No injury

TetO-YAPS127A [126]

TetO-YAPS127A ; Rbpjfl/f [126]

BEC differentiation is assessed by KRT19 and pan-CK

AAV8-TTR-Flp

Alb-Cre; Rbpjf/f; Hnf6f/f [127]

○ ∼ 100% of BECs derived from HCs assessed by KRT19 and wide-spectrum CK

Hnf4a-DreERT2; Sox9-CreERT2

DDC,[60] BDL[60]

○ ∼ 3.63% of BECs derived from HCs assessed by KRT19

No injury

HDTVI of CAGGS-GFP-IRES-SOX9 [148]

○ ∼ 18% of BECs derived from HCs assessed by KRT19

Abbreviations: HC, hepatocyte; DDC, 3,5-diethoxycarbonyl-1,4-dihydrocollidine; CDE, choline-deficient, ethionine-supplemented; CCl4, carbon tetrachloride; PHx, 70% partial hepatectomy; MCD, methionine- and choline-deficient; MCDE, methionine- and choline-deficient, 0.15% ethionine-supplemented; APAP, acetaminophen; BDL, bile duct ligation; DAPM, methylene dianiline; TAA, thioacetamide; LSL, loxP-Stop-loxP cassette; HDTVI, hydrodynamic tail vein injection; Ac-tub, acetylated tubulin.



#

Zebrafish

Given the strengths of zebrafish as a vertebrate model organism, including (1) rapid and external embryogenesis, (2) an easy drug administration, (3) a large number of progenies, and (4) a low maintenance cost, and similar cellular compositions in the liver between zebrafish and mammals albeit a difference between the two in the mode of connection between hepatocytes and bile ductules,[79] [80] zebrafish have been widely used for investigating liver diseases and for liver toxicology tests.[81] [82] [83] [84] [85] [86] [87] [88] [89] [90] [91] [92] [93] [94] [95] A decade ago, three groups independently developed a hepatocyte ablation model by generating the Tg(fabp10a:NTR) fish lines that express bacterial nitroreductase (NTR) specifically in hepatocytes.[91] [96] [97] Since NTR converts metronidazole (MTZ) into a cytotoxic drug, MTZ treatment ablates nearly all hepatocytes in Tg(fabp10a:NTR) fish. Following MTZ washout, the liver robustly and synchronously among animals regenerates through BEC-to-hepatocyte conversion. This liver regeneration occurs through four steps: (1) BEC-to-LPC dedifferentiation, (2) LPC proliferation, (3) LPC-to-hepatocyte differentiation, and (4) hepatocyte proliferation and maturation.[96] [97] In this model, LPCs are distinguished from BECs based on cell and nuclear shape and the expression of hepatocyte markers. Given the synchrony and robustness of BEC-driven liver regeneration in this ablation model combined with the general strengths of zebrafish as a vertebrate model organism, the ablation model has been actively used to reveal the molecular mechanisms underlying BEC-to-hepatocyte conversion. It was revealed using the zebrafish model that Notch signaling,[96] bromodomain and extraterminal (BET) proteins,[98] Dnmt1,[99] and mTORC1[100] regulate the first step of the conversion process, BEC-to-LPC dedifferentiation. It was also revealed that the second step, LPC proliferation, is positively regulated by BET proteins,[98] Tel2,[101] and Stat3[102] and negatively regulated by Notch[91] and FXR[103] signaling. Specifically, FXR activation did not only suppress LPC proliferation but also induced its death.[103] Given the correlation between LPC number and the severity of human liver diseases[21] [104] and the potential of regenerative therapy to promote LPC-to-hepatocyte differentiation in the diseased livers, the third step, LPC-to-hepatocyte differentiation, was more extensively investigated using the zebrafish model than the other steps. It was revealed that BMP signaling[105] and Tel2[101] positively regulate the third step through Tbx2b and Hhex, respectively, and that Notch signaling negatively regulates the step.[91] [96] Epigenetic regulators, Hdac1[106] and Dnmt1,[99] also positively control LPC-to-hepatocyte differentiation by repressing sox9b and tp53 expression, respectively. p53 inhibits the differentiation by suppressing BMP signaling.[99] It was also reported that FXR activation suppresses LPC-to-hepatocyte differentiation via the FXR–PTEN–PI3K–AKT–mTORC1 axis.[103] Conversely, another group reported using a similar ablation model that FXR is required for LPC-to-hepatocyte differentiation by regulating ERK1,[107] suggesting that FXR can play a dual role in this process. Regarding the last step of the conversion process, hepatocyte proliferation, and maturation, it was reported that BET and Wnt2bb regulate the hepatocyte proliferation[97] [98] and that Stat3 regulates its maturation.[102]

In addition to the hepatocyte ablation model, an oncogene-induced hepatocyte damage model has been used to study BEC-to-hepatocyte conversion, particularly LPC-to-hepatocyte differentiation. In Tg(fabp10a:pt-β-catenin) fish, hepatocyte-specific overexpression of the stable form of β-catenin triggers oncogene-induced senescence and apoptosis in hepatocytes, thereby inducing BEC-driven liver regeneration.[108] Indeed, in this model, both BECs and survived hepatocytes dedifferentiate into LPCs, and later, the LPCs differentiate into hepatocytes. Using this model, it was revealed that suppressing EGFR signaling promotes LPC-to-hepatocyte differentiation via the EGFR–ERK–SOX9 axis.[108] Importantly, this study suggests EGFR inhibitors as a potential regenerative therapeutic drug to promote LPC-to-hepatocyte differentiation in diseased livers.


#
#

Hepatocyte-to-BEC Conversion in Animal Models

Rodents

Hepatocyte-to-BEC conversion was first examined in rats[109] [110] followed by mice. Rats subjected to bile duct ligation (BDL) with pretreatment with the biliary toxin 4,4′-methylenedianiline (MDA) exhibited hepatocyte-to-BEC conversion.[109] In this model, BDL induces biliary damage and subsequent regeneration and MDA blocks BEC proliferation, which is in symmetry with the rat Solt-Farber model, in which PHx induces liver regeneration and 2-AAF blocks hepatocyte proliferation. In addition to the BDL-MDA combination, repeated administrations of MDA alone induced chronic biliary damage and hepatocyte-to-BEC conversion.[111] In these rat models, hepatocyte-derived BECs were identified based on dipeptidyl dipeptidase IV (DDPIV) expression using DDPIV-negative rats having hepatocytes from DDPIV-positive donors.

Using the Cre/loxP system for hepatocyte-lineage tracing, hepatocyte-to-BEC conversion was validated in mice with multiple liver injury models, including DDC, CCl4, MDA, and BDL[112] [113] [114] ([Table 1]). Compared with BEC-lineage tracing, hepatocyte-lineage tracing is much more robust and unquestionable owing to the availability and tracing efficiency of hepatocyte-specific Cre lines that are not activated in any other cell types. Given the essential role of Notch signaling in biliary formation during development,[115] [116] [117] it has been reported that in mice, Notch signaling controls hepatocyte-to-BEC conversion.[113] [114] [118] [119] Hepatocyte-specific deletion of Rbpj (the principal mediator of Notch signaling)[113] or Hes1 (a key effector of Notch signaling)[114] reduced the number of hepatocyte-derived BECs in mice fed a DDC diet, while hepatocyte-specific overexpression of Notch intracellular domain (NICD) induced hepatocyte-to-BEC conversion.[113] [118] [119] In addition to Notch signaling, Yap signaling plays a crucial role in both biliary development[117] [120] [121] and hepatocyte-to-BEC conversion.[122] [123] [124] [125] [126] Hepatocyte-specific deletion of Yap1 greatly reduced the number of DRs,[122] [123] while hepatocyte-specific overexpression of constitutive-active YAP1 induced hepatocyte-to-BEC conversion[124] [125] [126] through the induction of NOTCH2 and SOX9.[126]

Contrary to the prevailing thought that Notch signaling is indispensable for hepatocyte-to-BEC conversion, a Notch-independent mechanism for the conversion was recently identified.[147] In a mouse model that mimics Alagille syndrome, Alb-Cre; Rbpjf/f ; Hnf6f/f , intrahepatic peripheral bile ducts do not develop initially, but later, the bile ducts form via hepatocyte-to-BEC conversion. Additional deletion of Tgfbr2 blocked the bile duct recovery, whereas hepatocyte-specific overexpression of constitutive-active TGFBR1 promoted it, indicating the crucial role of TGFβ signaling in hepatocyte-to-BEC conversion in the absence of Notch signaling.[147]

Cholangiocarcinoma models in which hepatocyte-specific overexpression of certain oncogenes induces cholangiocarcinoma are also useful to study the molecular mechanisms of hepatocyte-to-BEC conversion, because the conversion is a prerequisite for the cancer formation. Using these models, it has been reported that not only Notch[119] [128] [129] [130] and Yap[118] [124] but also Dnmt1[118] play key roles in hepatocyte-to-BEC conversion. Particularly, NICD overexpression in hepatocytes induces the conversion through the NICD–YAP1–DNMT1 axis.[118]


#

Zebrafish

Several biliary injury models with genetic modifications causing BEC paucity were developed in zebrafish; however, hepatocyte-to-BEC conversion has not been investigated in these models.[81] [131] Given that severe liver injury is required for plasticity-mediated liver regeneration, severe biliary injury models may be needed to study hepatocyte-to-BEC conversion in zebrafish. We have recently developed a zebrafish model for the conversion, in which all regenerating BECs originate from hepatocytes.[132] Temporal Notch inhibition during BEC-driven liver regeneration triggered by hepatocyte ablation generates zebrafish that completely lack BECs in the liver. Subsequent removal of Notch inhibition permits a subset of hepatocytes to give rise to BECs. In this new zebrafish model, both Notch and Yap signaling control hepatocyte-to-BEC conversion,[132] consistent with the findings in mice.[114] [121] [126] Given the strengths of zebrafish as a vertebrate model organism, particularly chemical screening, we expect that our novel model as well as other zebrafish models to be developed will significantly contribute to a better understanding of the molecular mechanisms underlying hepatocyte-to-BEC conversion.


#
#

Regulating Hepatobiliary Plasticity as a Therapeutic Intervention

Orthotopic liver transplantation is a main curative approach for end-stage liver diseases. However, the shortage of organ donors results in many patients dying while waiting for transplantation. For such patients, there is a need for discovery of alternative therapies that could act as a bridge to support them until availability of liver donors. The development of effective cell replacement therapy could provide such a bridge and represent a promising approach to the treatment of liver diseases. Another alternative to liver transplant is leveraging the innate liver repair by harnessing mechanisms of cell plasticity.

Studies using zebrafish models of hepatocyte ablation have demonstrated key pathways that can be manipulated to promote LPC-to-hepatocyte differentiation ([Fig. 2]). Sox9b repression is important for this process[106]; hence, use of Notch inhibitor LY411575 that represses Sox9b demonstrated the enhanced induction of Hnf4a in LPCs.[133] Furthermore, pharmacological inhibition of EGFR or MEK/ERK promoted LPC-to-hepatocyte differentiation, demonstrating the prospects of the epidermal growth factor receptor (EGFR) signaling pathway as a candidate therapeutic target.[108] Manipulation of the BEC niche is shown to facilitate BEC-to-hepatocyte differentiation during chronic injury in mice.[62] Inhibition of laminin deposition using Iloprost, a synthetic analog of prostaglandin I2 known to block TGFb1-mediated fibrogenesis, enhanced the presence of differentiated hepatocytes after 3 weeks of CDE-induced liver injury in mice. However, the low efficiency of the process still posed a question about its clinical significance. We have recently demonstrated that VEGFA significantly enhanced BEC-to-hepatocyte conversion with a factor 5 to produce healthy hepatocytes in acute as well as chronically injured livers (unpublished data, Rizvi and Gouon-Evans, in preparation). Importantly, the results provide evidence of the therapeutic potential of VEGFA to harness BEC-driven liver regeneration with the use of nucleoside-modified mRNA-LNP, a tool that we have validated to express regenerative factors for preclinical therapeutic interventions in various murine liver diseases.[78] [134] Some studies have also explored pathway modulation to promote BEC differentiation to benefit biliary diseases. YAP activation in hepatocytes is required for hepatocyte-to-BEC conversion after DDC-induced liver injury.[135] YAP-mediated hepatocyte transdifferentiation was further confirmed in mouse models of alcohol exposure.[136]

Zoom Image
Fig. 2 Molecular mechanisms driving hepatobiliary plasticity. Studies in rodents and zebrafish together with observations in human specimens have revealed positive and negative molecular regulators implicated in BEC-to-hepatocytes conversion following hepatocyte injury (A) and in hepatocyte-to-BEC conversion following BEC injury (B). Factors written with capital and small letters are related to mouse and zebrafish studies, respectively. Factors with an asterisk ([*]) are related to both mouse and zebrafish studies. BEC, biliary epithelial cell.

Limitations and Concluding Remarks

The findings from the juxtaposition of both rodent and zebrafish liver injury models have indisputably revealed the ability of hepatocytes and BECs to generate each other when needed, a fact that is not that surprising as hepatocytes and BECs come from a common fetal progenitor, the hepatoblast. Even though this review largely focuses on the contribution of facultative LPCs to liver regeneration, the role of resident LPCs in this process is very likely. A critical limitation of the current lineage tracing models is that they do not discriminate between LPCs and BECs. Indeed, we are yet to discover markers specific to LPCs that are not expressed in BECs. Furthermore, with respect to cell plasticity in rodents, it is not clear whether the dedifferentiation of mature cells into facultative LPCs always precedes their conversion to different cell fates. Additional lineage tracing studies using yet-to-be discovered unique markers for LPCs, in combination with isolated LPC fate mapping investigation in ex vivo clonal cultures or following transplantation in vivo, will be instrumental to further reveal the true potential of LPCs in regenerating a damaged liver.

Interestingly, although animal studies indicate that LPC-driven liver regeneration restores liver parenchyma in liver diseases, it does not appear to benefit patients with advanced liver disease. In fact, the clinical benefits of LPCs may be questionable as the presence of DRs has been associated with poor prognosis in advanced human chronic liver diseases.[21] [32] However, a correlation between LPC numbers and disease severity in patients with chronic liver diseases may also imply that while LPCs are activated, their differentiation into hepatocytes may be ineffective. Indeed, persistent LPCs release profibrogenic factors that may induce inflammation and subsequent fibrosis, and instead aggravate the chronic liver disease.[137] [138] Yet, a recent study demonstrated a positive clinical outcome from BEC-derived hepatocytes in resolving human cirrhosis,[33] indicating that aberrant glutamine synthetase positivity in portal hepatocytes is significantly associated with regressed cirrhosis in humans. Moreover, in cases of severe intoxication with drugs such as acetaminophen, BEC-to-hepatocyte conversion is observed and associated with decreased DRs in patients.[139] However, this study concluded that the expansion and differentiation of BECs into hepatocyte-like cells take longer than required to prevent urgent liver transplantation. Hence, to become a viable and effective treatment, particularly for acute liver injuries, the cell conversion must be accelerated. Stimulating the cell plasticity is an attractive therapeutic option for patients with advanced liver disease. The ability to reliably identify a true progenitor population and to, thus, define druggable pathways that would accelerate their plasticity and differentiation into functional hepatocytes would immensely facilitate the therapeutic potential of BECs within the naturally occurring DRs in the vast majority of human liver diseases.


#
#
#

Conflict of interest

None declared.

Acknowledgments

We apologize for not being able to discuss and cite all work in this growing field due to space restraints. Illustrations were created using BioRender (biorender.com).

* These authors contributed equally to the research and retain the first authorship.


  • References

  • 1 Miyajima A, Tanaka M, Itoh T. Stem/progenitor cells in liver development, homeostasis, regeneration, and reprogramming. Cell Stem Cell 2014; 14 (05) 561-574
  • 2 Dorrell C, Erker L, Lanxon-Cookson KM. et al. Surface markers for the murine oval cell response. Hepatology 2008; 48 (04) 1282-1291
  • 3 Jensen CH, Jauho EI, Santoni-Rugiu E. et al. Transit-amplifying ductular (oval) cells and their hepatocytic progeny are characterized by a novel and distinctive expression of delta-like protein/preadipocyte factor 1/fetal antigen 1. Am J Pathol 2004; 164 (04) 1347-1359
  • 4 Kamiya A, Kakinuma S, Yamazaki Y, Nakauchi H. Enrichment and clonal culture of progenitor cells during mouse postnatal liver development in mice. Gastroenterology 2009; 137 (03) 1114-1126 , 1126.e1–1126.e14
  • 5 Okabe M, Tsukahara Y, Tanaka M. et al. Potential hepatic stem cells reside in EpCAM+ cells of normal and injured mouse liver. Development 2009; 136 (11) 1951-1960
  • 6 Qiu Q, Hernandez JC, Dean AM, Rao PH, Darlington GJ. CD24-positive cells from normal adult mouse liver are hepatocyte progenitor cells. Stem Cells Dev 2011; 20 (12) 2177-2188
  • 7 Rountree CB, Barsky L, Ge S, Zhu J, Senadheera S, Crooks GM. A CD133-expressing murine liver oval cell population with bilineage potential. Stem Cells 2007; 25 (10) 2419-2429
  • 8 Schmelzer E, Zhang L, Bruce A. et al. Human hepatic stem cells from fetal and postnatal donors. J Exp Med 2007; 204 (08) 1973-1987
  • 9 Suzuki A, Sekiya S, Onishi M. et al. Flow cytometric isolation and clonal identification of self-renewing bipotent hepatic progenitor cells in adult mouse liver. Hepatology 2008; 48 (06) 1964-1978
  • 10 Tanaka M, Okabe M, Suzuki K. et al. Mouse hepatoblasts at distinct developmental stages are characterized by expression of EpCAM and DLK1: drastic change of EpCAM expression during liver development. Mech Dev 2009; 126 (8-9): 665-676
  • 11 Yovchev MI, Grozdanov PN, Zhou H, Racherla H, Guha C, Dabeva MD. Identification of adult hepatic progenitor cells capable of repopulating injured rat liver. Hepatology 2008; 47 (02) 636-647
  • 12 Overi D, Carpino G, Cardinale V. et al. Contribution of resident stem cells to liver and biliary tree regeneration in human diseases. Int J Mol Sci 2018; 19 (10) 19
  • 13 Shin S, Kaestner KH. The origin, biology, and therapeutic potential of facultative adult hepatic progenitor cells. Curr Top Dev Biol 2014; 107: 269-292
  • 14 Sato K, Marzioni M, Meng F, Francis H, Glaser S, Alpini G. Ductular reaction in liver diseases: pathological mechanisms and translational significances. Hepatology 2019; 69 (01) 420-430
  • 15 Boulter L, Lu WY, Forbes SJ. Differentiation of progenitors in the liver: a matter of local choice. J Clin Invest 2013; 123 (05) 1867-1873
  • 16 Gouw AS, Clouston AD, Theise ND. Ductular reactions in human liver: diversity at the interface. Hepatology 2011; 54 (05) 1853-1863
  • 17 Roskams TA, Theise ND, Balabaud C. et al. Nomenclature of the finer branches of the biliary tree: canals, ductules, and ductular reactions in human livers. Hepatology 2004; 39 (06) 1739-1745
  • 18 Roskams T, Desmet V. Ductular reaction and its diagnostic significance. Semin Diagn Pathol 1998; 15 (04) 259-269
  • 19 Popper H, Kent G, Stein R. Ductular cell reaction in the liver in hepatic injury. J Mt Sinai Hosp N Y 1957; 24 (05) 551-556
  • 20 Turányi E, Dezsö K, Csomor J, Schaff Z, Paku S, Nagy P. Immunohistochemical classification of ductular reactions in human liver. Histopathology 2010; 57 (04) 607-614
  • 21 Lowes KN, Brennan BA, Yeoh GC, Olynyk JK. Oval cell numbers in human chronic liver diseases are directly related to disease severity. Am J Pathol 1999; 154 (02) 537-541
  • 22 Van Haele M, Snoeck J, Roskams T. Human liver regeneration: an etiology dependent process. Int J Mol Sci 2019; 20 (09) 20
  • 23 Roskams TA, Libbrecht L, Desmet VJ. Progenitor cells in diseased human liver. Semin Liver Dis 2003; 23 (04) 385-396
  • 24 Yoon SM, Gerasimidou D, Kuwahara R. et al. Epithelial cell adhesion molecule (EpCAM) marks hepatocytes newly derived from stem/progenitor cells in humans. Hepatology 2011; 53 (03) 964-973
  • 25 Stueck AE, Wanless IR. Hepatocyte buds derived from progenitor cells repopulate regions of parenchymal extinction in human cirrhosis. Hepatology 2015; 61 (05) 1696-1707
  • 26 Hytiroglou P, Theise ND. Regression of human cirrhosis: an update, 18 years after the pioneering article by Wanless et al. Virchows Arch 2018; 473 (01) 15-22
  • 27 Wanless IR, Nakashima E, Sherman M. Regression of human cirrhosis. Morphologic features and the genesis of incomplete septal cirrhosis. Arch Pathol Lab Med 2000; 124 (11) 1599-1607
  • 28 Falkowski O, An HJ, Ianus IA. et al. Regeneration of hepatocyte ‘buds’ in cirrhosis from intrabiliary stem cells. J Hepatol 2003; 39 (03) 357-364
  • 29 Fleming KE, Wanless IR. Glutamine synthetase expression in activated hepatocyte progenitor cells and loss of hepatocellular expression in congestion and cirrhosis. Liver Int 2013; 33 (04) 525-534
  • 30 Deng X, Zhang X, Li W. et al. Chronic liver injury induces conversion of biliary epithelial cells into hepatocytes. Cell Stem Cell 2018; 23 (01) 114-122 .e3
  • 31 Haque S, Haruna Y, Saito K. et al. Identification of bipotential progenitor cells in human liver regeneration. Lab Invest 1996; 75 (05) 699-705
  • 32 Sancho-Bru P, Altamirano J, Rodrigo-Torres D. et al. Liver progenitor cell markers correlate with liver damage and predict short-term mortality in patients with alcoholic hepatitis. Hepatology 2012; 55 (06) 1931-1941
  • 33 Hadi R, Shin K, Reder N. et al. Utility of glutamine synthetase immunohistochemistry in identifying features of regressed cirrhosis. Mod Pathol 2020; 33 (03) 448-455
  • 34 Lin WR, Lim SN, McDonald SA. et al. The histogenesis of regenerative nodules in human liver cirrhosis. Hepatology 2010; 51 (03) 1017-1026
  • 35 Fellous TG, Islam S, Tadrous PJ. et al. Locating the stem cell niche and tracing hepatocyte lineages in human liver. Hepatology 2009; 49 (05) 1655-1663
  • 36 Van Eyken P, Sciot R, Desmet VJ. A cytokeratin immunohistochemical study of cholestatic liver disease: evidence that hepatocytes can express ‘bile duct-type’ cytokeratins. Histopathology 1989; 15 (02) 125-135
  • 37 Bellizzi AM, LeGallo RD, Boyd JC, Iezzoni JC. Hepatocyte cytokeratin 7 expression in chronic allograft rejection. Am J Clin Pathol 2011; 135 (02) 238-244
  • 38 Ernst LM, Spinner NB, Piccoli DA, Mauger J, Russo P. Interlobular bile duct loss in pediatric cholestatic disease is associated with aberrant cytokeratin 7 expression by hepatocytes. Pediatr Dev Pathol 2007; 10 (05) 383-390
  • 39 Goldstein NS, Soman A, Gordon SC. Portal tract eosinophils and hepatocyte cytokeratin 7 immunoreactivity helps distinguish early-stage, mildly active primary biliary cirrhosis and autoimmune hepatitis. Am J Clin Pathol 2001; 116 (06) 846-853
  • 40 Quaglia A, Bhathal PS. Copper, copper-binding protein and cytokeratin 7 in biliary disorders. Histopathology 2017; 71 (06) 1006-1008
  • 41 Sakellariou S, Michaelides C, Voulgaris T. et al. Keratin 7 expression in hepatic cholestatic diseases. Virchows Arch 2021; 479 (04) 815-824
  • 42 Sjöblom N, Boyd S, Manninen A. et al. Chronic cholestasis detection by a novel tool: automated analysis of cytokeratin 7-stained liver specimens. Diagn Pathol 2021; 16 (01) 41
  • 43 Fabris L, Cadamuro M, Guido M. et al. Analysis of liver repair mechanisms in Alagille syndrome and biliary atresia reveals a role for notch signaling. Am J Pathol 2007; 171 (02) 641-653
  • 44 Carpino G, Cardinale V, Folseraas T. et al. Hepatic stem/progenitor cell activation differs between primary sclerosing and primary biliary cholangitis. Am J Pathol 2018; 188 (03) 627-639
  • 45 Chatzipantelis P, Lazaris AC, Kafiri G. et al. Cytokeratin-7, cytokeratin-19, and c-Kit: immunoreaction during the evolution stages of primary biliary cirrhosis. Hepatol Res 2006; 36 (03) 182-187
  • 46 Limaye PB, Alarcón G, Walls AL. et al. Expression of specific hepatocyte and cholangiocyte transcription factors in human liver disease and embryonic development. Lab Invest 2008; 88 (08) 865-872
  • 47 Cazzagon N, Sarcognato S, Floreani A. et al. Cholangiocyte senescence in primary sclerosing cholangitis is associated with disease severity and prognosis. JHEP Rep 2021; 3 (03) 100286
  • 48 Harada K, Furubo S, Ozaki S, Hiramatsu K, Sudo Y, Nakanuma Y. Increased expression of WAF1 in intrahepatic bile ducts in primary biliary cirrhosis relates to apoptosis. J Hepatol 2001; 34 (04) 500-506
  • 49 Meadows V, Baiocchi L, Kundu D. et al. Biliary epithelial senescence in liver disease: there will be SASP. Front Mol Biosci 2021; 8: 803098
  • 50 Sasaki M, Ikeda H, Yamaguchi J, Nakada S, Nakanuma Y. Telomere shortening in the damaged small bile ducts in primary biliary cirrhosis reflects ongoing cellular senescence. Hepatology 2008; 48 (01) 186-195
  • 51 Tabibian JH, O'Hara SP, Splinter PL, Trussoni CE, LaRusso NF. Cholangiocyte senescence by way of N-ras activation is a characteristic of primary sclerosing cholangitis. Hepatology 2014; 59 (06) 2263-2275
  • 52 Tabibian JH, Trussoni CE, O'Hara SP, Splinter PL, Heimbach JK, LaRusso NF. Characterization of cultured cholangiocytes isolated from livers of patients with primary sclerosing cholangitis. Lab Invest 2014; 94 (10) 1126-1133
  • 53 Crosby HA, Hubscher SG, Joplin RE, Kelly DA, Strain AJ. Immunolocalization of OV-6, a putative progenitor cell marker in human fetal and diseased pediatric liver. Hepatology 1998; 28 (04) 980-985
  • 54 Crosby HA, Hubscher S, Fabris L. et al. Immunolocalization of putative human liver progenitor cells in livers from patients with end-stage primary biliary cirrhosis and sclerosing cholangitis using the monoclonal antibody OV-6. Am J Pathol 1998; 152 (03) 771-779
  • 55 Tarlow BD, Pelz C, Naugler WE. et al. Bipotential adult liver progenitors are derived from chronically injured mature hepatocytes. Cell Stem Cell 2014; 15 (05) 605-618
  • 56 Alison M, Golding M, Lalani EN, Nagy P, Thorgeirsson S, Sarraf C. Wholesale hepatocytic differentiation in the rat from ductular oval cells, the progeny of biliary stem cells. J Hepatol 1997; 26 (02) 343-352
  • 57 Evarts RP, Nagy P, Nakatsukasa H, Marsden E, Thorgeirsson SS. In vivo differentiation of rat liver oval cells into hepatocytes. Cancer Res 1989; 49 (06) 1541-1547
  • 58 Li Q, Gong Y, Wang Y. et al. Sirt1 promotes the restoration of hepatic progenitor cell (HPC)-mediated liver fatty injury in NAFLD through activating the Wnt/β-catenin signal pathway. Front Nutr 2021; 8: 791861
  • 59 Manco R, Clerbaux LA, Verhulst S. et al. Reactive cholangiocytes differentiate into proliferative hepatocytes with efficient DNA repair in mice with chronic liver injury. J Hepatol 2019; 70 (06) 1180-1191
  • 60 Han X, Wang Y, Pu W. et al. Lineage tracing reveals the bipotency of SOX9+ hepatocytes during liver regeneration. Stem Cell Reports 2019; 12 (03) 624-638
  • 61 Shin S, Upadhyay N, Greenbaum LE, Kaestner KH. Ablation of Foxl1-Cre-labeled hepatic progenitor cells and their descendants impairs recovery of mice from liver injury. Gastroenterology 2015; 148 (01) 192-202 .e3
  • 62 Español-Suñer R, Carpentier R, Van Hul N. et al. Liver progenitor cells yield functional hepatocytes in response to chronic liver injury in mice. Gastroenterology 2012; 143 (06) 1564-1575 .e7
  • 63 Shin S, Walton G, Aoki R. et al. Foxl1-Cre-marked adult hepatic progenitors have clonogenic and bilineage differentiation potential. Genes Dev 2011; 25 (11) 1185-1192
  • 64 Malato Y, Naqvi S, Schürmann N. et al. Fate tracing of mature hepatocytes in mouse liver homeostasis and regeneration. J Clin Invest 2011; 121 (12) 4850-4860
  • 65 Dorrell C, Erker L, Schug J. et al. Prospective isolation of a bipotential clonogenic liver progenitor cell in adult mice. Genes Dev 2011; 25 (11) 1193-1203
  • 66 Sackett SD, Li Z, Hurtt R. et al. Foxl1 is a marker of bipotential hepatic progenitor cells in mice. Hepatology 2009; 49 (03) 920-929
  • 67 Lu WY, Bird TG, Boulter L. et al. Hepatic progenitor cells of biliary origin with liver repopulation capacity. Nat Cell Biol 2015; 17 (08) 971-983
  • 68 Raven A, Lu WY, Man TY. et al. Cholangiocytes act as facultative liver stem cells during impaired hepatocyte regeneration. Nature 2017; 547 (7663): 350-354
  • 69 Russell JO, Lu WY, Okabe H. et al. Hepatocyte-specific β-catenin deletion during severe liver injury provokes cholangiocytes to differentiate into hepatocytes. Hepatology 2019; 69 (02) 742-759
  • 70 Wood MJ, Gadd VL, Powell LW, Ramm GA, Clouston AD. Ductular reaction in hereditary hemochromatosis: the link between hepatocyte senescence and fibrosis progression. Hepatology 2014; 59 (03) 848-857
  • 71 Gadd VL, Skoien R, Powell EE. et al. The portal inflammatory infiltrate and ductular reaction in human nonalcoholic fatty liver disease. Hepatology 2014; 59 (04) 1393-1405
  • 72 Weng HL, Feng DC, Radaeva S. et al. IFN-γ inhibits liver progenitor cell proliferation in HBV-infected patients and in 3,5-diethoxycarbonyl-1,4-dihydrocollidine diet-fed mice. J Hepatol 2013; 59 (04) 738-745
  • 73 Richardson MM, Jonsson JR, Powell EE. et al. Progressive fibrosis in nonalcoholic steatohepatitis: association with altered regeneration and a ductular reaction. Gastroenterology 2007; 133 (01) 80-90
  • 74 Clouston AD, Powell EE, Walsh MJ, Richardson MM, Demetris AJ, Jonsson JR. Fibrosis correlates with a ductular reaction in hepatitis C: roles of impaired replication, progenitor cells and steatosis. Hepatology 2005; 41 (04) 809-818
  • 75 Roskams T, Yang SQ, Koteish A. et al. Oxidative stress and oval cell accumulation in mice and humans with alcoholic and nonalcoholic fatty liver disease. Am J Pathol 2003; 163 (04) 1301-1311
  • 76 Minnis-Lyons SE, Ferreira-González S, Aleksieva N. et al. Notch-IGF1 signaling during liver regeneration drives biliary epithelial cell expansion and inhibits hepatocyte differentiation. Sci Signal 2021; 14 (688) 14
  • 77 Aloia L, McKie MA, Vernaz G. et al. Epigenetic remodelling licences adult cholangiocytes for organoid formation and liver regeneration. Nat Cell Biol 2019; 21 (11) 1321-1333
  • 78 Rizvi F, Everton E, Smith AR. et al. Murine liver repair via transient activation of regenerative pathways in hepatocytes using lipid nanoparticle-complexed nucleoside-modified mRNA. Nat Commun 2021; 12 (01) 613
  • 79 Ota N, Shiojiri N. Comparative study on a novel lobule structure of the zebrafish liver and that of the mammalian liver. Cell Tissue Res 2022; 388 (02) 287-299
  • 80 Caviglia S, Unterweger IA, Gasiūnaitė A. et al. FRaeppli: a multispectral imaging toolbox for cell tracing and dense tissue analysis in zebrafish. Development 2022; 149 (16) 149
  • 81 Zhao C, Lancman JJ, Yang Y. et al. Intrahepatic cholangiocyte regeneration from an Fgf-dependent extrahepatic progenitor niche in a zebrafish model of Alagille Syndrome. Hepatology 2022; 75 (03) 567-583
  • 82 Wang J, Leng X, Wang G, Wan X, Cao H. The construction of intrahepatic cholangiocarcinoma model in zebrafish. Sci Rep 2017; 7 (01) 13419
  • 83 Goessling W, Sadler KC. Zebrafish: an important tool for liver disease research. Gastroenterology 2015; 149 (06) 1361-1377
  • 84 Evason KJ, Francisco MT, Juric V. et al. Identification of chemical inhibitors of β-catenin-driven liver tumorigenesis in zebrafish. PLoS Genet 2015; 11 (07) e1005305
  • 85 Den Broeder MJ, Kopylova VA, Kamminga LM, Legler J. Zebrafish as a model to study the role of peroxisome proliferating-activated receptors in adipogenesis and obesity. PPAR Res 2015; 2015: 358029
  • 86 Nguyen AT, Emelyanov A, Koh CH. et al. A high level of liver-specific expression of oncogenic Kras(V12) drives robust liver tumorigenesis in transgenic zebrafish. Dis Model Mech 2011; 4 (06) 801-813
  • 87 Vogt A, Cholewinski A, Shen X. et al. Automated image-based phenotypic analysis in zebrafish embryos. Dev Dyn 2009; 238 (03) 656-663
  • 88 Liu W, Chen JR, Hsu CH. et al. A zebrafish model of intrahepatic cholangiocarcinoma by dual expression of hepatitis B virus X and hepatitis C virus core protein in liver. Hepatology 2012; 56 (06) 2268-2276
  • 89 Shwartz A, Goessling W, Yin C. Macrophages in zebrafish models of liver diseases. Front Immunol 2019; 10: 2840
  • 90 Pham DH, Yin C. Zebrafish as a model to study cholestatic liver diseases. Methods Mol Biol 2019; 1981: 273-289
  • 91 Huang M, Chang A, Choi M, Zhou D, Anania FA, Shin CH. Antagonistic interaction between Wnt and Notch activity modulates the regenerative capacity of a zebrafish fibrotic liver model. Hepatology 2014; 60 (05) 1753-1766
  • 92 Ko S, Shin D. Chemical screening using a zebrafish model for liver progenitor cell-driven liver regeneration. Methods Mol Biol 2019; 1905: 83-90
  • 93 Cox AG, Saunders DC, Kelsey Jr PB. et al. S-nitrosothiol signaling regulates liver development and improves outcome following toxic liver injury. Cell Rep 2014; 6 (01) 56-69
  • 94 He JH, Guo SY, Zhu F. et al. A zebrafish phenotypic assay for assessing drug-induced hepatotoxicity. J Pharmacol Toxicol Methods 2013; 67 (01) 25-32
  • 95 North TE, Babu IR, Vedder LM. et al. PGE2-regulated Wnt signaling and N-acetylcysteine are synergistically hepatoprotective in zebrafish acetaminophen injury. Proc Natl Acad Sci U S A 2010; 107 (40) 17315-17320
  • 96 He J, Lu H, Zou Q, Luo L. Regeneration of liver after extreme hepatocyte loss occurs mainly via biliary transdifferentiation in zebrafish. Gastroenterology 2014; 146 (03) 789-800 .e8
  • 97 Choi TY, Ninov N, Stainier DY, Shin D. Extensive conversion of hepatic biliary epithelial cells to hepatocytes after near total loss of hepatocytes in zebrafish. Gastroenterology 2014; 146 (03) 776-788
  • 98 Ko S, Choi TY, Russell JO, So J, Monga SPS, Shin D. Bromodomain and extraterminal (BET) proteins regulate biliary-driven liver regeneration. J Hepatol 2016; 64 (02) 316-325
  • 99 He J, Zhou Y, Qian C. et al. DNA methylation maintenance at the p53 locus initiates biliary-mediated liver regeneration. NPJ Regen Med 2022; 7 (01) 21
  • 100 He J, Chen J, Wei X. et al. Mammalian target of rapamycin complex 1 signaling is required for the dedifferentiation from biliary cell to bipotential progenitor cell in zebrafish liver regeneration. Hepatology 2019; 70 (06) 2092-2106
  • 101 Zhang J, Zhou Y, Li S. et al. Tel2 regulates redifferentiation of bipotential progenitor cells via Hhex during zebrafish liver regeneration. Cell Rep 2022; 39 (01) 110596
  • 102 Khaliq M, Ko S, Liu Y. et al. Stat3 regulates liver progenitor cell-driven liver regeneration in zebrafish. Gene Expr 2018; 18 (03) 157-170
  • 103 Jung K, Kim M, So J, Lee SH, Ko S, Shin D. Farnesoid X receptor activation impairs liver progenitor cell-mediated liver regeneration via the PTEN-PI3K-AKT-mTOR axis in zebrafish. Hepatology 2021; 74 (01) 397-410
  • 104 Miyoshi T, Hidaka M, Miyamoto D. et al. Successful induction of human chemically induced liver progenitors with small molecules from damaged liver. J Gastroenterol 2022; 57 (06) 441-452
  • 105 Choi TY, Khaliq M, Tsurusaki S. et al. Bone morphogenetic protein signaling governs biliary-driven liver regeneration in zebrafish through tbx2b and id2a. Hepatology 2017; 66 (05) 1616-1630
  • 106 Ko S, Russell JO, Tian J. et al. Hdac1 regulates differentiation of bipotent liver progenitor cells during regeneration via Sox9b and Cdk8. Gastroenterology 2019; 156 (01) 187-202 .e14
  • 107 Cai P, Mao X, Zhao J. et al. Farnesoid X receptor is required for the redifferentiation of bipotential progenitor cells during biliary-mediated zebrafish liver regeneration. Hepatology 2021; 74 (06) 3345-3361
  • 108 So J, Kim M, Lee SH. et al. Attenuating the epidermal growth factor receptor-extracellular signal-regulated kinase-sex-determining region Y-box 9 axis promotes liver progenitor cell-mediated liver regeneration in zebrafish. Hepatology 2021; 73 (04) 1494-1508
  • 109 Michalopoulos GK, Barua L, Bowen WC. Transdifferentiation of rat hepatocytes into biliary cells after bile duct ligation and toxic biliary injury. Hepatology 2005; 41 (03) 535-544
  • 110 Limaye PB, Bowen WC, Orr AV, Luo J, Tseng GC, Michalopoulos GK. Mechanisms of hepatocyte growth factor-mediated and epidermal growth factor-mediated signaling in transdifferentiation of rat hepatocytes to biliary epithelium. Hepatology 2008; 47 (05) 1702-1713
  • 111 Limaye PB, Bowen WC, Orr A, Apte UM, Michalopoulos GK. Expression of hepatocytic- and biliary-specific transcription factors in regenerating bile ducts during hepatocyte-to-biliary epithelial cell transdifferentiation. Comp Hepatol 2010; 9: 9
  • 112 Nagahama Y, Sone M, Chen X. et al. Contributions of hepatocytes and bile ductular cells in ductular reactions and remodeling of the biliary system after chronic liver injury. Am J Pathol 2014; 184 (11) 3001-3012
  • 113 Yanger K, Zong Y, Maggs LR. et al. Robust cellular reprogramming occurs spontaneously during liver regeneration. Genes Dev 2013; 27 (07) 719-724
  • 114 Sekiya S, Suzuki A. Hepatocytes, rather than cholangiocytes, can be the major source of primitive ductules in the chronically injured mouse liver. Am J Pathol 2014; 184 (05) 1468-1478
  • 115 Lorent K, Moore JC, Siekmann AF, Lawson N, Pack M. Reiterative use of the notch signal during zebrafish intrahepatic biliary development. Dev Dyn 2010; 239 (03) 855-864
  • 116 Zong Y, Panikkar A, Xu J. et al. Notch signaling controls liver development by regulating biliary differentiation. Development 2009; 136 (10) 1727-1739
  • 117 Lemaigre FP. Development of the intrahepatic and extrahepatic biliary tract: a framework for understanding congenital diseases. Annu Rev Pathol 2020; 15: 1-22
  • 118 Hu S, Molina L, Tao J. et al. NOTCH-YAP1/TEAD-DNMT1 axis drives hepatocyte reprogramming into intrahepatic cholangiocarcinoma. Gastroenterology 2022; 163 (02) 449-465
  • 119 Fan B, Malato Y, Calvisi DF. et al. Cholangiocarcinomas can originate from hepatocytes in mice. J Clin Invest 2012; 122 (08) 2911-2915
  • 120 Molina L, Nejak-Bowen K, Monga SP. Role of YAP1 signaling in biliary development, repair, and disease. Semin Liver Dis 2022; 42 (01) 17-33
  • 121 Lee DH, Park JO, Kim TS. et al. LATS-YAP/TAZ controls lineage specification by regulating TGFβ signaling and Hnf4α expression during liver development. Nat Commun 2016; 7: 11961
  • 122 Planas-Paz L, Sun T, Pikiolek M. et al. YAP, but not RSPO-LGR4/5, signaling in biliary epithelial cells promotes a ductular reaction in response to liver injury. Cell Stem Cell 2019; 25 (01) 39-53 .e10
  • 123 Li W, Yang L, He Q. et al. A homeostatic Arid1a-dependent permissive chromatin state licenses hepatocyte responsiveness to liver-injury-associated YAP signaling. Cell Stem Cell 2019; 25 (01) 54-68 .e5
  • 124 Liu Y, Zhuo S, Zhou Y. et al. Yap-Sox9 signaling determines hepatocyte plasticity and lineage-specific hepatocarcinogenesis. J Hepatol 2022; 76 (03) 652-664
  • 125 Wang J, Dong M, Xu Z. et al. Notch2 controls hepatocyte-derived cholangiocarcinoma formation in mice. Oncogene 2018; 37 (24) 3229-3242
  • 126 Yimlamai D, Christodoulou C, Galli GG. et al. Hippo pathway activity influences liver cell fate. Cell 2014; 157 (06) 1324-1338
  • 127 Schaub JR, Huppert KA, Kurial SNT. et al. De novo formation of the biliary system by TGFβ-mediated hepatocyte transdifferentiation. Nature 2018; 557 (7704): 247-251
  • 128 Tschaharganeh DF, Xue W, Calvisi DF. et al. p53-dependent Nestin regulation links tumor suppression to cellular plasticity in liver cancer. Cell 2014; 158 (03) 579-592
  • 129 Jeliazkova P, Jörs S, Lee M. et al. Canonical Notch2 signaling determines biliary cell fates of embryonic hepatoblasts and adult hepatocytes independent of Hes1. Hepatology 2013; 57 (06) 2469-2479
  • 130 Guo J, Fu W, Xiang M. et al. Notch1 drives the formation and proliferation of intrahepatic cholangiocarcinoma. Curr Med Sci 2019; 39 (06) 929-937
  • 131 Pham DH, Kudira R, Xu L. et al. Deleterious variants in ABCC12 are detected in idiopathic chronic cholestasis and cause intrahepatic bile duct loss in model organisms. Gastroenterology 2021; 161 (01) 287-300 .e16
  • 132 Lee SH, So J, Shin D. Hepatocyte-to-cholangiocyte conversion occurs through transdifferentiation independently of proliferation in zebrafish. Hepatology 2023, in press
  • 133 Russell JO, Ko S, Monga SP, Shin D. Notch inhibition promotes differentiation of liver progenitor cells into hepatocytes via sox9b repression in zebrafish. Stem Cells Int 2019; 2019: 8451282
  • 134 Everton E, Rizvi F, Smith AR. et al. Transient yet robust expression of proteins in the mouse liver via intravenous injection of lipid nanoparticle-encapsulated nucleoside-modified mRNA. Bio Protoc 2021; 11 (19) e4184
  • 135 Pepe-Mooney BJ, Dill MT, Alemany A. et al. Single-cell analysis of the liver epithelium reveals dynamic heterogeneity and an essential role for YAP in homeostasis and regeneration. Cell Stem Cell 2019; 25 (01) 23-38 .e8
  • 136 Bou Saleh M, Louvet A, Ntandja-Wandji LC. et al. Loss of hepatocyte identity following aberrant YAP activation: a key mechanism in alcoholic hepatitis. J Hepatol 2021; 75 (04) 912-923
  • 137 Glaser SS, Gaudio E, Miller T, Alvaro D, Alpini G. Cholangiocyte proliferation and liver fibrosis. Expert Rev Mol Med 2009; 11: e7
  • 138 Kaur S, Siddiqui H, Bhat MH. Hepatic progenitor cells in action: liver regeneration or fibrosis?. Am J Pathol 2015; 185 (09) 2342-2350
  • 139 Katoonizadeh A, Nevens F, Verslype C, Pirenne J, Roskams T. Liver regeneration in acute severe liver impairment: a clinicopathological correlation study. Liver Int 2006; 26 (10) 1225-1233
  • 140 Rodrigo-Torres D, Affò S, Coll M. et al. The biliary epithelium gives rise to liver progenitor cells. Hepatology 2014; 60 (04) 1367-1377
  • 141 Furuyama K, Kawaguchi Y, Akiyama H. et al. Continuous cell supply from a Sox9-expressing progenitor zone in adult liver, exocrine pancreas and intestine. Nat Genet 2011; 43 (01) 34-41
  • 142 Font-Burgada J, Shalapour S, Ramaswamy S. et al. Hybrid periportal hepatocytes regenerate the injured liver without giving rise to cancer. Cell 2015; 162 (04) 766-779
  • 143 Lesaffer B, Verboven E, Van Huffel L. et al. Comparison of the Opn-CreER and Ck19-CreER drivers in bile ducts of normal and injured mouse livers. Cells 2019; 8 (04) 8
  • 144 Morales-Ibanez O, Domínguez M, Ki SH. et al. Human and experimental evidence supporting a role for osteopontin in alcoholic hepatitis. Hepatology 2013; 58 (05) 1742-1756
  • 145 Uede T. Osteopontin, intrinsic tissue regulator of intractable inflammatory diseases. Pathol Int 2011; 61 (05) 265-280
  • 146 Zweerink S, Mueck V, Kraemer LP. et al. Repolarization precedes oval cell-mediated hepatocytic regeneration in the CDE diet mouse model. J Histochem Cytochem 2022; 70 (05) 377-389
  • 147 Sekiya S, Suzuki A. Intrahepatic cholangiocarcinoma can arise from Notch-mediated conversion of hepatocytes. J Clin Invest 2012; 122 (11) 3914-3918
  • 148 Yoshii D, Shimata K, Yokouchi Y. et al. SOX9 contributes to the progression of ductular reaction for the protection from chronic liver injury. Hum Cell 2022; 35 (02) 721-734

Address for correspondence

Donghun Shin, PhD
Department of Developmental Biology
University of Pittsburgh School of Medicine, 3501 5th Avenue #5063 Pittsburgh, PA 15260

Publication History

Article published online:
10 February 2023

© 2023. The Author(s). This is an open access article published by Thieme under the terms of the Creative Commons Attribution-NonDerivative-NonCommercial License, permitting copying and reproduction so long as the original work is given appropriate credit. Contents may not be used for commercial purposes, or adapted, remixed, transformed or built upon. (https://creativecommons.org/licenses/by-nc-nd/4.0/)

Thieme Medical Publishers, Inc.
333 Seventh Avenue, 18th Floor, New York, NY 10001, USA

  • References

  • 1 Miyajima A, Tanaka M, Itoh T. Stem/progenitor cells in liver development, homeostasis, regeneration, and reprogramming. Cell Stem Cell 2014; 14 (05) 561-574
  • 2 Dorrell C, Erker L, Lanxon-Cookson KM. et al. Surface markers for the murine oval cell response. Hepatology 2008; 48 (04) 1282-1291
  • 3 Jensen CH, Jauho EI, Santoni-Rugiu E. et al. Transit-amplifying ductular (oval) cells and their hepatocytic progeny are characterized by a novel and distinctive expression of delta-like protein/preadipocyte factor 1/fetal antigen 1. Am J Pathol 2004; 164 (04) 1347-1359
  • 4 Kamiya A, Kakinuma S, Yamazaki Y, Nakauchi H. Enrichment and clonal culture of progenitor cells during mouse postnatal liver development in mice. Gastroenterology 2009; 137 (03) 1114-1126 , 1126.e1–1126.e14
  • 5 Okabe M, Tsukahara Y, Tanaka M. et al. Potential hepatic stem cells reside in EpCAM+ cells of normal and injured mouse liver. Development 2009; 136 (11) 1951-1960
  • 6 Qiu Q, Hernandez JC, Dean AM, Rao PH, Darlington GJ. CD24-positive cells from normal adult mouse liver are hepatocyte progenitor cells. Stem Cells Dev 2011; 20 (12) 2177-2188
  • 7 Rountree CB, Barsky L, Ge S, Zhu J, Senadheera S, Crooks GM. A CD133-expressing murine liver oval cell population with bilineage potential. Stem Cells 2007; 25 (10) 2419-2429
  • 8 Schmelzer E, Zhang L, Bruce A. et al. Human hepatic stem cells from fetal and postnatal donors. J Exp Med 2007; 204 (08) 1973-1987
  • 9 Suzuki A, Sekiya S, Onishi M. et al. Flow cytometric isolation and clonal identification of self-renewing bipotent hepatic progenitor cells in adult mouse liver. Hepatology 2008; 48 (06) 1964-1978
  • 10 Tanaka M, Okabe M, Suzuki K. et al. Mouse hepatoblasts at distinct developmental stages are characterized by expression of EpCAM and DLK1: drastic change of EpCAM expression during liver development. Mech Dev 2009; 126 (8-9): 665-676
  • 11 Yovchev MI, Grozdanov PN, Zhou H, Racherla H, Guha C, Dabeva MD. Identification of adult hepatic progenitor cells capable of repopulating injured rat liver. Hepatology 2008; 47 (02) 636-647
  • 12 Overi D, Carpino G, Cardinale V. et al. Contribution of resident stem cells to liver and biliary tree regeneration in human diseases. Int J Mol Sci 2018; 19 (10) 19
  • 13 Shin S, Kaestner KH. The origin, biology, and therapeutic potential of facultative adult hepatic progenitor cells. Curr Top Dev Biol 2014; 107: 269-292
  • 14 Sato K, Marzioni M, Meng F, Francis H, Glaser S, Alpini G. Ductular reaction in liver diseases: pathological mechanisms and translational significances. Hepatology 2019; 69 (01) 420-430
  • 15 Boulter L, Lu WY, Forbes SJ. Differentiation of progenitors in the liver: a matter of local choice. J Clin Invest 2013; 123 (05) 1867-1873
  • 16 Gouw AS, Clouston AD, Theise ND. Ductular reactions in human liver: diversity at the interface. Hepatology 2011; 54 (05) 1853-1863
  • 17 Roskams TA, Theise ND, Balabaud C. et al. Nomenclature of the finer branches of the biliary tree: canals, ductules, and ductular reactions in human livers. Hepatology 2004; 39 (06) 1739-1745
  • 18 Roskams T, Desmet V. Ductular reaction and its diagnostic significance. Semin Diagn Pathol 1998; 15 (04) 259-269
  • 19 Popper H, Kent G, Stein R. Ductular cell reaction in the liver in hepatic injury. J Mt Sinai Hosp N Y 1957; 24 (05) 551-556
  • 20 Turányi E, Dezsö K, Csomor J, Schaff Z, Paku S, Nagy P. Immunohistochemical classification of ductular reactions in human liver. Histopathology 2010; 57 (04) 607-614
  • 21 Lowes KN, Brennan BA, Yeoh GC, Olynyk JK. Oval cell numbers in human chronic liver diseases are directly related to disease severity. Am J Pathol 1999; 154 (02) 537-541
  • 22 Van Haele M, Snoeck J, Roskams T. Human liver regeneration: an etiology dependent process. Int J Mol Sci 2019; 20 (09) 20
  • 23 Roskams TA, Libbrecht L, Desmet VJ. Progenitor cells in diseased human liver. Semin Liver Dis 2003; 23 (04) 385-396
  • 24 Yoon SM, Gerasimidou D, Kuwahara R. et al. Epithelial cell adhesion molecule (EpCAM) marks hepatocytes newly derived from stem/progenitor cells in humans. Hepatology 2011; 53 (03) 964-973
  • 25 Stueck AE, Wanless IR. Hepatocyte buds derived from progenitor cells repopulate regions of parenchymal extinction in human cirrhosis. Hepatology 2015; 61 (05) 1696-1707
  • 26 Hytiroglou P, Theise ND. Regression of human cirrhosis: an update, 18 years after the pioneering article by Wanless et al. Virchows Arch 2018; 473 (01) 15-22
  • 27 Wanless IR, Nakashima E, Sherman M. Regression of human cirrhosis. Morphologic features and the genesis of incomplete septal cirrhosis. Arch Pathol Lab Med 2000; 124 (11) 1599-1607
  • 28 Falkowski O, An HJ, Ianus IA. et al. Regeneration of hepatocyte ‘buds’ in cirrhosis from intrabiliary stem cells. J Hepatol 2003; 39 (03) 357-364
  • 29 Fleming KE, Wanless IR. Glutamine synthetase expression in activated hepatocyte progenitor cells and loss of hepatocellular expression in congestion and cirrhosis. Liver Int 2013; 33 (04) 525-534
  • 30 Deng X, Zhang X, Li W. et al. Chronic liver injury induces conversion of biliary epithelial cells into hepatocytes. Cell Stem Cell 2018; 23 (01) 114-122 .e3
  • 31 Haque S, Haruna Y, Saito K. et al. Identification of bipotential progenitor cells in human liver regeneration. Lab Invest 1996; 75 (05) 699-705
  • 32 Sancho-Bru P, Altamirano J, Rodrigo-Torres D. et al. Liver progenitor cell markers correlate with liver damage and predict short-term mortality in patients with alcoholic hepatitis. Hepatology 2012; 55 (06) 1931-1941
  • 33 Hadi R, Shin K, Reder N. et al. Utility of glutamine synthetase immunohistochemistry in identifying features of regressed cirrhosis. Mod Pathol 2020; 33 (03) 448-455
  • 34 Lin WR, Lim SN, McDonald SA. et al. The histogenesis of regenerative nodules in human liver cirrhosis. Hepatology 2010; 51 (03) 1017-1026
  • 35 Fellous TG, Islam S, Tadrous PJ. et al. Locating the stem cell niche and tracing hepatocyte lineages in human liver. Hepatology 2009; 49 (05) 1655-1663
  • 36 Van Eyken P, Sciot R, Desmet VJ. A cytokeratin immunohistochemical study of cholestatic liver disease: evidence that hepatocytes can express ‘bile duct-type’ cytokeratins. Histopathology 1989; 15 (02) 125-135
  • 37 Bellizzi AM, LeGallo RD, Boyd JC, Iezzoni JC. Hepatocyte cytokeratin 7 expression in chronic allograft rejection. Am J Clin Pathol 2011; 135 (02) 238-244
  • 38 Ernst LM, Spinner NB, Piccoli DA, Mauger J, Russo P. Interlobular bile duct loss in pediatric cholestatic disease is associated with aberrant cytokeratin 7 expression by hepatocytes. Pediatr Dev Pathol 2007; 10 (05) 383-390
  • 39 Goldstein NS, Soman A, Gordon SC. Portal tract eosinophils and hepatocyte cytokeratin 7 immunoreactivity helps distinguish early-stage, mildly active primary biliary cirrhosis and autoimmune hepatitis. Am J Clin Pathol 2001; 116 (06) 846-853
  • 40 Quaglia A, Bhathal PS. Copper, copper-binding protein and cytokeratin 7 in biliary disorders. Histopathology 2017; 71 (06) 1006-1008
  • 41 Sakellariou S, Michaelides C, Voulgaris T. et al. Keratin 7 expression in hepatic cholestatic diseases. Virchows Arch 2021; 479 (04) 815-824
  • 42 Sjöblom N, Boyd S, Manninen A. et al. Chronic cholestasis detection by a novel tool: automated analysis of cytokeratin 7-stained liver specimens. Diagn Pathol 2021; 16 (01) 41
  • 43 Fabris L, Cadamuro M, Guido M. et al. Analysis of liver repair mechanisms in Alagille syndrome and biliary atresia reveals a role for notch signaling. Am J Pathol 2007; 171 (02) 641-653
  • 44 Carpino G, Cardinale V, Folseraas T. et al. Hepatic stem/progenitor cell activation differs between primary sclerosing and primary biliary cholangitis. Am J Pathol 2018; 188 (03) 627-639
  • 45 Chatzipantelis P, Lazaris AC, Kafiri G. et al. Cytokeratin-7, cytokeratin-19, and c-Kit: immunoreaction during the evolution stages of primary biliary cirrhosis. Hepatol Res 2006; 36 (03) 182-187
  • 46 Limaye PB, Alarcón G, Walls AL. et al. Expression of specific hepatocyte and cholangiocyte transcription factors in human liver disease and embryonic development. Lab Invest 2008; 88 (08) 865-872
  • 47 Cazzagon N, Sarcognato S, Floreani A. et al. Cholangiocyte senescence in primary sclerosing cholangitis is associated with disease severity and prognosis. JHEP Rep 2021; 3 (03) 100286
  • 48 Harada K, Furubo S, Ozaki S, Hiramatsu K, Sudo Y, Nakanuma Y. Increased expression of WAF1 in intrahepatic bile ducts in primary biliary cirrhosis relates to apoptosis. J Hepatol 2001; 34 (04) 500-506
  • 49 Meadows V, Baiocchi L, Kundu D. et al. Biliary epithelial senescence in liver disease: there will be SASP. Front Mol Biosci 2021; 8: 803098
  • 50 Sasaki M, Ikeda H, Yamaguchi J, Nakada S, Nakanuma Y. Telomere shortening in the damaged small bile ducts in primary biliary cirrhosis reflects ongoing cellular senescence. Hepatology 2008; 48 (01) 186-195
  • 51 Tabibian JH, O'Hara SP, Splinter PL, Trussoni CE, LaRusso NF. Cholangiocyte senescence by way of N-ras activation is a characteristic of primary sclerosing cholangitis. Hepatology 2014; 59 (06) 2263-2275
  • 52 Tabibian JH, Trussoni CE, O'Hara SP, Splinter PL, Heimbach JK, LaRusso NF. Characterization of cultured cholangiocytes isolated from livers of patients with primary sclerosing cholangitis. Lab Invest 2014; 94 (10) 1126-1133
  • 53 Crosby HA, Hubscher SG, Joplin RE, Kelly DA, Strain AJ. Immunolocalization of OV-6, a putative progenitor cell marker in human fetal and diseased pediatric liver. Hepatology 1998; 28 (04) 980-985
  • 54 Crosby HA, Hubscher S, Fabris L. et al. Immunolocalization of putative human liver progenitor cells in livers from patients with end-stage primary biliary cirrhosis and sclerosing cholangitis using the monoclonal antibody OV-6. Am J Pathol 1998; 152 (03) 771-779
  • 55 Tarlow BD, Pelz C, Naugler WE. et al. Bipotential adult liver progenitors are derived from chronically injured mature hepatocytes. Cell Stem Cell 2014; 15 (05) 605-618
  • 56 Alison M, Golding M, Lalani EN, Nagy P, Thorgeirsson S, Sarraf C. Wholesale hepatocytic differentiation in the rat from ductular oval cells, the progeny of biliary stem cells. J Hepatol 1997; 26 (02) 343-352
  • 57 Evarts RP, Nagy P, Nakatsukasa H, Marsden E, Thorgeirsson SS. In vivo differentiation of rat liver oval cells into hepatocytes. Cancer Res 1989; 49 (06) 1541-1547
  • 58 Li Q, Gong Y, Wang Y. et al. Sirt1 promotes the restoration of hepatic progenitor cell (HPC)-mediated liver fatty injury in NAFLD through activating the Wnt/β-catenin signal pathway. Front Nutr 2021; 8: 791861
  • 59 Manco R, Clerbaux LA, Verhulst S. et al. Reactive cholangiocytes differentiate into proliferative hepatocytes with efficient DNA repair in mice with chronic liver injury. J Hepatol 2019; 70 (06) 1180-1191
  • 60 Han X, Wang Y, Pu W. et al. Lineage tracing reveals the bipotency of SOX9+ hepatocytes during liver regeneration. Stem Cell Reports 2019; 12 (03) 624-638
  • 61 Shin S, Upadhyay N, Greenbaum LE, Kaestner KH. Ablation of Foxl1-Cre-labeled hepatic progenitor cells and their descendants impairs recovery of mice from liver injury. Gastroenterology 2015; 148 (01) 192-202 .e3
  • 62 Español-Suñer R, Carpentier R, Van Hul N. et al. Liver progenitor cells yield functional hepatocytes in response to chronic liver injury in mice. Gastroenterology 2012; 143 (06) 1564-1575 .e7
  • 63 Shin S, Walton G, Aoki R. et al. Foxl1-Cre-marked adult hepatic progenitors have clonogenic and bilineage differentiation potential. Genes Dev 2011; 25 (11) 1185-1192
  • 64 Malato Y, Naqvi S, Schürmann N. et al. Fate tracing of mature hepatocytes in mouse liver homeostasis and regeneration. J Clin Invest 2011; 121 (12) 4850-4860
  • 65 Dorrell C, Erker L, Schug J. et al. Prospective isolation of a bipotential clonogenic liver progenitor cell in adult mice. Genes Dev 2011; 25 (11) 1193-1203
  • 66 Sackett SD, Li Z, Hurtt R. et al. Foxl1 is a marker of bipotential hepatic progenitor cells in mice. Hepatology 2009; 49 (03) 920-929
  • 67 Lu WY, Bird TG, Boulter L. et al. Hepatic progenitor cells of biliary origin with liver repopulation capacity. Nat Cell Biol 2015; 17 (08) 971-983
  • 68 Raven A, Lu WY, Man TY. et al. Cholangiocytes act as facultative liver stem cells during impaired hepatocyte regeneration. Nature 2017; 547 (7663): 350-354
  • 69 Russell JO, Lu WY, Okabe H. et al. Hepatocyte-specific β-catenin deletion during severe liver injury provokes cholangiocytes to differentiate into hepatocytes. Hepatology 2019; 69 (02) 742-759
  • 70 Wood MJ, Gadd VL, Powell LW, Ramm GA, Clouston AD. Ductular reaction in hereditary hemochromatosis: the link between hepatocyte senescence and fibrosis progression. Hepatology 2014; 59 (03) 848-857
  • 71 Gadd VL, Skoien R, Powell EE. et al. The portal inflammatory infiltrate and ductular reaction in human nonalcoholic fatty liver disease. Hepatology 2014; 59 (04) 1393-1405
  • 72 Weng HL, Feng DC, Radaeva S. et al. IFN-γ inhibits liver progenitor cell proliferation in HBV-infected patients and in 3,5-diethoxycarbonyl-1,4-dihydrocollidine diet-fed mice. J Hepatol 2013; 59 (04) 738-745
  • 73 Richardson MM, Jonsson JR, Powell EE. et al. Progressive fibrosis in nonalcoholic steatohepatitis: association with altered regeneration and a ductular reaction. Gastroenterology 2007; 133 (01) 80-90
  • 74 Clouston AD, Powell EE, Walsh MJ, Richardson MM, Demetris AJ, Jonsson JR. Fibrosis correlates with a ductular reaction in hepatitis C: roles of impaired replication, progenitor cells and steatosis. Hepatology 2005; 41 (04) 809-818
  • 75 Roskams T, Yang SQ, Koteish A. et al. Oxidative stress and oval cell accumulation in mice and humans with alcoholic and nonalcoholic fatty liver disease. Am J Pathol 2003; 163 (04) 1301-1311
  • 76 Minnis-Lyons SE, Ferreira-González S, Aleksieva N. et al. Notch-IGF1 signaling during liver regeneration drives biliary epithelial cell expansion and inhibits hepatocyte differentiation. Sci Signal 2021; 14 (688) 14
  • 77 Aloia L, McKie MA, Vernaz G. et al. Epigenetic remodelling licences adult cholangiocytes for organoid formation and liver regeneration. Nat Cell Biol 2019; 21 (11) 1321-1333
  • 78 Rizvi F, Everton E, Smith AR. et al. Murine liver repair via transient activation of regenerative pathways in hepatocytes using lipid nanoparticle-complexed nucleoside-modified mRNA. Nat Commun 2021; 12 (01) 613
  • 79 Ota N, Shiojiri N. Comparative study on a novel lobule structure of the zebrafish liver and that of the mammalian liver. Cell Tissue Res 2022; 388 (02) 287-299
  • 80 Caviglia S, Unterweger IA, Gasiūnaitė A. et al. FRaeppli: a multispectral imaging toolbox for cell tracing and dense tissue analysis in zebrafish. Development 2022; 149 (16) 149
  • 81 Zhao C, Lancman JJ, Yang Y. et al. Intrahepatic cholangiocyte regeneration from an Fgf-dependent extrahepatic progenitor niche in a zebrafish model of Alagille Syndrome. Hepatology 2022; 75 (03) 567-583
  • 82 Wang J, Leng X, Wang G, Wan X, Cao H. The construction of intrahepatic cholangiocarcinoma model in zebrafish. Sci Rep 2017; 7 (01) 13419
  • 83 Goessling W, Sadler KC. Zebrafish: an important tool for liver disease research. Gastroenterology 2015; 149 (06) 1361-1377
  • 84 Evason KJ, Francisco MT, Juric V. et al. Identification of chemical inhibitors of β-catenin-driven liver tumorigenesis in zebrafish. PLoS Genet 2015; 11 (07) e1005305
  • 85 Den Broeder MJ, Kopylova VA, Kamminga LM, Legler J. Zebrafish as a model to study the role of peroxisome proliferating-activated receptors in adipogenesis and obesity. PPAR Res 2015; 2015: 358029
  • 86 Nguyen AT, Emelyanov A, Koh CH. et al. A high level of liver-specific expression of oncogenic Kras(V12) drives robust liver tumorigenesis in transgenic zebrafish. Dis Model Mech 2011; 4 (06) 801-813
  • 87 Vogt A, Cholewinski A, Shen X. et al. Automated image-based phenotypic analysis in zebrafish embryos. Dev Dyn 2009; 238 (03) 656-663
  • 88 Liu W, Chen JR, Hsu CH. et al. A zebrafish model of intrahepatic cholangiocarcinoma by dual expression of hepatitis B virus X and hepatitis C virus core protein in liver. Hepatology 2012; 56 (06) 2268-2276
  • 89 Shwartz A, Goessling W, Yin C. Macrophages in zebrafish models of liver diseases. Front Immunol 2019; 10: 2840
  • 90 Pham DH, Yin C. Zebrafish as a model to study cholestatic liver diseases. Methods Mol Biol 2019; 1981: 273-289
  • 91 Huang M, Chang A, Choi M, Zhou D, Anania FA, Shin CH. Antagonistic interaction between Wnt and Notch activity modulates the regenerative capacity of a zebrafish fibrotic liver model. Hepatology 2014; 60 (05) 1753-1766
  • 92 Ko S, Shin D. Chemical screening using a zebrafish model for liver progenitor cell-driven liver regeneration. Methods Mol Biol 2019; 1905: 83-90
  • 93 Cox AG, Saunders DC, Kelsey Jr PB. et al. S-nitrosothiol signaling regulates liver development and improves outcome following toxic liver injury. Cell Rep 2014; 6 (01) 56-69
  • 94 He JH, Guo SY, Zhu F. et al. A zebrafish phenotypic assay for assessing drug-induced hepatotoxicity. J Pharmacol Toxicol Methods 2013; 67 (01) 25-32
  • 95 North TE, Babu IR, Vedder LM. et al. PGE2-regulated Wnt signaling and N-acetylcysteine are synergistically hepatoprotective in zebrafish acetaminophen injury. Proc Natl Acad Sci U S A 2010; 107 (40) 17315-17320
  • 96 He J, Lu H, Zou Q, Luo L. Regeneration of liver after extreme hepatocyte loss occurs mainly via biliary transdifferentiation in zebrafish. Gastroenterology 2014; 146 (03) 789-800 .e8
  • 97 Choi TY, Ninov N, Stainier DY, Shin D. Extensive conversion of hepatic biliary epithelial cells to hepatocytes after near total loss of hepatocytes in zebrafish. Gastroenterology 2014; 146 (03) 776-788
  • 98 Ko S, Choi TY, Russell JO, So J, Monga SPS, Shin D. Bromodomain and extraterminal (BET) proteins regulate biliary-driven liver regeneration. J Hepatol 2016; 64 (02) 316-325
  • 99 He J, Zhou Y, Qian C. et al. DNA methylation maintenance at the p53 locus initiates biliary-mediated liver regeneration. NPJ Regen Med 2022; 7 (01) 21
  • 100 He J, Chen J, Wei X. et al. Mammalian target of rapamycin complex 1 signaling is required for the dedifferentiation from biliary cell to bipotential progenitor cell in zebrafish liver regeneration. Hepatology 2019; 70 (06) 2092-2106
  • 101 Zhang J, Zhou Y, Li S. et al. Tel2 regulates redifferentiation of bipotential progenitor cells via Hhex during zebrafish liver regeneration. Cell Rep 2022; 39 (01) 110596
  • 102 Khaliq M, Ko S, Liu Y. et al. Stat3 regulates liver progenitor cell-driven liver regeneration in zebrafish. Gene Expr 2018; 18 (03) 157-170
  • 103 Jung K, Kim M, So J, Lee SH, Ko S, Shin D. Farnesoid X receptor activation impairs liver progenitor cell-mediated liver regeneration via the PTEN-PI3K-AKT-mTOR axis in zebrafish. Hepatology 2021; 74 (01) 397-410
  • 104 Miyoshi T, Hidaka M, Miyamoto D. et al. Successful induction of human chemically induced liver progenitors with small molecules from damaged liver. J Gastroenterol 2022; 57 (06) 441-452
  • 105 Choi TY, Khaliq M, Tsurusaki S. et al. Bone morphogenetic protein signaling governs biliary-driven liver regeneration in zebrafish through tbx2b and id2a. Hepatology 2017; 66 (05) 1616-1630
  • 106 Ko S, Russell JO, Tian J. et al. Hdac1 regulates differentiation of bipotent liver progenitor cells during regeneration via Sox9b and Cdk8. Gastroenterology 2019; 156 (01) 187-202 .e14
  • 107 Cai P, Mao X, Zhao J. et al. Farnesoid X receptor is required for the redifferentiation of bipotential progenitor cells during biliary-mediated zebrafish liver regeneration. Hepatology 2021; 74 (06) 3345-3361
  • 108 So J, Kim M, Lee SH. et al. Attenuating the epidermal growth factor receptor-extracellular signal-regulated kinase-sex-determining region Y-box 9 axis promotes liver progenitor cell-mediated liver regeneration in zebrafish. Hepatology 2021; 73 (04) 1494-1508
  • 109 Michalopoulos GK, Barua L, Bowen WC. Transdifferentiation of rat hepatocytes into biliary cells after bile duct ligation and toxic biliary injury. Hepatology 2005; 41 (03) 535-544
  • 110 Limaye PB, Bowen WC, Orr AV, Luo J, Tseng GC, Michalopoulos GK. Mechanisms of hepatocyte growth factor-mediated and epidermal growth factor-mediated signaling in transdifferentiation of rat hepatocytes to biliary epithelium. Hepatology 2008; 47 (05) 1702-1713
  • 111 Limaye PB, Bowen WC, Orr A, Apte UM, Michalopoulos GK. Expression of hepatocytic- and biliary-specific transcription factors in regenerating bile ducts during hepatocyte-to-biliary epithelial cell transdifferentiation. Comp Hepatol 2010; 9: 9
  • 112 Nagahama Y, Sone M, Chen X. et al. Contributions of hepatocytes and bile ductular cells in ductular reactions and remodeling of the biliary system after chronic liver injury. Am J Pathol 2014; 184 (11) 3001-3012
  • 113 Yanger K, Zong Y, Maggs LR. et al. Robust cellular reprogramming occurs spontaneously during liver regeneration. Genes Dev 2013; 27 (07) 719-724
  • 114 Sekiya S, Suzuki A. Hepatocytes, rather than cholangiocytes, can be the major source of primitive ductules in the chronically injured mouse liver. Am J Pathol 2014; 184 (05) 1468-1478
  • 115 Lorent K, Moore JC, Siekmann AF, Lawson N, Pack M. Reiterative use of the notch signal during zebrafish intrahepatic biliary development. Dev Dyn 2010; 239 (03) 855-864
  • 116 Zong Y, Panikkar A, Xu J. et al. Notch signaling controls liver development by regulating biliary differentiation. Development 2009; 136 (10) 1727-1739
  • 117 Lemaigre FP. Development of the intrahepatic and extrahepatic biliary tract: a framework for understanding congenital diseases. Annu Rev Pathol 2020; 15: 1-22
  • 118 Hu S, Molina L, Tao J. et al. NOTCH-YAP1/TEAD-DNMT1 axis drives hepatocyte reprogramming into intrahepatic cholangiocarcinoma. Gastroenterology 2022; 163 (02) 449-465
  • 119 Fan B, Malato Y, Calvisi DF. et al. Cholangiocarcinomas can originate from hepatocytes in mice. J Clin Invest 2012; 122 (08) 2911-2915
  • 120 Molina L, Nejak-Bowen K, Monga SP. Role of YAP1 signaling in biliary development, repair, and disease. Semin Liver Dis 2022; 42 (01) 17-33
  • 121 Lee DH, Park JO, Kim TS. et al. LATS-YAP/TAZ controls lineage specification by regulating TGFβ signaling and Hnf4α expression during liver development. Nat Commun 2016; 7: 11961
  • 122 Planas-Paz L, Sun T, Pikiolek M. et al. YAP, but not RSPO-LGR4/5, signaling in biliary epithelial cells promotes a ductular reaction in response to liver injury. Cell Stem Cell 2019; 25 (01) 39-53 .e10
  • 123 Li W, Yang L, He Q. et al. A homeostatic Arid1a-dependent permissive chromatin state licenses hepatocyte responsiveness to liver-injury-associated YAP signaling. Cell Stem Cell 2019; 25 (01) 54-68 .e5
  • 124 Liu Y, Zhuo S, Zhou Y. et al. Yap-Sox9 signaling determines hepatocyte plasticity and lineage-specific hepatocarcinogenesis. J Hepatol 2022; 76 (03) 652-664
  • 125 Wang J, Dong M, Xu Z. et al. Notch2 controls hepatocyte-derived cholangiocarcinoma formation in mice. Oncogene 2018; 37 (24) 3229-3242
  • 126 Yimlamai D, Christodoulou C, Galli GG. et al. Hippo pathway activity influences liver cell fate. Cell 2014; 157 (06) 1324-1338
  • 127 Schaub JR, Huppert KA, Kurial SNT. et al. De novo formation of the biliary system by TGFβ-mediated hepatocyte transdifferentiation. Nature 2018; 557 (7704): 247-251
  • 128 Tschaharganeh DF, Xue W, Calvisi DF. et al. p53-dependent Nestin regulation links tumor suppression to cellular plasticity in liver cancer. Cell 2014; 158 (03) 579-592
  • 129 Jeliazkova P, Jörs S, Lee M. et al. Canonical Notch2 signaling determines biliary cell fates of embryonic hepatoblasts and adult hepatocytes independent of Hes1. Hepatology 2013; 57 (06) 2469-2479
  • 130 Guo J, Fu W, Xiang M. et al. Notch1 drives the formation and proliferation of intrahepatic cholangiocarcinoma. Curr Med Sci 2019; 39 (06) 929-937
  • 131 Pham DH, Kudira R, Xu L. et al. Deleterious variants in ABCC12 are detected in idiopathic chronic cholestasis and cause intrahepatic bile duct loss in model organisms. Gastroenterology 2021; 161 (01) 287-300 .e16
  • 132 Lee SH, So J, Shin D. Hepatocyte-to-cholangiocyte conversion occurs through transdifferentiation independently of proliferation in zebrafish. Hepatology 2023, in press
  • 133 Russell JO, Ko S, Monga SP, Shin D. Notch inhibition promotes differentiation of liver progenitor cells into hepatocytes via sox9b repression in zebrafish. Stem Cells Int 2019; 2019: 8451282
  • 134 Everton E, Rizvi F, Smith AR. et al. Transient yet robust expression of proteins in the mouse liver via intravenous injection of lipid nanoparticle-encapsulated nucleoside-modified mRNA. Bio Protoc 2021; 11 (19) e4184
  • 135 Pepe-Mooney BJ, Dill MT, Alemany A. et al. Single-cell analysis of the liver epithelium reveals dynamic heterogeneity and an essential role for YAP in homeostasis and regeneration. Cell Stem Cell 2019; 25 (01) 23-38 .e8
  • 136 Bou Saleh M, Louvet A, Ntandja-Wandji LC. et al. Loss of hepatocyte identity following aberrant YAP activation: a key mechanism in alcoholic hepatitis. J Hepatol 2021; 75 (04) 912-923
  • 137 Glaser SS, Gaudio E, Miller T, Alvaro D, Alpini G. Cholangiocyte proliferation and liver fibrosis. Expert Rev Mol Med 2009; 11: e7
  • 138 Kaur S, Siddiqui H, Bhat MH. Hepatic progenitor cells in action: liver regeneration or fibrosis?. Am J Pathol 2015; 185 (09) 2342-2350
  • 139 Katoonizadeh A, Nevens F, Verslype C, Pirenne J, Roskams T. Liver regeneration in acute severe liver impairment: a clinicopathological correlation study. Liver Int 2006; 26 (10) 1225-1233
  • 140 Rodrigo-Torres D, Affò S, Coll M. et al. The biliary epithelium gives rise to liver progenitor cells. Hepatology 2014; 60 (04) 1367-1377
  • 141 Furuyama K, Kawaguchi Y, Akiyama H. et al. Continuous cell supply from a Sox9-expressing progenitor zone in adult liver, exocrine pancreas and intestine. Nat Genet 2011; 43 (01) 34-41
  • 142 Font-Burgada J, Shalapour S, Ramaswamy S. et al. Hybrid periportal hepatocytes regenerate the injured liver without giving rise to cancer. Cell 2015; 162 (04) 766-779
  • 143 Lesaffer B, Verboven E, Van Huffel L. et al. Comparison of the Opn-CreER and Ck19-CreER drivers in bile ducts of normal and injured mouse livers. Cells 2019; 8 (04) 8
  • 144 Morales-Ibanez O, Domínguez M, Ki SH. et al. Human and experimental evidence supporting a role for osteopontin in alcoholic hepatitis. Hepatology 2013; 58 (05) 1742-1756
  • 145 Uede T. Osteopontin, intrinsic tissue regulator of intractable inflammatory diseases. Pathol Int 2011; 61 (05) 265-280
  • 146 Zweerink S, Mueck V, Kraemer LP. et al. Repolarization precedes oval cell-mediated hepatocytic regeneration in the CDE diet mouse model. J Histochem Cytochem 2022; 70 (05) 377-389
  • 147 Sekiya S, Suzuki A. Intrahepatic cholangiocarcinoma can arise from Notch-mediated conversion of hepatocytes. J Clin Invest 2012; 122 (11) 3914-3918
  • 148 Yoshii D, Shimata K, Yokouchi Y. et al. SOX9 contributes to the progression of ductular reaction for the protection from chronic liver injury. Hum Cell 2022; 35 (02) 721-734

Zoom Image
Fig. 1 Mechanisms of hepatobiliary plasticity. Mature hepatocytes and BECs are known to divide to overcome minor injuries (no plasticity, left panel); however, when injuries are chronic or more severe, plasticity of the hepatobiliary compartment is observed (plasticity, right panel). In this case, various options have been proposed: (A) Resident LPCs self-renew and differentiate into both hepatocytes and BECs, (B) Facultative LPCs emerge following liver injury, self-renew and differentiate as the resident LPCs do, and (C) mature hepatocytes and BECs transdifferentiate into each other without transitioning through a LPC stage. BECs, biliary epithelial cells; LPC, liver progenitor cell.
Zoom Image
Fig. 2 Molecular mechanisms driving hepatobiliary plasticity. Studies in rodents and zebrafish together with observations in human specimens have revealed positive and negative molecular regulators implicated in BEC-to-hepatocytes conversion following hepatocyte injury (A) and in hepatocyte-to-BEC conversion following BEC injury (B). Factors written with capital and small letters are related to mouse and zebrafish studies, respectively. Factors with an asterisk ([*]) are related to both mouse and zebrafish studies. BEC, biliary epithelial cell.