• Non ci sono risultati.

MAP Kinase Pathways in the Control of Hepatocyte Growth, Metabolism and Survival

N/A
N/A
Protected

Academic year: 2022

Condividi "MAP Kinase Pathways in the Control of Hepatocyte Growth, Metabolism and Survival"

Copied!
16
0
0

Testo completo

(1)

MAP Kinase Pathways

in the Control of Hepatocyte Growth, Metabolism and Survival

Paul Dent

19

19.1

Introduction

In the past 15 years, multiple new signal transduction pathways within cells have been discovered. Many of these pathways belong to what is now termed “the mitogen-activated protein kinase (MAPK) super- family.” The roles of each MAPK pathway in the control of hepatocyte growth, metabolism and sur- vival are still under investigation. It is believed that exposure of hepatocytes to growth factors induces simultaneous activation of multiple MAPK path- ways that act in a coordinated manner to control cell growth, survival and metabolism. In addition, bile acids have more recently been shown to activate growth factor and death receptors in hepatocytes, which can interact with growth factor-induced sig- naling to generate a further layer of complexity in the activation of MAPK pathways. The expression and release of growth factors such as tumor necro- sis factor- α (TNF-α) from Kupffer cells during liver regeneration also has the potential to modulate the responses of MAPK pathways in hepatocytes. Thus the ability of growth factors and bile acids to acti- vate MAPK signaling pathways in hepatocytes may depend on the expression of multiple growth factor receptors, the synthesis of paracrine factors and bile acid levels. This review will describe the enzymes within four major MAPK signaling pathways and discuss their activation and roles in hepatic prolif- erative, survival and metabolic responses.

19.2

The "Classical" Mitogen-Activated Protein Kinase/Extracellular Regulated Kinase Pathway

"Mitogen-activated protein-2 kinase" was first re- ported by the laboratory of Dr. Thomas Sturgill in 1986 [131]. This protein kinase was originally described as a 42-kDa insulin-stimulated protein

kinase activity whose tyrosine phosphorylation in- creased after insulin exposure, and which phospho- rylated the cytoskeletal protein MAP-2. Contempo- raneous studies from the laboratory of Dr. Melanie Cobb identified an additional 44-kDa isoform of this enzyme, termed extracellular signal-regulated kinase (ERK1) [9]. Since many growth factors and mitogens could activate these enzymes, the acronym for this enzyme was subsequently changed to denote mitogen-activated protein (MAP) kinase. Addition- al studies demonstrated that the p42 (ERK2) and p44 (ERK1) MAP kinases regulated another protein kinase activity (p90

rsk

) [132], and that they were themselves regulated by protein kinase activities designated MKK1/2 (MAPK kinase; MAP2K), also termed MEK1/2 [154, 155]. MKK1 and MKK2 were also regulated by reversible phosphorylation. The protein kinase responsible for catalyzing MKK1/2 activation was initially described as the proto-on- cogene Raf-1 [19, 41]. This was soon followed by an- other MEK1/2-activating kinase, termed MEKK1, which was a mammalian homolog with similarity to the yeast Ste 11 and Byr 2 genes [74]. However, fur- ther studies have shown that the primary function of MEKK1 is to regulate the c-Jun NH2-terminal kinase (JNK)1/2, rather than the ERK1/2, pathway [157].

19.2.1

Growth Factor Receptors, Ras and Raf

Plasma membrane receptors transduce signals through the membrane to its inner leaflet, leading to the recruitment and activation of guanine nucle- otide exchange factors, which increase the amount of GTP bound to membrane-associated GTP-bind- ing proteins, in particular Ras family proteins [97]

(Fig. 19.1). There are four widely recognized isofor-

ms of Ras: Harvey (H), Kirsten (K4A, K4B) and Neu-

roblastoma (N) [114]. Receptor-stimulated guanine

nucleotide exchange of “Ras” to the GTP-bound

form permits Raf proteins and P110 phosphatidyl

(2)

inositol (PI) 3-kinase to associate with “Ras”, result- ing in kinase translocation to the plasma membrane environment where activation of these kinases takes place. Ras contains a GTPase activity that converts bound GTP to GDP resulting in inactivation of the Ras molecule. Mutation of Ras results in a loss of GTPase activity, generating a constitutively active Ras molecule that can lead to elevated activity with- in downstream signaling pathways. Approximately one third of human cancers have Ras mutations, pri- marily the K-Ras isoforms; however, Ras mutation is not a common occurrence in hepatoma, whereas it is relatively common in cholangiocarcinoma [136]. Of note, B-Raf mutations have also been shown to play a role in tumorigenesis in thyroid and cholangiocar- cinoma but not in hepatoma [63, 136].

Raf-1 is a member of a family of serine/threonine protein kinases also comprising B-Raf and A-Raf [104]. All “Raf” family members can phosphorylate MKK1/2 and activate the ERK1/2 pathway [140].

Thus the “Raf” kinases act at the level of a MAPK

kinase kinase (MAP3K). The NH

2

domain of Raf- 1 can reversibly interact with Ras family members in the plasma membrane and the ability of Raf-1 to associate with Ras is dependent upon the Ras mol- ecule being in the GTP-bound state [84, 92, 143].

Additional protein serine/threonine and tyrosine phosphorylation(s) are also known to play a role increasing Raf-1 activity when in the plasma mem- brane environment [20]. PKC isoforms, which can be activated by bile acids, have been proposed to be Raf-1-activating kinases [11]. The initial biochemi- cal analyses of purified Raf-1 demonstrated consti- tutive Y340-Y341 phosphorylation when the protein was co-expressed with Src [20, 41]. PI-3 kinase-de- pendent phosphorylation of S338 by PAK family en- zymes may facilitate Raf-1 tyrosine phosphorylation by Src family members, leading to full activation of Raf-1 [64]. PAK enzymes may also phosphorylate MEK1, promoting Raf-1–MEK1 association and thus pathway activation [14, 28]. In this regard, bile acids, compared to epidermal growth factor (EGF), potently enhance tyrosine phosphorylation of Raf-1 [111].

Phosphorylation of Raf-1 at S259 by either Akt or the cAMP-dependent-protein kinase (PKA) can inhibit Raf-1 activity and its activation by upstream stimuli [24]. Phosphorylation of Raf-1 at S43 by PKA inhibits the interaction of Raf-1 with Ras molecules, thereby blocking Raf-1 translocation to the plasma membrane and its Ras-dependent activation [156].

In contrast to Raf-1, the B-Raf isoform does not con- tain an equivalent to S43, but contains multiple sites of Akt (and potentially PKA)-mediated phosphor- ylation in addition to the B-Raf equivalent of S259 [42]. B-Raf can be activated by both Ras and cAMP via the Rap1 GTPase [162]. Thus the regulation of the MEK1/2-ERK1/2 pathway is very complex, and in some cell types may be both inhibited by cAMP/

PKA, through Raf-1, and stimulated by cAMP/Rap1, through B-Raf.

These findings may have considerable impact on catecholamine signaling in hepatocytes during liver regeneration. During hepatocyte proliferation/liver regeneration, the expression of α

1

-adrenergic re- ceptors declines and the levels of β

2

-adrenergic re- ceptors increase. In non-proliferating hepatocytes, the α

1

-adrenergic receptor agonist phenylephrine enhances Raf-1 and ERK1/2 activity in a similar manner to growth factors such as EGF, and gluca- gon inhibits phenylephrine or EGF-induced ERK1/2 activity by blocking Raf-1 activation [129]. In prolif- erating hepatocytes, phenylephrine cannot activate Raf-1/ERK1/2 and furthermore, the β

2

-adrenergic receptor agonist isoproteranol inhibits EGF-in- duced ERK1/2 activation. As ERK1/2 signaling has been linked to a proliferative response in hepato-

Fig. 19.1. Some of the characterized signal transduction path- ways in mammalian cells. Growth factor receptors, e.g., the ERBB family, insulin receptor can signal down through GTP-binding proteins into multiple intracellular signal transduction path- ways. Predominant among these pathways are the MAP kinase superfamily of cascades (ERK1/2, ERK5, JNK, p38). Growth fac- tor receptors, Ras proteins and downstream pathways are often activated in tumor cells and inhibitors have been developed to block the function of these molecules, thereby slowing cell growth and promoting cell death responses. Multiple inhibi- tors for the ERBB family receptors have been developed, e.g., the ERBB1 inhibitor AG1478. Inhibitors of Ras farnesylation (and geranylgeranylation) are in clinical trials, as are inhibitors of the ERK1/2/5 pathway. It should be noted that MEK1/2 inhibitors PD98059 and U0126 are also capable of inhibiting the “Big” MAP kinase pathway via blocking activation of MEK5

(3)

cytes, these findings suggest that catecholamines may initially promote, and then inhibit, rodent liver regeneration.

In gene deletion studies, loss of Raf-1 function was embryonically lethal due to weak placental an- giogenesis and hepatoblast apoptosis [152, 153]: in Raf-1 null hepatocytes, ERK1/2 was activated by B- Raf rather than Raf-1, which suggests: (a) fetal hepa- tocytes may utilize different “Raf” molecules to ac- tivate the ERK1/2 pathway in comparison to adult hepatocytes and (b) deletion of Raf-1 caused hepa- tocytes to survive by recruiting in compensation B- Raf as the MAP3K activator for ERK1/2, which will tend to promote growth arrest over proliferation.

With reference to comment (a), the relative ability of Raf-1 and B-Raf to activate the ERK1/2 pathway in primary hepatocytes and established HepG2 and HuH7 hepatoma cells is also different: in primary hepatocytes and HepG2 cells, ERK1/2 pathway acti- vation by growth factors is dependent on Raf-1 and inhibited by cAMP [3], whereas in HuH7 cells, B-Raf and ERK1/2 activation can be enhanced by cAMP (P.B. Hylemon and P. Dent, unpublished data). This implies primary and fetal hepatocytes, and estab- lished liver-derived cell lines can have very different signaling behavioral characteristics.

19.2.2

Hepatocyte Proliferation and ERK1/2 Signaling

There are multiple studies that link ERK1/2 signal- ing to hepatocyte proliferation and present evidence that inhibition of MEK1/2/5 blunts the induction of proteins essential for G1 phase progression and S phase entry, such as cyclin D1, PCNA and cyclin A [15, 22, 27, 34, 51, 61, 77, 100, 101, 106, 113, 135, 140]. Activation of ERK1/2 by constitutive or induc- ible Raf-1 and MEK1 constructs can enhance cyclin D1 expression in hepatocytes [22, 34, 100, 101]. One in vivo study using PD98059 has shown inhibition of liver regeneration; however, as PD98059 causes acute kidney toxicity, these results may not be solely due to direct effects on liver biology [135]. PD98059 acts as an inhibitor of “Raf”-mediated phosphoryla- tion of MEK1/2/5 molecules but it is not a kinase domain inhibitor of “Raf” proteins or of activated MEK1: in cells PD98059 is a relatively good inhibitor of MEK1 and MEK5 (IC50 ~5 µM) but a much poorer inhibitor of MEK2 (IC50 ~40 µM) [17]. U0126 is re- ported as a more equipotent inhibitor of MEK1/2/5 that blocks both activating phosphorylations on the MEK1/2/5 proteins as well as their kinase domain activity, with IC50 values in the ~0.5 µM range [17].

The clinically used MEK1/2 inhibitor PD184352

(CI1040) acts in a similar potent manner to U0126 and at concentrations below 10 µM has shown spe- cificity for MEK1/2 over MEK5 [91, 130]; it may be a useful tool to demarcate between MEK1/2-ERK1/2 and MEK5-ERK5 signaling dependencies.

The livers of mice deleted for insulin growth factor 1-binding protein, which regenerate poorly, have more recently been shown to lack an ERK1/2 activation response after partial hepatectomy as well as expression of the ERK1/2 downstream tar- get C/EBP β [6, 77]. The role of the cyclin-depend- ent kinase (CDK) inhibitor p21

Cip-1/WAF1/mda6

(p21) and other CDK inhibitors in hepatocyte growth has also been extensively investigated [3, 6, 52, 87, 93, 142, 148]. These findings are very similar to those in other cell types. Hepatocytes from older rodents, which have reduced liver regeneration capability, have higher basal expression of the cyclin kinase inhibitor p21 and lower growth factor-induced ac- tivation of ERK1/2 [138]. These findings are also in agreement with data arguing that the livers of p21 – /– mice regenerate more quickly than those of wild- type animals [1]. In hepatocytes ERK1/2 signaling positively regulates the p21 promoter via the tran- scription factors C/EBP β and Ets2, and to a lesser extent C/EBP α [100, 101]. In addition, ERK1/2 sig- naling enhances p21 mRNA stability and increases p21 protein half life (potentially by phosphorylation at T57 and S130) [62]. In hepatocytes that cannot express p21, proliferation is stimulated by intense ERK1/2 signaling that is dependent on C/EBP β and Ets2 [10, 101].

A number of other studies using various exog-

enous agents that cause prolonged intense ERK1/2

pathway activation in hepatocytes and hepatoma

cells have suggested that this signal is antiprolif-

erative. The novel vitamin K analog Cpd5 and vi-

tamin K3 (menadione) potently enhanced ERK1/2

signaling in hepatocytes, which was associated with

growth suppression [57, 145]. In HepG2 cells stably

transfected with the α

1

-adrenergic receptor, phe-

nylephrine induced prolonged intense ERK1/2 sig-

naling that was responsible for induction of p21 and

growth arrest; antisense mRNA ablation of p21 pro-

tein expression converted prolonged intense ERK1/2

activation into a growth-promoting signal [3, 4],

which was subsequently confirmed [142]. Acute

ethanol exposure of hepatocytes in vitro combined

with growth factor treatment promotes superinduc-

tion of ERK1/2 signaling, and of p21 and p16 expres-

sion, leading to growth arrest, which is in general

agreement with the known antiproliferative effects

of ethanol on liver regeneration [140]. Chronic etha-

nol exposure of animals appears to block liver regen-

eration, not by promoting, but instead by inhibiting,

growth factor-induced activation of ERK1/2 [13] and

(4)

by promoting a p38 MAPK-dependent increase in p21 expression [68]. Chronic exposure of animals to xenobiotic chemical agents that block liver regener- ation can also alter the ability of the liver function- ally to activate the ERK1/2 pathway, for example, the carcinogenic drug 2-acetylamino fluorine does not block EGF-induced activation of ERK1/2 but pre- vents nuclear translocation of ERK1/2, blunting en- hanced transcription factor function(s) [127]. Hepa- titis B virus (HBV) X protein enhances or suppress- es p21 expression in a cell type-dependent manner [67, 75, 76, 102, 106]. For example, in some estab-

lished hepatocyte cell types, X protein enhances ERK1/2 and promotes cell proliferation [7, 66], whereas in primary cells it can promote activation of MAPK pathways and enhance expression of p21 and p27

Kip-1

, leading to growth arrest [75, 106]. Genetic differences between primary and established hepa- tocyte cells as well as different expression levels of X protein in cells may explain these discrepancies.

In many tumor cell types, it was initially noted that ERK1/2 activity was elevated compared to non- transformed “normal” tissues. Immunohistochemi- cal analyses have shown that liver tumors have ele- vated levels of phospho-ERK1/2 [51]. A strong corre- lation between reduced hepatoma cell proliferation and the pathway inhibitory actions of MEK1/2/5 in- hibitors has been noted, suggesting that the ERK1/2 pathway could be a useful therapeutic target in the treatment of hepatocellular carcinoma [e.g., 3, 53].

19.2.3

ERK1/2 Signaling, Apoptosis and Bile Acids

Activation of the ERK1/2 pathway in many cell types is associated not only with proliferation but also with protection from toxic stresses. For exam- ple, exposure of hepatoma cells to transforming growth factor beta (TGF β) can increase ERK1/2, JNK1/2 and p38 activity: inhibition of ERK1/2 sig- naling promotes further JNK1/2 activation and en- hances cell death [103]. Serum and insulin have also

Fig. 19.2a, b. Inhibition of ERK1/2 and PI-3 kinase enhances bile acid toxicity in hepatocytes. a Hepatocytes were treated with the indicated bile acid (50 µM) in the presence or absence of either PD98059 (50 µM), U0126 (10 µM) or PD184352 (10 µM) for 6 h. Ap- optosis was determined by morphological examination of slides as noted in [105]. b Hepatocytes were treated with either 1 µM AG1478 (ERBB1 inhibitor) or 1 µM AG1024 (INS-R, IGF-1-R inhibi- tor) in the presence or absence of DCA (50 µM) for 0–120 min.

ERK1/2, AKT and GSK3 activity was determined by phospho-im- munoblotting as described [105, 110]

Fig. 19.3. Inhibition of ERBB1 using AG1478 promotes and in- hibits DCA-induced apoptosis, which is dependent on the con- centration of DCA. Hepatocytes were treated with 1 µM AG1478 (ERBB1 inhibitor) in the presence or absence of DCA (50 µM;

250 µM; 450 µM) for 6 h. Apoptosis was determined by mor- phological examination of slides as noted in [105, 110]. Arrows indicate enhancement, no effect or inhibition of apoptosis by AG1478

(5)

been shown to mediate their anti-apoptotic effects in hepatoma cells via ERK1/2 signaling [56, 90]. In some instances, prolonged intense ERK1/2 activa- tion has been shown to promote cell death. The agent Cpd5 can cause prolonged intense ERBB1-ERK1/2 activation in hepatocytes, which is responsible for cell death [144]. Menadione and fatty acids promote intense ERK1/2 activation in established RALA255- 10G hepatocytes: yet inhibition of ERK1/2 signaling promotes menadione toxicity but inhibits fatty acid toxicity (Dr. M. Czaja, personal communication).

In general, it appears that ERK1/2 (and PI-3 kinase) activation is a compensatory survival re- sponse to balance potential toxic signals, such as bile acid exposure. This laboratory has shown that unconjugated forms of deoxycholic acid (DCA), CDCA and ursodeoxycholic acid (UDCA) all acti- vate the ERK1/2 pathway in primary hepatocytes, which was partially dependent on the PI-3 ki- nase pathway, and that inhibition of MEK1/2 with

PD98059, U0126 or PD184352 enhances bile acid toxicity [105, 108–110] (Fig. 19.2a). Similar data for ERK1/2 protection against DCA toxicity have been generated in colon cancer cells [107]. Inhibition of PI-3 kinase with LY294002 is relatively less effec- tive at promoting unconjugated bile acid-induced apoptosis compared to that caused by a MEK1/2 inhibitor [105, 108]. In contrast to these findings, MEK1/2 inhibitors more weakly enhanced the tox- icity of conjugated bile acids, whereas PI-3 kinase inhibition more efficaciously promoted conjugated bile acid toxicity (P. Dent and P.B. Hylemon, unpub- lished observations) [119, 134, 147]. One likely possi- bility is that conjugated and unconjugated bile acids differentially activate alternate cassettes of growth factor receptors, in different subcellular locations, and thus the ERK1/2 and PI-3 kinase signaling path- ways, in a dose-dependent fashion.

Several independent studies have demonstrated that bile acid-induced ERK1/2 signaling is depend-

Fig. 19.4. Possible mechanisms by which ERK1/2 signaling could protect hepatocytes from toxic stresses. Apoptosis occurs following activation of effector caspases (e.g., caspases-3, -6, -7), which can be triggered by either the extrinsic or intrinsic path- ways. The extrinsic pathway is characteristically initiated by liga- tion of the Fas ligand with its receptor, leading to formation of the death-inducing signaling complex (DISC), which permits the Fas-associated death domain (FADD) to cleave and activate pro- caspase-8. Activated caspase-8 can activate effector caspases such as procaspase-3 or initiate mitochondrial injury via BID. The intrinsic, or mitochondrial pathway, becomes engaged follow- ing mitochondrial injury (e.g., loss of mitochondrial membrane potential; and/or release of pro-apoptotic proteins such as cyto- chrome c). Cytochrome c, in association with dATP, promotes the caspase-9-mediated activation of procaspase-3. Considerable

crosstalk exists between the intrinsic and extrinsic pathways.

For example, while caspase-8 directly activates caspase-3, it also cleaves and activates the pro-apoptotic BH3-only domain Bcl-2 family member BID, which then triggers cytochrome c release and results in further procaspase-3 activation. A large and ex- panding group of pro- and anti-apoptotic Bcl-2 family proteins has been described, which may act by modulating BAD/BIM interactions and mitochondrial pore function, or, in the case of IAP proteins, including c-FLIP molecules, by directly inhibiting caspase activation. The relevance of apoptosis for hepatic cell biology is underscored by accumulating evidence that diverse signaling pathways downstream of ERBB receptors, e.g., the ERK1/2 pathway, regulate cell survival and response to bile acids by modulating the apoptotic threshold

(6)

ent on the activation of growth factor receptors in hepatocytes, predominantly of the ERBB family, and, furthermore, in hepatoma cells, by paracrine ligands such as TGF α [47, 105, 108, 149, 160]. Oth- ers have shown some bile acids (TLCA) can bind to and directly activate serpentine receptors [60]. Ac- tivation of ERBB1 and the insulin receptor in pri- mary rodent hepatocytes by bile acids is dependent on mitochondria-dependent generation of reactive oxygen/nitrogen species and the inactivation of pro- tein tyrosine phosphatases (PTPase) [33, 105]. The generation and/or modulation of cellular reactive oxygen species (ROS) levels by bile acids has been described in detail by several groups [116, 117, 159].

As PTPases are ~100-fold more active as enzymes than protein kinases, a small reduction in cellu- lar PTPase activity will result in an increase in the steady-state tyrosine phosphorylation of the pro- tein kinases, e.g., receptor tyrosine kinases, whose phosphorylation they suppress. Hence, inhibition of DCA-stimulated ERBB1 or the insulin receptor sign- aling significantly blunts DCA-induced ERK1/2 ac- tivation (Fig. 19.2b). In some cell types, the insulin receptor can promote activation of ERBB1 via release of TGF α, as has been observed for bile acid-induced activation of ERBB1 in HuH7 hepatoma cells [149].

Treatment of primary hepatocytes with 50 µM DCA causes almost no induction of apoptosis within 6 h;

however, DCA toxicity is significantly enhanced by AG1478, an ERBB1 inhibitor (Fig. 19.3) [105]. Treat- ment of hepatocytes with either 250 µM or 450 µM DCA causes a considerably greater increase in hepa- tocyte apoptosis within 6 h: AG1478 does not poten- tiate 250 µM DCA-induced killing, despite signifi- cantly blocking PI-3 kinase and ERK1/2 activation [47, 105], and reduces 450 µM DCA-induced hepa- tocyte death. These findings may explain the con- flicting findings reported by different groups using lower and higher concentrations of toxic bile acids, e.g., [105, 112]. Hence bile acid-induced activation of growth factor receptors may promote both protec- tive and toxic cellular responses that are dependent on the concentration, and presumably duration, of bile acid exposure. It should be noted that all of the above findings in bile acid-treated hepatocytes have similarities to observations made in carcinoma cells exposed to UV and ionizing radiation [21].

The mechanisms by which ERK1/2 signaling protects hepatocytes from toxic stresses are likely to be found at multiple levels within the cell (Fig. 19.4).

First, ERK1/2 signaling regulates the activities of transcription factors that are known to elicit cy- toprotective responses, e.g., CREB, C/EBP β [101, 108]. These factors in turn can alter the transcrip- tion of diverse pro- and anti-apoptotic genes such as PUMA, c-FLIP isoforms, Mcl-1 and Bcl-

XL

[8, 55,

98, 110]. Second, in various cells ERK1/2 and p90

rsk

may directly phosphorylate and inhibit the func- tions of pro-apoptotic proteins, e.g., BAD, BIM, pro- caspase 9 [2, 32, 69, 78, 79, 105, 110, 121, 147]. Third, ERK1/2 pathway signaling can promote or inhibit the expression of both protective and toxic growth factors, e.g., TNF- α, EGF [49]. Finally, bile acid-in- duced ERK1/2 signaling could potentially regulate both the expression of MDM2, the E3 ligase that can ubiquinate the tumor suppressor p53; as well as ex- pression of p19

ARF

, which blocks the ubiquination of p53 by MDM2. In a variety of cell types the bal- ance of MDM2 versus p19 expression can determine p53 levels, and those of p21, as well as the apoptotic threshold [115]. While expression of p21 in many cell systems has been associated with a protection of toxic stresses, p21 can enhance bile acid toxicity in hepatocytes [72, 109]. Thus ERK1/2 signaling may have a dual protective and toxic nature in hepato- cytes; short-term activation of ERK1/2 promotes cell survival after a toxic insult, whereas prolonged ERK1/2 signaling (>~24 h) may enhance cell killing [144].

19.2.4

ERK1/2 Signaling and Metabolism

Activation of ERK1/2 can enhance plasma mem- brane translocation of the glucose transporter GLUT4 in adipocytes and myotubes [120]. In some cell types, ERK1/2 signaling can also control the activity of glycogen synthase kinase 3 (GSK3) and that of serine/threonine protein phosphatase 1, co- localized on the glycogen particles, which could in- fluence the rate of glycogen synthesis [30, 122]. Bile acids regulate glycogen synthase activity in hepato- cytes via the insulin-R/PI-3 kinase/GSK3 pathway to the same degree as exposure to 50 nM insulin [47]. In hepatocytes and hepatoma cells, growth fac- tor and bile acid-induced ERK1/2 signaling controls the expression of the LDL receptor and plays a role in stimulating the excretion of bile salts leading to increased canalicular transport capacity [70, 111].

19.3

The JNK Pathway

c-Jun NH

2

-terminal kinase 1 and 2 were initially de-

scribed biochemically to be stress-induced protein

kinase activities that phosphorylated the NH

2

-ter-

minus of the transcription factor c-Jun; hence the

pathway is often called the stress-activated protein

kinase (SAPK) pathway [23, 48]. Multiple stresses

(7)

increase JNK1/2 (and the subsequently discovered JNK3) activity including UV- and γ-irradiation, cy- totoxic drugs, bile acids and ROS (e.g., H

2

O

2

). Phos- phorylation of the NH

2

-terminal sites Ser63 and Ser73 in c-Jun increases its ability to trans-activate AP-1 enhancer elements in the promoters of many genes. It has been suggested that JNK can phos- phorylate the NH2-terminus of c-Myc, potentially playing a role in both proliferative and apoptotic signaling [96]. In a similar manner to the previously described ERK1/2 MAPK pathway, JNK1/2 activities were regulated by dual threonine and tyrosine phos- phorylation, which were found to be catalyzed by a protein kinase analogous to MKK1/2, termed stress- activated extracellular regulated kinase 1 (SEK1), also called MKK4 [23, 141]. An additional isoform of MKK4, termed MKK7, was subsequently discovered [141]. As in the case of MKK1 and MKK2, MKK4 and MKK7 were regulated by dual serine phosphoryla- tion.

In contrast to the ERK1/2 pathway, however, which appears primarily to utilize the three protein kinases of the Raf family to activate MKK1/2, at least ten protein kinases are known to phosphorylate and activate MKK4/7, including the Ste 11/Byr2-ho- mologs MKKK1–4, as well as proteins such as TAK-1 and Tpl-2 [74, 123, 124, 157]. Cleavage of MEKK1 by caspase molecules into a constitutively active mol- ecule may play an amplifying role in the execution of apoptotic processes [124, 151].

Upstream of the MAP3K enzymes are another layer of JNK pathway protein kinases, e.g., Ste20-ho- mologs and low molecular weight GTP-binding pro- teins of the Rho family, in particular Cdc42 and Rac1 (Fig. 19.1) [12, 36, 163]. It is not clear how growth factor receptors, e.g., ERBB1, activate the Rho fam- ily low molecular weight GTP-binding proteins; one mechanism may be via the Ras proto-oncogene, whereas others have suggested via PI-3 kinase and/

or protein kinase C isoforms [139]. In addition, other groups have shown that agonists acting through the TNF- α and FAS receptors, via sphingomyelinase en- zymes generating the lipid second-messenger cera- mide, can activate the JNK pathway by mechanisms that may act through Rho family GTPases.

19.3.1

Hepatocytes and JNK1/2 Signaling

c-Jun NH

2

-terminal kinase signaling plays an essen- tial role in hepatocyte proliferation. Genomic dele- tion of c-Jun was shown to be embryonically lethal due to a lack of liver development, and in a more recently published study, deletion of MKK4 (SEK1) has also been found to block liver development in

vivo [37]. The in utero lethality of c-Jun –/– and MKK4 –/– animals was enhanced in c-Jun –/– MKK4 –/– embryos, suggesting that the JNK pathway sig- naling regulates transcription factors in addition to c-Jun to promote liver growth and survival [29, 37, 95, 146, 150]. In contrast, the embryonic lethality of MEK1 and Raf-1 deletions has been shown to be dependent on a lack of vasculature development and to a lesser extent on hepatoblast survival [39].

In vitro studies examining the role of JNK pathway signaling in the control of hepatocyte DNA synthe- sis have largely agreed with the embryonic lethality studies. Interruption of JNK pathway signaling by use of dominant negative MKK4, JNK1 and c-Jun blocks growth factor-induced activation of the AP- 1 transcription factor complex and hepatocyte DNA synthesis in vitro [5, 26, 150]. Studies using high doses of the JNK1/2 inhibitor SP600125 have also suggested JNK signaling controls expression of cy- clin D1 and liver regeneration in vivo; however, at the concentration used, SP600125 may also have had non-specific inhibitory effects on other MAPK path- ways [125]. Liver regeneration/hepatocyte prolifera- tion is controlled by many growth factors, some of which potently activate the JNK pathway, includ- ing TNF- α, insulin and to a lesser extent EGF and hepatocyte growth factor (HGF). TNF- α, primarily generated by Kupffer cells [126], can stimulate the proliferation of hepatocytes. TNF- α signaling also activates the transcription factor NF- κB, which acts in a coordinated fashion with JNK1/2 and ERK1/2 to promote hepatocyte growth and protect hepato- cytes from TNF- α toxicity [65, 118].

19.3.2

JNK1/2 Signaling, Apoptosis and Bile Acids

In hepatocytes both pro- and anti-apoptotic actions of JNK signaling have been observed and appear to be stimulus-dependent [5, 25, 38, 81, 83, 89]. Lig- and-induced activation of the TNF- α receptor in hepatocytes is important in proliferative responses, while at the same time, many groups have argued that TNF- α-induced JNK signaling is toxic, par- ticularly when combined with oxidative stress [83].

Others have argued that TNF- α and ceramide/PKC zeta-induced JNK signaling plays a protective role, and that loss of both JNK and NF- κB induction pro- mote TNF- α toxicity [81], including that induced by chemotherapeutic agents [85, 86]. Activation of MEKK1, either by upstream stimuli or by caspase- dependent cleavage has been linked to many of the above observations [151].

Bile acids can activate the JNK pathway in hepa-

tocytes. DCA increases ceramide production in

(8)

hepatocytes and DCA- and TCA-induced JNK acti- vation is abolished in FAS-R –/– and acidic sphin- gomyelinase –/– mouse hepatocytes, whereas the ERBB1 inhibitor or ROS/RNS-quenching agent N- acetyl cysteine has no inhibitory effect on JNK ac- tivity [43, 44]. In FAS-R –/– cells, TNF- α can still activate the JNK pathway and generate ceramide, and DCA can still activate the ERK1/2 pathway, ar- guing that bile acids do not functionally activate the TNF- α receptor towards the JNK pathway and that ERBB1 signaling is intact in FAS-R –/– cells. This is of note because ceramide has been linked to altered Raf-1–RAS interactions in some cell systems. Treat- ment of FAS-R –/– hepatocytes with exogenous neu- tral or acidic sphingomyelinase does not activate the JNK pathway, suggesting ceramide generation is upstream of the initial bile acid target (possibly acidic sphingomyelinase) and the FAS-R.

The roles of each JNK isoform appear to be dif- ferent in hepatocyte survival responses. Activation of JNK2 enhances the apoptotic response of mouse hepatocytes exposed to DCA, whereas inhibition of JNK1 function by expression of dominant nega- tive JNK1 reduces the apoptotic response of mouse hepatocytes exposed to DCA [110]. Treatment of hepatocytes with Brefeldin A, which blocks the bile acid-induced translocation of FAS-R from intrac- ellular stores to the plasma membrane [128], does not block bile acid-induced JNK activation [43]. The above findings are of particular note as ligand-inde- pendent activation of the FAS-R by bile acids, and to a lesser extent TRAIL receptor family members, has been causally linked to bile acid-stimulated apopto- sis [35, 49, 82, 105]. Thus bile acids activation of the JNK pathway in hepatocytes, via the activation of the FAS-R at an intracellular site, presumably is in a dynamic regulatory balance for cell survival with FAS-R-stimulated activation of pro-caspases. Pre- sumably, the dependence of bile acid-induced JNK activation on ceramide generation argues that clus- tering of FAS-R molecules into raft-like structures is required for downstream signal transduction, since ceramide generation causes the clustering of death receptors into rafts, which activate JNK [16, 99].

19.3.3

JNK1/2 Signaling and Metabolism

JNK1/2 signaling in non-hepatic cells has gener- ally not been linked to altered metabolic processes in cells. However, in hepatocytes the rapid activa- tion of JNK/AP-1 signaling has been linked to the downregulation of cholesterol 7 alpha hydroxylase (CYP7A1), the rate-limiting enzyme in the neutral pathway of bile acid biosynthesis [44]. More recent

studies have suggested that prolonged activation of the sterol-regulated transcription factor FXR can promote the expression of the growth factor FGF-19, leading to activation of the FGF receptor 4 and downstream JNK pathway, thereby controlling CYP7A1 expression [50].

19.4

The p38 MAPK Pathway

The p38 MAPK pathway was originally described as a mammalian homolog to a yeast osmolarity sens- ing pathway [45]. It was soon discovered that many cellular stresses activated the p38 MAPK pathway, in a manner not dissimilar to that described for the JNK pathway. Rho family GTPases appear to play an important role as upstream activators of the p38 MAPK pathway and via several MAP3K enzymes, e.g., the PAK family, regulate the MAP2K enzymes MKK3 and MKK6 [46]. At least four isoforms of p38 MAPK exist, termed p38 α, β, γ and δ [73]. There are several protein kinases downstream of p38 MAPK enzymes that are activated following phosphoryla- tion by p38 isoforms, including p90

MAPKAPK2

and MSK1/2 [18]. P90

MAPKAPK2

phosphorylates and acti- vates HSP27, while MSK1/2 can phosphorylate/acti- vate transcription factors that regulate e.g., CREB.

Bile acid-induced activation of CREB in primary hepatocytes appears dependent on the ERK1/2 path- way, rather than p38 signaling [110].

The role of p38 MAPK signaling in cellular re- sponses is diverse, depending on the cell type and stimulus. For example, p38 MAPK signaling has been shown to promote cell death as well as to en- hance cell growth and survival [83, 161]. The ability of bile acids to regulate p38 MAPK activity in cells derived from the liver appears to be highly variable, with different groups reporting either no activation, weak activation or strong activation [40, 110, 111], which may be due to differences between primary and established cells. This is in contrast to the clas- sical MAPK–ERK1/2 pathway where bile acid-in- duced activation has been consistently observed by many groups.

19.4.1

Hepatocyte Proliferation and p38 Signaling

A positive role for p38 signaling in hepatocyte prolif-

eration was first noted in 1997, and subsequently by

a variety of groups [13, 129] (see Fig. 19.5). Ethanol-

induced impairment of liver regeneration has been

(9)

linked to both a reduction in stimulated p38 phos- phorylation and sustained p38 activity, which was correlated to high basal expression of p21 [13, 68].

The growth-promoting role of p38 may be stimulus specific as activation of G-protein-coupled recep- tors, including those for prostaglandin F(2alpha), vasopressin and norepinephrine, does not appear to stimulate hepatocyte growth via either the ERK1/2 or p38 pathways [94]. Expression of HBV X protein has been shown to activate the p38 pathway, which may be responsible for activation of STAT3 and in- creased cyclin D1 levels [137]: however, other studies have argued that STAT3 and p38 signaling can sup- press cyclin D1 levels in a differentiation-specific manner in liver-derived cells [88].

19.4.2

P38 Signaling, Apoptosis and Metabolism

Activation of the p38 pathway has been linked to the induction of apoptosis: TGF β-induced apoptosis has been shown to utilize p38 signaling as a pro-ap- optotic signal [80, 103]. However, hepatitis C virus core protein has also been shown to inhibit FAS-de- pendent p38 signaling and apoptosis [158]. In HuH7 cells, bile acid-induced apoptosis also recruits p38 as a pro-apoptotic stimulus that is inhibited by c- FLIP-L [40]; of note, c-FLIP-S expression is PI-3 kinase and ERK1/2 dependent in DCA-treated pri- mary rodent hepatocytes [105, 110]. Hepatocellular carcinoma cells have been noted to have lower levels of apoptosis and p38 activity [54]. P38 signaling has also been linked to choloretic efflux of bile acids, by regulating the distribution of the translocation to the plasma membrane of the bile salt export pump [71].

19.5

The MEK5-ERK5 "Big MAP Kinase" Pathway

The "big MAP kinase" pathway was first described in 1995 [164]. The term "big" derives from the fact that, whereas the molecular masses of ERK1/2 and JNK1/2 are 42/44 kDa and 46/54 kDa, respectively, ERK5 has a mass of ~90 kDa. The upstream activa- tors of ERK5, the MEK5 isoforms, have a similar molecular mass to other MAP2K molecules [31] and display different subcellular locations. The response of the MEK5-ERK5 pathway to growth factors such as EGF is similar to that of the MEK1/2-ERK1/2 pathway, including in many, but not all cell types, a dependency on Ras signaling [31, 58, 59]. Fur- thermore, MEK5 is also partially inhibited by the

previously described MEK1/2 inhibitors PD98059 and U0126. It appears that PD184352 is much more specific at lower concentrations (<10–20 µM) for MEK1/2 than MEK5 [91]. In this regard, we have found that PD184352 potentiated DCA-induced ap- optosis in primary hepatocytes to the same extent as U0126 and dominant negative MEK1 (S217A), but not dominant negative MEK1 (K97M), suggest- ing MEK5 does not signal to survival after DCA ex- posure (Fig. 19.2a) [105] (L. Qiao and P. Dent, un- published findings). The MAP3K enzymes recently shown to phosphorylate MEK5, MEKK2 and MEKK3 have been previously linked to signaling through the JNK pathway [133]. ERK5 has been proposed to phosphorylate and activate the transcription factors MEF2C, Sap 1a and also c-Myc. In a similar man- ner to the ERK1/2 pathway, the ERK5 cascade has been proposed to play a key role in growth factor- stimulated cell growth and cell survival processes (Fig. 19.1). The activation of MEK5-ERK5 in hepa- tocytes is unknown, although preliminary studies suggest it can be activated by EGF and conjugated/

unconjugated bile acids in primary hepatocytes (C.

Mitchell and P. Dent, unpublished observations).

Based on the fact that EGFR/Ras signaling can pro- mote MEK5-ERK5 activation, and bile acids also ac- tivate these molecules, it is probable that this path-

Fig. 19.5. Inhibition of p38 alpha function blunts hepatocyte DNA synthesis. Rats were subjected to either partial hepatecto- my or sham operation. Twenty-four hours after surgery, primary hepatocytes were isolated and plated in vitro. Hepatocytes were plated in the presence of vehicle (DMSO), the p38 inhibi- tor SB203580 (5 µM), its inactive analog SKF106978 (5 µM) or the MEK1/2/5 inhibitor PD98059 (50 µM). Cells were incubated with 10 µCi [3H]thymidine and DNA synthesis was determined 48 h after plating by [3H]thymidine incorporation into DNA [129]. As indicated by the solid lines, DNA synthesis is strongly inhibited by SB203580 but only weakly inhibited by the MEK1/2/5 inhibi- tor PD98059

(10)

way will be stimulated following exposure of cells to many growth factors.

19.6 Conclusions

Multiple MAPK pathways play critical roles in hepa- tocyte proliferation, survival and metabolism con- trol. Although much has been discovered using iso- lated hepatocytes in vitro, significantly more work needs to be performed to understand how MAPK pathway signaling within the intact liver is control- led.

Acknowledgments

P.D. was supported in part by National Institute of Health grants DK52875, CA72955, CA88906, Depart- ment of Defense grants BC98-0148 and BC02-0335, the Jim Valvano "V Foundation" and the Universal Leaf Corporation Professorship in Signal Transduc- tion Research.

Selected Reading

Sturgill TW, Ray LB. Muscle proteins related to microtubule as- sociated protein-2 are substrates for an insulin-stimulatable kinase. Biochem Biophys Res Commun 1986;134:565–571.

Sturgill TW, Ray LB, Erikson E, Maller JL. Insulin-stimulated MAP- 2 kinase phosphorylates and activates ribosomal protein S6 kinase II. Nature 1988;334:715–718.

Derijard B, Raingeaud J, Barrett T et al. Independent human MAP-kinase signal transduction pathways defined by MEK and MKK isoforms. Science 1995;267:682–685.

Qiao L, Studer E, Leach K et al. Deoxycholic acid (DCA) causes ligand-independent activation of epidermal growth factor receptor (EGFR) and FAS receptor in primary hepatocytes:

inhibition of EGFR/mitogen-activated protein kinase-signal- ing module enhances DCA-induced apoptosis. Mol Biol Cell 2001;12:2629–2645.

In Sturgill and Ray, the discovery of "classical" insu- lin-stimulated MAP kinase (now termed ERK2) was reported. In Sturgill et al., the ability of this MAP kinase to phosphorylate and activate ribosomal p90 S6 kinase was noted: these findings represent the first report of a cascade of insulin-stimulated pro- tein kinases, and ultimately led to the elucidation of the component parts of the "insulin black box."

Studies in Derijard et al. describe the discovery of MAP kinase family pathways parallel to the initial classical ERK/MAP kinase cascade. As an example of the complex interplay between MAP kinase path- ways in the liver, Qiao et al. present evidence as to how bile acids, via activation of receptor tyrosine kinases such as the EGF, HER2 and insulin recep- tors, and other receptors such as the FAS receptor, promote simultaneous activation of multiple MAP kinase pathways in hepatocytes.

References

1. Albrecht JH, Poon RY, Ahonen CL et al. Involvement of p21 and p27 in the regulation of CDK activity and cell cycle pro- gression in the regenerating liver. Oncogene 1998;16:2141–

2150.

2. Allan LA, Morrice N, Brady S et al. Inhibition of caspase-9 through phosphorylation at Thr 125 by ERK MAPK. Nat Cell Biol 2003;5:647–654.

3. Auer KL, Spector MS, Tombes RM et al. Alpha-adrener- gic inhibition of proliferation in HepG2 cells stably trans- fected with the alpha1B-adrenergic receptor through a p42MAPkinase/p21Cip1/WAF1-dependent pathway. FEBS Lett 1998;436:131–138.

4. Auer KL, Park JS, Seth P et al. Prolonged activation of the mitogen-activated protein kinase pathway promotes DNA synthesis in primary hepatocytes from p21Cip-1/WAF1-null mice, but not in hepatocytes from p16INK4a-null mice. Bio- chem J 1998;336:551–560.

5. Auer KL, Contessa J, Brenz-Verca S et al. The Ras/Rac1/Cdc42/

SEK/JNK/c-Jun cascade is a key pathway by which agonists stimulate DNA synthesis in primary cultures of rat hepato- cytes. Mol Biol Cell 1998;9:561–573.

6. Barreca A, Voci A, Minuto F et al. Effect of epidermal growth factor on insulin-like growth factor-I (IGF-I) and IGF-binding protein synthesis by adult rat hepatocytes. Mol Cell Endocri- nol 1992;84:119-126.

7. Benn J, Su F, Doria M, Schneider RJ. Hepatitis B virus HBx pro- tein induces transcription factor AP-1 by activation of extra- cellular signal-regulated and c-Jun N-terminal mitogen-acti- vated protein kinases. J Virol 1996;70:4978–4985.

8. Boucher MJ, Morisset J, Vachon PH et al. MEK/ERK signaling pathway regulates the expression of Bcl-2, Bcl-X(L), and Mcl- 1 and promotes survival of human pancreatic cancer cells. J Cell Biochem 2000;79:355–369.

9. Boulton TG, Cobb MH. Identification of multiple extracellular signal-regulated kinases (ERKs) with antipeptide antibodies.

Cell Regul 1991;2:357–371.

10. Buck M, Poli V, van der Geer P et al. Phosphorylation of rat serine 105 or mouse threonine 217 in C/EBP beta is required for hepatocyte proliferation induced by TGF alpha. Mol Cell 1999;4:1087–1092.

(11)

11. Cai H, Smola U, Wixler V et al. Role of diacylglycerol-regu- lated protein kinase C isotypes in growth factor activation of the Raf-1 protein kinase. Mol Cell Biol 1997;17:732–741.

12. Chadee DN, Yuasa T, Kyriakis JM. Direct activation of mi- togen-activated protein kinase kinase kinase MEKK1 by the Ste20p homologue GCK and the adapter protein TRAF2. Mol Cell Biol 2002;22:737–749.

13. Chen J, Ishac EJ, Dent P et al. Effects of ethanol on mitogen- activated protein kinase and stress-activated protein ki- nase cascades in normal and regenerating liver. Biochem J 1998;334:669–676.

14. Coles LC, Shaw PE. PAK1 primes MEK1 for phosphorylation by Raf-1 kinase during cross-cascade activation of the ERK pathway. Oncogene 2002;21:2236–2244.

15. Coutant A, Rescan C, Gilot D et al. PI3K-FRAP/mTOR path- way is critical for hepatocyte proliferation whereas MEK/

ERK supports both proliferation and survival. Hepatology 2002;36:1079–1088.

16. Cremesti A, Paris F, Grassme H et al. Ceramide enables fas to cap and kill. J Biol Chem 2001;276:23954–23961.

17. Davies SP, Reddy H, Caivano M, Cohen P. Specificity and mechanism of action of some commonly used protein ki- nase inhibitors. Biochem J 2000;351:95–105.

18. Deak MM, Clifton AD, Lucocq LM, Alessi DR. Mitogen- and stress-activated protein kinase-1 (MSK1) is directly activated by MAPK and SAPK2/p38, and may mediate activation of CREB. EMBO J 1998;17:4426–4441.

19. Dent P, Haser W, Haystead TA et al. Activation of mitogen- activated protein kinase kinase by v-Raf in NIH 3T3 cells and in vitro. Science 1992;257:1404–1407.

20. Dent P, Reardon DB, Morrison DK, Sturgill TW. Regulation of Raf-1 and Raf-1 mutants by Ras-dependent and Ras-inde- pendent mechanisms in vitro. Mol Cell Biol 1995;15:4125–

4135.

21. Dent P, Reardon DB, Park JS et al. Radiation-induced release of transforming growth factor alpha activates the epider- mal growth factor receptor and mitogen-activated protein kinase pathway in carcinoma cells, leading to increased pro- liferation and protection from radiation-induced cell death.

Mol Biol Cell 1999;10:2493–2506.

22. Dent P, Qiao L, Gilfor D et al. The regulation of tumor suppres- sor genes by oncogenes. Methods Mol Biol 2003;222:269–

292.

23. Derijard B, Hibi M, Wu IH et al. JNK1: a protein kinase stimu- lated by UV light and Ha-Ras that binds and phosphorylates the c-Jun activation domain. Cell 1994;76:1025–1037.

24. Dhillon AS, Pollock C, Steen H et al. Cyclic AMP-dependent kinase regulates Raf-1 kinase mainly by phosphorylation of serine 259. Mol Cell Biol 2002;22:3237–3246.

25. Diao J, Khine AA, Sarangi F et al. X Protein of hepatitis B virus inhibits Fas-mediated apoptosis and is associated with up-regulation of the SAPK/JNK pathway. J Biol Chem 2001;276:8328–8340.

26. Diehl AM, Yin M, Fleckenstein J et al. Tumor necrosis factor- alpha induces c-jun during the regenerative response to liver injury. Am J Physiol 1994;267:G552–G561.

27. Dixon M, Agius L, Yeaman SJ, Day CP. Inhibition of rat hepa- tocyte proliferation by transforming growth factor beta and glucagon is associated with inhibition of ERK2 and p70 S6 kinase. Hepatology 1999;29:1418–1424.

28. Eblen ST, Slack JK, Weber MJ, Catling AD. Rac-PAK signaling stimulates extracellular signal-regulated kinase (ERK) acti- vation by regulating formation of MEK1-ERK complexes. Mol Cell Biol 2002;22:6023–6033.

29. Eferl R, Sibilia M, Hilberg F et al. Functions of c-Jun in liver and heart development. J Cell Biol 1999;145:1049–1061.

30. Eldar-Finkelman H, Seger R, Vandenheede JR, Krebs EG. Inac- tivation of glycogen synthase kinase-3 by epidermal growth factor is mediated by mitogen-activated protein kinase/p90 ribosomal protein S6 kinase signaling pathway in NIH/3T3 cells. J Biol Chem 1995;270:987–990.

31. English JM, Vanderbilt CA, Xu S et al. Isolation of MEK5 and differential expression of alternatively spliced forms. J Biol Chem 1995;270:28897-28902.

32. Erhardt P, Schremser EJ, Cooper GM. B-Raf inhibits pro- grammed cell death downstream of cytochrome c release from mitochondria by activating the MEK/Erk pathway. Mol Cell Biol 1999;19:5308–5315.

33. Fang Y, Han SI, Mitchell C et al. Bile acids induce mitochon- drial ROS, which promote activation of receptor tyrosine kinases and signaling pathways in rat hepatocytes. Hepatol- ogy 2004;40:961.

34. Fassett JT, Tobolt D, Nelsen CJ et al. The role of collagen structure in mitogen stimulation of ERK, cyclin D1 expres- sion, and G1-S progression in rat hepatocytes. J Biol Chem 2003;278:31691–31700.

35. Faubion WA, Guicciardi ME, Miyoshi H et al. Toxic bile salts induce rodent hepatocyte apoptosis via direct activation of Fas. J Clin Invest 1999;103:137–145.

36. Frost JA, Xu S, Hutchison MR et al. Actions of Rho family small G proteins and p21-activated protein kinases on mitogen- activated protein kinase family members. Mol Cell Biol 1996;16:3707–3713.

37. Ganiatsas S, Kwee L, Fujiwara Y et al. SEK1 deficiency reveals mitogen-activated protein kinase cascade crossregulation and leads to abnormal hepatogenesis. Proc Natl Acad Sci USA 1998;95:6881–6886.

38. Gilot D, Loyer P, Corlu A et al. Liver protection from apoptosis requires both blockage of initiator caspase activities and in- hibition of ASK1/JNK pathway via glutathione S-transferase regulation. J Biol Chem 2002;277:49220–49229.

39. Giroux S, Tremblay M, Bernard D et al. Embryonic death of Mek1-deficient mice reveals a role for this kinase in angio- genesis in the labyrinthine region of the placenta. Curr Biol 1999;9:369–372.

40. Grambihler A, Higuchi H, Bronk SF, Gores GJ. cFLIP-L in- hibits p38 MAPK activation: an additional anti-apoptotic mechanism in bile acid-mediated apoptosis. J Biol Chem 2003;278:26831–26837.

41. Grant S, Qiao L, Dent P. Roles of ERBB family receptor tyrosine kinases, and downstream signaling pathways, in the control of cell growth and survival. Front Biosci 2002;7:376–389.

(12)

42. Guan KL, Figueroa C, Brtva TR et al. Negative regulation of the serine/threonine kinase B-Raf by Akt. J Biol Chem 2000;275:27354–27359.

43. Gupta S, Fang Y, Dent P, Hylemon PB. Bile acids activate the JNK pathway in hepatocytes via acid sphingomyeli- nase/ceramide/FAS-R-dependent mechanism J Biol Chem 2004;279:5821–5828.

44. Gupta S, Stravitz RT, Dent P, Hylemon PB. Down-regulation of cholesterol 7alpha-hydroxylase (CYP7A1) gene expression by bile acids in primary rat hepatocytes is mediated by the c- Jun N-terminal kinase pathway. J Biol Chem 2001;276:15816–

15822.

45. Han J, Lee JD, Bibbs L, Ulevitch RJ. A MAP kinase targeted by endotoxin and hyperosmolarity in mammalian cells. Science 1994;265:808–811.

46. Han J, Lee JD, Jiang Y et al. Characterization of the structure and function of a novel MAP kinase kinase (MKK6). J Biol Chem 1996;271:2886–2891.

47. Han SI, Fang Y, Gupta S et al. Bile acids enhance the activ- ity of the insulin receptor and glycogen synthase in primary rodent hepatocytes. Hepatology 2004;39:456–463.

48. Hibi M, Lin A, Smeal T et al. Identification of an oncoprotein- and UV-responsive protein kinase that binds and potentiates the c-Jun activation domain. Genes Dev 1993;7:2135–2148.

49. Higuchi H, Bronk SF, Taniai M et al. Cholestasis increases tumor necrosis factor-related apoptotis-inducing ligand (TRAIL)-R2/DR5 expression and sensitizes the liver to TRAIL- mediated cytotoxicity. J Pharmacol Exp Ther 2002;303:461–

467.

50. Holt JA, Luo G, Billin AN et al. Definition of a novel growth factor-dependent signal cascade for the suppression of bile acid biosynthesis. Genes Dev 2003;17:1581–1591.

51. Huynh HT, Nguyen TT, Chow PK et al. Over-expression of MEK-MAPK in hepatocellular carcinoma: its role in tumor progression and apoptosis. BMC Gastroenterol 2003;3:19.

52. Ilyin GP, Glaise D, Gilot D et al. Regulation and role of p21 and p27 cyclin-dependent kinase inhibitors during hepatocyte differentiation and growth. Am J Physiol Gastrointest Liver Physiol 2003;285:G115–G127.

53. Ito Y, Sasaki Y, Horimoto M et al. Activation of mitogen- activated protein kinases/extracellular signal-regulated kinases in human hepatocellular carcinoma. Hepatology 1998;27:951–958.

54 Iyoda K, Sasaki Y, Horimoto M et al. Involvement of the p38 mitogen-activated protein kinase cascade in hepatocellular carcinoma. Cancer 2003;97:3017–3026.

55. Jost M, Huggett TM, Kari C et al. Epidermal growth factor receptor-dependent control of keratinocyte survival and Bcl-xL expression through a MEK-dependent pathway. J Biol Chem 2001;276:6320–6326.

56. Kang S, Song J, Kang H et al. Insulin can block apoptosis by decreasing oxidative stress via phosphatidylinositol 3-ki- nase- and extracellular signal-regulated protein kinase-de- pendent signaling pathways in HepG2 cells. Eur J Endocrinol 2003;148:147–155.

57. Kar S, Adachi T, Carr BI. EGFR-independent activation of ERK1/2 mediates growth inhibition by a PTPase antagoniz- ing K-vitamin analog. J Cell Physiol 2002;190:356–364.

58. Kato Y, Kravchenko VV, Tapping RI et al. BMK1/ERK5 regulates serum-induced early gene expression through transcription factor MEF2C. EMBO J 1997;16:7054–7066.

59. Kato Y, Tapping RI, Huang S et al. Bmk1/Erk5 is required for cell proliferation induced by epidermal growth factor. Na- ture 1998;395:713–716.

60. Kawamata Y, Fujii R, Hosoya M et al. A G protein-coupled re- ceptor responsive to bile acids. J Biol Chem 2003;278:9435–

9440.

61. Khamzina L, Gruppuso PA, Wands JR. Insulin signaling through insulin receptor substrate 1 and 2 in normal liver development. Gastroenterology 2003;125:572–585.

62. Kim GY, Mercer SE, Ewton DZ et al. The stress-activated pro- tein kinases p38 alpha and JNK1 stabilize p21(Cip1) by phos- phorylation. J Biol Chem 2002;277:29792–29802.

63. Kimura ET, Nikiforova MN, Zhu Z et al. High prevalence of BRAF mutations in thyroid cancer: genetic evidence for constitutive activation of the RET/PTC-RAS-BRAF signal- ing pathway in papillary thyroid carcinoma. Cancer Res 2003;63:1454–1457.

64. King AJ, Wireman RS, Hamilton M, Marshall MS. Phosphor- ylation site specificity of the Pak-mediated regulation of Raf- 1 and cooperativity with Src. FEBS Lett 2001;497:6–14.

65. Kirillova I, Chaisson M, Fausto N. Tumor necrosis factor in- duces DNA replication in hepatic cells through nuclear fac- tor kappaB activation. Cell Growth Differ 1999;10:819–828.

66. Klein NP, Schneider RJ. Activation of Src family kinases by hepatitis B virus HBx protein and coupled signaling to Ras.

Mol Cell Biol 1997;17:6427–6436.

67. Kobayashi S, Matsushita K, Saigo K et al. P21WAF1/CIP1 mes- senger RNA expression in hepatitis B, C virus-infected hu- man hepatocellular carcinoma tissues. Cancer 2001;91:2096–

2103.

68. Koteish A, Yang S, Lin H et al. Ethanol induces redox-sensitive cell-cycle inhibitors and inhibits liver regeneration after par- tial hepatectomy. Alcohol Clin Exp Res 2002;26:1710–1718.

69. Kovalovich K, Li W, Deangelis R et al. IL-6 protects against Fas-mediated death by establishing a critical level of antia- poptotic hepatic proteins FLIP, Bcl-2 and Bcl-xL. J Biol Chem 2001;276:26605–26613.

70. Kurz AK, Block C, Graf D et al. Phosphoinositide 3 kinase- dependent Ras activation by tauroursodesoxycholate in rat liver. Biochem J 2000;350:207–213.

71. Kurz AK, Graf D, Schmitt M et al. Tauroursodesoxycholate-in- duced choleresis involves p38(MAPK) activation and trans- location of the bile salt export pump in rats. Gastroenterol- ogy 2001;121:407–419.

72. Kwon YH, Jovanovic A, Serfas MS, Tyner AL. The Cdk inhibitor p21 is required for necrosis, but it inhibits apoptosis follow- ing toxin-induced liver injury. J Biol Chem 2003;278:30348–

30355.

(13)

73. Kyriakis JM, Avruch J. Mammalian mitogen-activated pro- tein kinase signal transduction pathways activated by stress and inflammation. Physiol Rev 2001;81:807–869.

74. Lange-Carter CA, Pleiman CM, Gardner AM et al. A diver- gence in the MAP kinase regulatory network defined by MEK kinase and Raf. Science 1993;260:315–319.

75. Leach JK, Qiao L, Fang Y et al. Regulation of p21 and p27 ex- pression by the hepatitis B virus X protein and the alternate initiation site X proteins, AUG2 and AUG3. J Gastroenterol Hepatol 2003;18:376–385.

76. Lee S, Tarn C, Wang WH et al. Hepatitis B virus X protein dif- ferentially regulates cell cycle progression in X-transforming versus non-transforming hepatocyte (AML12) cell lines. J Biol Chem 2002;277:8730–8740.

77. Leu JI, Crissey MA, Craig LE, Taub R. Impaired hepatocyte DNA synthetic response posthepatectomy in insulin-like growth factor-binding protein 1-deficient mice with defects in C/EBP beta and mitogen-activated protein kinase/ex- tracellular signal-regulated kinase regulation. Mol Cell Biol 2003;23:1251–1259.

78. Leu CM, Chang C, Hu C. Epidermal growth factor (EGF) sup- presses staurosporine-induced apoptosis by inducing mcl-1 via the mitogen-activated protein kinase pathway. Onco- gene 2000;19:1665–1675.

79. Ley R, Balmanno K, Hadfield K et al. Activation of the ERK1/2 signaling pathway promotes phosphorylation and protea- some-dependent degradation of the BH3-only protein, Bim.

J Biol Chem 2003;278:18811–18816.

80. Liao JH, Chen JS, Chai MQ et al. The involvement of p38 MAPK in transforming growth factor beta1-induced apopto- sis in murine hepatocytes. Cell Res 2001;11:89–94.

81. Liedtke C, Plümpe J, Kubicka S et al. Jun kinase modulates tumor necrosis factor-dependent apoptosis in liver cells.

Hepatology 2002;36:315–325.

82. Lieser MJ, Park J, Natori S et al. Cholestasis confers resistance to the rat liver mitochondrial permeability transition. Gas- troenterology 1998;115:693–701.

83. Liu H, Lo CR, Czaja MJ. NF- B inhibition sensitizes hepato- cytes to TNF-induced apoptosis through a sustained activa- tion of JNK and c-Jun. Hepatology 2002;35:772–778.

84. Li N, Batzer A, Daly R et al. Guanine-nucleotide-releasing fac- tor hSos1 binds to Grb2 and links receptor tyrosine kinases to Ras signaling. Nature 1993;363:85–88.

85. Liu B, Fang M, Lu Y et al. Involvement of JNK-mediated path- way in EGF-mediated protection against paclitaxel-induced apoptosis in SiHa human cervical cancer cells. Br J Cancer 2001;85:303–311.

86. Liu H, Jones BE, Bradham C, Czaja MJ. Increased cytochrome P-450 2E1 expression sensitizes hepatocytes to c-Jun-medi- ated cell death from TNF-alpha. J Physiol Gastrointest Liver Physiol 2002;282:G257–G266.

87. Luedde T, Rodriguez ME, Tacke F et al. p18(INK4c) collabo- rates with other CDK-inhibitory proteins in the regenerating liver. Hepatology 2003;37:833–841.

88. Matsui T, Kinoshita T, Hirano T et al. STAT3 down-regulates the expression of cyclin D during liver development. J Biol Chem 2002;277:36167–36173.

89. Matsumaru K, Ji C, Kaplowitz N. Mechanisms for sensitization to TNF-induced apoptosis by acute glutathione depletion in murine hepatocytes. Hepatology 2003;37:1425–1434.

90. Mitsui H, Takuwa N, Maruyama T et al. The MEK1-ERK map kinase pathway and the PI 3-kinase-Akt pathway independ- ently mediate anti-apoptotic signals in HepG2 liver cancer cells. Int J Cancer 2001;92:55–62.

91. Mody N, Leitch J, Armstrong C et al. Effects of MAP kinase cascade inhibitors on the MKK5/ERK5 pathway. FEBS Lett 2001;502:21–24.

92. Moodie SA, Willumsen BM, Weber MJ, Wolfman A. Com- plexes of Ras.GTP with Raf-1 and mitogen-activated protein kinase kinase. Science 1993;260:1658–1661.

93. Nagaki M, Sugiyama A, Naiki T et al. Control of cyclins, cy- clin-dependent kinase inhibitors, p21 and p27, and cell cy- cle progression in rat hepatocytes by extracellular matrix. J Hepatol 2000;32:488–496.

94. Nilssen LS, Hege Thoresen G, Christoffersen T, Sandnes D.

Differential role of MAP kinases in stimulation of hepato- cyte growth by EGF and G-protein-coupled receptor ago- nists. Biochem Biophys Res Commun 2002;291:588–592.

95. Nishina H, Vaz C, Billia P et al. Defective liver formation and liver cell apoptosis in mice lacking the stress signaling ki- nase SEK1/MKK4. Development 1999;126:505–516.

96. Noguchi K, Kitanaka C, Yamana H et al. Regulation of c-Myc through phosphorylation at Ser-62 and Ser-71 by c-Jun N- terminal kinase. J Biol Chem 1999;274:32580–32587.

97. Olivier JP, Raabe T, Henkemeyer M et al. A Drosophila SH2- SH3 adaptor protein implicated in coupling the sevenless tyrosine kinase to an activator of Ras guanine nucleotide exchange, Sos. Cell 1993;73:179–191.

98. Pardo OE, Arcaro A, Salerno G et al. Fibroblast growth fac- tor-2 induces translational regulation of Bcl-XL and Bcl-2 via a MEK-dependent pathway: correlation with resistance to etoposide-induced apoptosis. J Biol Chem 2002;277:12040–

12046.

99. Paris F, Grassme H, Cremesti A et al. Natural ceramide re- verses Fas resistance of acid sphingomyelinase(–/–) hepa- tocytes. J Biol Chem 2001;276:8297–8305.

100. Park JS, Boyer S, Mitchell K et al. Expression of human pap- illoma virus E7 protein causes apoptosis and inhibits DNA synthesis in primary hepatocytes via increased expression of p21(Cip-1/WAF1/MDA6). J Biol Chem 2000;275:18–28.

101. Park JS, Qiao L, Gilfor D et al. A role for both Ets and C/EBP transcription factors and mRNA stabilization in the MAPK- dependent increase in p21 (Cip-1/WAF1/mda6) protein levels in primary hepatocytes. Mol Biol Cell 2000;11:2915–

2932.

102. Park US, Park SK, Lee YI et al. Hepatitis B virus-X protein upregulates the expression of p21waf1/cip1 and prolongs G1→S transition via a p53-independent pathway in human hepatoma cells. Oncogene 2000;19:3384–3394.

(14)

103. Park HJ, Kim BC, Kim SJ, Choi KS. Role of MAP kinases and their cross-talk in TGF-beta1-induced apoptosis in FaO rat hepatoma cell line. Hepatology 2002;35:1360–1371.

104. Pritchard CA, Samuels ML, Bosch E, McMahon M. Condition- ally oncogenic forms of the A-Raf and B-Raf protein kinases display different biological and biochemical properties in NIH 3T3 cells. Mol Cell Biol 1995;15:6430–6442.

105. Qiao L, Studer E, Leach K et al. Deoxycholic acid (DCA) causes ligand-independent activation of epidermal growth factor receptor (EGFR) and FAS receptor in primary hepa- tocytes: inhibition of EGFR/mitogen-activated protein ki- nase-signaling module enhances DCA-induced apoptosis.

Mol Biol Cell 2001;12:2629–2645.

106. Qiao L, Leach K, McKinstry R et al. Hepatitis B virus X pro- tein increases expression of p21(Cip-1/WAF1/MDA6) and p27(Kip-1) in primary mouse hepatocytes, leading to re- duced cell cycle progression. Hepatology 2001;34:906–

917.

107. Qiao D, Stratagouleas ED, Martinez JD. Activation and role of mitogen-activated protein kinases in deoxycholic acid- induced apoptosis. Carcinogenesis 2001;22:35–41.

108. Qiao L, Yacoub A, Studer E et al. Inhibition of the MAPK and PI3 K pathways enhances UDCA-induced apoptosis in pri- mary rodent hepatocytes. Hepatology 2002;35:779–789.

109. Qiao L, McKinstry R, Gupta S et al. Cyclin kinase inhibitor p21 potentiates bile acid-induced apoptosis in hepatocytes that is dependent on p53. Hepatology 2002;36:39–48.

110. Qiao L, Han SI, Fang Y et al. Bile acid regulation of C/EBP- beta, CREB, and c-Jun function, via the extracellular signal- regulated kinase and c-Jun NH2-terminal kinase pathways, modulates the apoptotic response of hepatocytes. Mol Cell Biol 2003;23:3052–3066.

111. Rao YP, Studer EJ, Stravitz RT et al. Activation of the Raf- 1/MEK/ERK cascade by bile acids occurs via the epidermal growth factor receptor in primary rat hepatocytes. Hepa- tology 2002;35:307–314.

112. Reinehr R, Graf D, Haussinger D. Bile salt-induced hepato- cyte apoptosis involves epidermal growth factor receptor- dependent CD95 tyrosine phosphorylation. Gastroenterol- ogy 2003;125:839–853.

113. Rescan C, Coutant A, Talarmin H et al. Mechanism in the se- quential control of cell morphology and S phase entry by epidermal growth factor involves distinct MEK/ERK activa- tions. Mol Biol Cell 2001;12:725–738.

114. Reuther GW, Der CJ. The RAS branch of small GTPases: RAS family members don't fall far from the tree. Curr Opin Cell Biol 2000;12:157–165.

115. Ries S, Biederer C, Woods D et al. Opposing effects of Ras on p53: transcriptional activation of mdm2 and induction of p19ARF. Cell 2000;103:321–330.

116. Rodrigues CM, Fan G, Wong PY et al. Ursodeoxycholic acid may inhibit deoxycholic acid-induced apoptosis by modu- lating mitochondrial transmembrane potential and reac- tive oxygen species production. Mol Med 1998;4:165–178.

117. Rodrigues CM, Fan G, Ma X et al. A novel role for ursodeoxy- cholic acid in inhibiting apoptosis by modulating mitochon-

drial membrane perturbation. J Clin Invest 1998;101:2790–

2799.

118. Rosenfeld ME, Prichard L, Shiojiri N, Fausto N. Prevention of hepatic apoptosis and embryonic lethality in RelA/TNFR-1 double knockout mice. Am J Pathol 2000;156:997–1007.

119. Rust C, Karnitz LM, Paya CV et al. The bile acid tauroche- nodeoxycholate activates a phosphatidylinositol 3-ki- nase-dependent survival signaling cascade. J Biol Chem 2000;275:20210–20216.

120. Sajan MP, Bandyopadhyay G, Kanoh Y et al. Sorbitol acti- vates atypical protein kinase C and GLUT4 glucose trans- porter translocation/glucose transport through proline-rich tyrosine kinase-2, the extracellular signal-regulated kinase pathway and phospholipase D. Biochem J 2002;362:665–

674.

121. Scheid MP, Schubert KM, Duronio V. Regulation of bad phos- phorylation and association with Bcl-x(L) by the MAPK/Erk kinase. J Biol Chem 1999;274:31108–31113.

122. Shaw M, Cohen P. Role of protein kinase B and the MAP ki- nase cascade in mediating the EGF-dependent inhibition of glycogen synthase kinase 3 in Swiss 3T3 cells. FEBS Lett 1999;461:120–124.

123. Schlesinger TK, Fanger GR, Yujiri T, Johnson GL. The TAO of MEKK. Front Biosci 1998;3:D1181–D1186.

124. Schlesinger TK, Bonvin C, Jarpe MB et al. Apoptosis stimu- lated by the 91-kDa caspase cleavage MEKK1 fragment requires translocation to soluble cellular compartments. J Biol Chem 2002;277:10283–10291.

125. Schwabe RF, Bradham CA, Uehara T et al. c-Jun-N-terminal kinase drives cyclin D1 expression and proliferation during liver regeneration. Hepatology 2003;37:824–832.

126. Selzner N, Selzner M, Odermatt B et al. ICAM-1 triggers liver regeneration through leukocyte recruitment and Kupffer cell-dependent release of TNF-alpha/IL-6 in mice. Gastro- enterology 2003;124:692–700.

127. Skarpen E, Lindeman B, Thoresen GH et al. Impaired nu- clear accumulation and shortened phosphorylation of ERK after growth factor stimulation in cultured hepatocytes from rats exposed to 2-acetylaminofluorene. Mol Carcinog 2000;28:84–96.

128. Sodeman T, Bronk SF, Roberts PJ et al. Bile salts mediate hepatocyte apoptosis by increasing cell surface traffick- ing of Fas. Am J Physiol Gastrointest Liver Physiol 2000;278:

G992–G999.

129. Spector MS, Auer KL, Jarvis WD et al. Differential regula- tion of the mitogen-activated protein and stress-activated protein kinase cascades by adrenergic agonists in quies- cent and regenerating adult rat hepatocytes. Mol Cell Biol 1997;17:3556–3565.

130. Squires MS, Nixon PM, Cook SJ. Cell-cycle arrest by PD184352 requires inhibition of extracellular signal-regulated kinases (ERK) 1/2 but not ERK5/BMK1. Biochem J 2002;366:673–

680.

131. Sturgill TW, Ray LB. Muscle proteins related to microtubule associated protein-2 are substrates for an insulin-stimulat-

Riferimenti

Documenti correlati

Su questo punto Regione Lombardia intende avviare un’iniziativa con il Governo per ottenere alcune modifiche legislative: le dighe non siano beni dei concessionari, bensì beni

We then build a phenomenological model, consisting of a power law (to account for the con- tamination of NGC 5765A in the NuSTAR spectrum 4 ), a soft diffuse X-ray emission

The results of simulations obtained with an amplifier based on the Doherty technique (6) made it possible to check that this technique allows to maintain the Power Added

Liguria ha iniziato a partecipare alla gestione della rete sentieristica, con diverse iniziative, tra cui il progetto Alta Via dei Monti Liguri (AVML) e il

To conclude, mechanical AGN feedback tightly self-regulated via CCA is able to quench the cooling rates by 20 fold while pre- serving the cool-core structure (the mass sink problem);

Centinaia di espositori, attentamente selezionati, riempiono per tre giorni il centro storico di Cuneo, invadendo di profumi, sapori e colori le principali piazze della parte

Thus as shown in Table 2B, four proteins (not significantly induced in the 8% ethanol cultures), identified as clpL2 protein (spot 428), heat.. Taking into account the results